Second Edition Evolution and Geological Significance of Larger Benthic Foraminifera  is a unique, comprehensive reference work on the larger benthic foraminifera. E vo l u t i o n this second edition is substantially revised, including extensive re-analysis of  the most recent work on cenozoic forms. it provides documentation of the biostratigraphic ranges and paleoecological significance of the larger foraminifera, which is essential for understanding many major oil-bearing sedimentary basins. in addition, it offers a paleogeographic interpretation of a n D G E o l o G i c a l the shallow marine late Paleozoic to cenozoic world. Marcelle K. BouDagher-Fadel collects and significantly adds to the information already published on the larger benthic foraminifera. new research in the Far East, the Middle East, South africa, tibet and the americas has provided fresh insights into the evolution and paleographic S i G n i F i c a n c E significance of these vital reef-forming forms. With the aid of new and precise biostratigraphic dating, she presents revised phylogenies and ranges of the larger foraminifera. the book is illustrated throughout, with examples of different families and groups at the generic levels. Key species are discussed and their biostratigraphic ranges are depicted in comparative charts (available o F l a r G E r B E n t h i c separately online). PROF MaRcelle K. BOuDagheR-FaDel is a Professorial research Fellow in the office of the vice-Provost (research) at ucl. She graduated with a BSc from the lebanese university and has an MSc and PhD from ucl. She has an extensive publication record, having written three major books and F o r a M i n i F E r a over 130 papers. She is an established consultant with several oil companies, lectures widely, and supervises PhD students from around the world. Marcelle K. BouDagher-Fadel FRONT AND BACK COVER IMAGES: Marcelle K. BouDagher-Fadel Cover design: Rawshock design www.ucl.ac.uk/ucl-press Evolution anD GEoloGical SiGniFicancE oF larGEr MarcEllE K. BouDaGhEr-FaDEl BEnthic ForaMiniFEra EVOLUTION AND GEOLOGICAL SIGNIFICANCE OF LARGER BENTHIC FORAMINIFERA EVOLUTION AND GEOLOGICAL SIGNIFICANCE OF LARGER BENTHIC FORAMINIFERA SECOND EDITION MARCELLE K. BOUDAGHER- FADEL This Edition published in 2018 by UCL Press University College London Gower Street London WC1E 6BT First edition published in 2008 by Elsevier. Available to download free: www.ucl.ac.uk/u cl- press Text © Marcelle K. BouDagher- Fadel, 2018 Images © Marcelle K. BouDagher- Fadel and copyright owners named in the captions, 2018 Marcelle K. BouDagher- Fadel has asserted her right under the Copyright, Designs and Patents Act 1988 to be identified as the author of this work. A CIP catalogue record for this book is available from The British Library. This book is published under a Creative Commons 4.0 International license (CC BY 4.0).This license allows you to share, copy, distribute and transmit the work; to adapt the work and to make commercial use of the work providing attribution is made to the authors (but not in any way that suggests that they endorse you or your use of the work). Attribution should include the following information: BouDagher- Fadel M.K. 2018. Evolution and Geological Significance of Larger Benthic Foraminifera, Second edition. London, UCL Press. DOI: https:// doi.org/ 10.14324/ 111.9781911576938 Further details about Creative Commons licenses are available at http:// creativecommons.org/ licenses/ ISBN: 978- 1- 911576- 95- 2 (Hbk.) ISBN: 978- 1- 911576- 94- 5 (Pbk.) ISBN: 978-1 - 911576- 93- 8 (PDF) ISBN: 978- 1-9 11576- 96-9  (epub) ISBN: 978- 1- 911576- 97-6  (mobi) ISBN: 978- 1- 911576- 98-3  (html) DOI: https:// doi.org/ 10.14324/ 111.9781911576938 Acknowledgements Following the success of making the second edition of Biostratigraphic and Geological Significance of Planktonic Foraminifera freely available in open access format via UCL Press, I decided to publish a second edition of this book, in the same way. This second edition contains extensive revisions, additional figures, and a significant update to a large number of orders and families of the Larger Benthic Foraminifera. During the course of writing this book, I have been helped by numerous friends and colleagues. I would like to thank Prof Pamela Hallock Muller, University of South Florida, for carefully editing and reviewing Chapter 1, and Prof Vladimir Davydov of Florida International University, for reviewing early drafts of Chapter 2. I am grateful to the late Prof Lucas Hottinger, and to Dr G. Wyn Hughes for carefully reviewing vari- ous sections of this book. I would like to thank Ebrahim Mohammadi, Iran University for carefully reading the first edition and Prof Ahmad Aftab for sending me informa- tion on Pakistan localities. I am also grateful to Prof Felix Schlagintweit and Prof Ercan Özcan, Department of Geology, Maslak, for allowing me to publish the original illustrations of some of their new genera and species, Mr Nguyen Van Sang Vova for access to his Vietnam material, the South East Asia Group (SEA), Royal Holloway University London, for access to their Indonesian material, Gyongyver Jennifer Fischer and Pascal Kindler, University of Geneva for access to their Mayaguana Bank (SE Bahamas) material, and to Prof Hu Xiumian for access to his Cretaceous and Paleogene Tibetan material. As with the first edition, in creating this edition I have been greatly aided and sup- ported by my dear colleague and friend Prof David Price. I would also like to acknowl- edge the help of Chris Penfold from UCL Press who has been invaluable in the creation of this open access work. There are many photographs and illustrations in this book. Most are original, but some are reproduced from standard sources. I have tried to contact or reference all potential copyright holders. If I have overlooked any or been inaccurate in any acknowledgement, I apologise unreservedly and I will ensure that suitable corrections are made in subsequent editions. The charts mentioned in this book, are freely available separately online at https://doi.org/10.14324/111.9781911576938. In conclusion, I should repeat here the acknowledgement from the first edition (BouDagher-Fadel, 2008) as in the course of writing that edition, I was helped by numerous other friends and colleagues: I would like to thank Prof Ron Blakey for his permission to use his exquisite palaeo- geographic maps as illustrations in my book. Many more of these splendid maps can be found on his website http://jan.ucc.nau.edu/~rcb7/. Prof Rudolf Röttger has been most supportive, and has given me photographs of living larger foraminifera, and corrected Chapter 1 of my book. I also relied heavily vi Acknowledgements on his work in my discussions of the biology of the larger foraminifera as presented in Chapter 1. I would like to refer readers to Prof Lukas Hottinger’s outstanding web pages “Illustrated glossary of terms used in foraminiferal research” which can be found at http://paleopolis.rediris.es/cg/CG2006_M02/4_droite.htm. Some of these illustrations are reproduced in this book, courtesy of Prof Hottinger. I would like to thank Prof McMillan for access to his South African collection of larger foraminifera, Dr Michelle Ferrandini, Université de Corse, for access to her Corsican collections and Prof K. Matsumaru for access to some of his original material. I would also like to thank the Natural History Museum, London for giving me access to their excellent collection, which includes type species of many early work- ers. I would like to thank all scientists who contributed to this collection and thus to my book. My gratitude is also expressed to the Senckenberg-Forschungsinstitut und Naturmuseum, Germany for their Permian collection and UCL Geological Sciences, Micropalaeontology unit collections. I am particularly grateful for the assistance of Mr Jim Davy, UCL, and in Mr Clive Jones, NHM. Mr Jones was very helpful in locat- ing specimens and his methods of filing and storing the NHM collection were so very useful. Finally, I am especially grateful for the careful editing and reviewing carried out by Prof Alan Lord (of the Senckenberg-Forschungsinstitut und Naturmuseum, Germany) and Prof David Price (UCL). Prof Price’ s advice throughout the book, and our use- ful discussions on the causes of extinctions gave me many ideas on the relationship between sensitive, small, living organisms, such as the larger foraminifera, and large scale geological processes. I also thank him for helping me to look into the wider pro- cesses involved in evolution and for his encouragement. Marcelle BouDagher- Fadel London September 2017 newgenprepdf Contents 1. Biology and Evolutionary History of Larger Benthic Foraminifera 1 2. The Palaeozoic Larger Benthic Foraminifera. 45 3. The Mesozoic Larger Benthic Foraminifera: The Triassic 161 4. The Mesozoic Larger Benthic Foraminifera: The Jurassic 203 5. The Mesozoic Larger Benthic Foraminifera: The Cretaceous 285 6. The Cenozoic Larger Benthic Foraminifera: The Paleogene 387 7. The Cenozoic Larger Benthic Foraminifera: The Neogene 543 References 639 Subject Index 667 1 Chapter 1 Biology and Evolutionary History of Larger Benthic Foraminifera 1.1 Biological Classification of Foraminifera 1.1.1 Introduction Foraminifera are unicellular eukaryotes characterized by streaming granular ecto- plasm usually supported by an endoskeleton or “test” made of various materials. They are considered to fall within the phylum Retaria, which in turn is within the infrakingdom Rhizaria (Ruggiero et al., 2015). Their cellular cytoplasm is organised into a complex structure by internal membranes, and contains a nucleus (Plate 1.1, Figs. 1– 2), mitochondria, chloroplasts (when present) and Golgi bodies (Plate 1.1, Figs. 3–5 ; Plate 1.2). In foraminifera, the cytoplasm is subdivided into the endo- plasm, in which the nucleus (or nuclei, as many foraminifera are multinucleate) and other organelles are concentrated, and ectoplasm, which contains microtubules and mitochondria (Hemleben et al., 1977; Anderson et al., 1979; Alexander, 1985). Foraminifera are characterised by specialized pseudopodia (temporary organic pro- jections) known as granuloreticulopodia (also called rhizopodia), which are thread- like, granular, branched and anastomosing filaments that emerge from the cell body (Fig. 1.1). The unique ability of the foraminiferal ectoplasm to assemble and dis- assemble microtubules allows them rapidly to extend or retract their rhizopodia (Bowser and Travis, 2002). The functions of the rhizopodia include movement, feed- ing, and construction of the test. Both living and fossil foraminifera come in a wide variety of shapes and sizes. Academically, the study of their preserved tests is referred to as micropalaeontology, and although their typical size is sub- millimetric, they have occurred in the geologi- cal past with sizes up to ~150mm. In addition, they occur in many different environ- ments, from freshwater to the deep sea, and from near surface to the ocean floor. Their remains are extremely abundant in most marine sediments and they live in nearly all marine to brackish habitats (Fig. 1.2). Foraminifera that dwell in freshwater do not produce tests (Pawlowski et al., 2003), however most marine foraminiferal species grow an elaborate test or endoskeleton made of a series of chambers (Fig. 1.3). These single- celled organisms have inhabited the oceans for more than 500 million years. The complexity of their fossilised test structures (and their evolution in time) is the basis of their geological usefulness. The earliest known foraminifera, mostly forms that had an organic wall or produced a test by agglutinating particles within an organic or mineralized matrix, appeared in the Cambrian, and were common in the 2 Evolution and Geological Significance of Larger Benthic Foraminifera Fig. 1.1. Larger foraminifera Heterostegina depressa with thread- like, granular, branched and anastomos- ing filaments that emerge from the cell body (courtesy of Prof Röttger). Early Paleozoic (Platon et al., 2001). Forms with calcareous tests appeared by the Early Carboniferous, becoming diverse and abundant, with the evolutionary development of taxa with relatively large and complicated test architecture by the Late Paleozoic. Their long, diverse and well- documented evolutionary record makes Foraminifera of outstanding value in zonal stratigraphy, and in paleoenvironmental, palaeobiological and palaeoceanographic interpretation and analysis. Fossil and living foraminifera have been known and studied for centuries. They were noted by Herodotus (in his Histories written in the 5th century BC) as occurring in the limestone of the Egyptian pyramids, which in fact contain fossils of the larger ben- thic Foraminifera Nummulites. The name Foraminifera derives from the apertures and the “foramen” connecting successive chambers seen in their tests. The test surfaces of many foraminiferal species are covered with microscopic holes (foramen), normally visible at about x40 magnification (Fig. 1.4). Among the earliest workers who described and drew foraminiferal tests were Anthony van Leeuwenhoek in 1600, and Robert Hooke in 1665, but an accurate description of foraminiferal architecture was not given until the 19th century (Carpenter et al., 1862). Biology and Evolutionary History of Larger Benthic Foraminifera 3 Calcareous assemblage Stenohaline benthic Rotaliida ra Euryhaline simple complex inif e Rotaliida for am e.g. Elphidiidae ate d Planktonic tin Rotaliida glu Textulariidae Ag Ataxophragmiidae Miliolida Verneulinidae simple complex Globigerinidae s Lituolidae Cl i n- ceaO ing s fac che gro ve ea Man rift b p dal swa m ito r L MIXED REEF/ Laggon Reef top TERRESTRIAL SEDIMENTS BACK REEF/LAGOON REEF BIOHERMS FORE-REEF BASIN FORE-REEF SEDIMENTS Fig. 1.2. The ecological distribution of foraminifera. Aperture Proloculus Planispiral coiling a Chamber embrace evolute involute Trochospiral coiling a u a u a Miliolid coiling ventral dorsal Fig. 1.3. The different shapes of foraminiferal test; a = axis of the test; u = umbilicus. all f w Re e 4 Evolution and Geological Significance of Larger Benthic Foraminifera Fig. 1.4. An enlargement of the surface of a Heterostegina shell showing two types of holes. 1) the many small pores which are characteristic of all foraminifera. They do not form open connections between the test lumen and the sea water, but are closed by a membrane. Only small molecules like nutrition salts may pene- trate, which are important for the nutrition of the algal endosymbionts. 2) the larger openings on the lateral test surface are openings of the canal system of the chamber walls and chamberlet walls (shown in Fig. 1.17) with the outside world. In Heterostegina depressa, and other nummulitids, the protoplasm emerges through these openings and forms a thin veil covering the test surface in living specimens, which is also responsible for the secretion of the elastic inanimate protective sheath with radiating processes that covers the test, attaching it to the algal or rock surface. This function is described and illustrated by Röttger (1983). The apertures in the last chamber are masked in Heterostegina. The first attempts to taxonomically classify Foraminifera placed them within the genus Nautilus, a member of the phylum Mollusca. In 1781, Spengler was among the first to note that foraminiferal chambers are in fact divided by septa. In 1826, d’Orbigny, having made the same observation, named the group Foraminifères. In 1835, Foraminifera were recognised by Dujardin as protozoa, and shortly afterwards d’Orbigny produced the first classification of foraminifera, which was based on test morphology. Modern workers normally use the structure and composition of the test wall as a basis of primary classification, and this approach will be followed in this book. Despite the diversity and usefulness of the foraminifera, the phylogenetic relation- ship of Foraminifera to other eukaryotes has only recently emerged. Early genetic work on the origin of the Foraminifera postulated that the foraminiferal taxa are a divergent “alveolate” lineage, within the major eukaryotic radiation (Wray et al., 1995; Baldauf, 2003). Subsequently, many researchers have tried to determine the origin of the foramin- ifera, but molecular data from Foraminifera generated conflicting conclusions. Molecular phylogenetic trees have assigned most of the characterised eukaryotes to one of eight major groups. Baldauf (2003) tried to resolve the relationships among these groups to find “the deep roots of the eukaryotes”. He placed them in the “Cercozoa” group. Cercozoans are amoebae, with filose pseudopodia, often living within tests, some of which can be very elaborate. The phylum Cercozoa was originally erected by Cavalier- Smith (1998) to accommodate the euglyphid filose amoebae, along with the heterotrophic cercomonadids and thaumatomonad flagellates, which were shown to be related by Cavalier- Smith and Chao (1997). However, the origins of both Cercozoa and Foraminifera have been evolutionary puzzles because foraminiferal ribosomal RNA gene sequences are generally divergent, Biology and Evolutionary History of Larger Benthic Foraminifera 5 and show dramatic fluctuations in evolutionary rates that conflict with fossil evidence. Ribosomal RNA gene trees have suggested that Foraminifera are closely related to slime moulds and amoebae (Pawlowski et al., 1994), or alternatively used to suggest that they are an extremely ancient eukaryotic lineage (Pawlowski et al., 1996). In 2003, Archibald et al. found that cercozoan and foraminiferal polyubiquitin genes (76 amino acid proteins) contain a shared derived character, a unique insertion, which implies that Foraminifera and Cercozoa indeed share a common ancestor. Archibald et al. (2003) proposed a cercozoan- foraminiferan supergroup to unite these two large and diverse eukaryotic groups. In recent molecular phylogenetic studies, Nikolaev et  al. (2004) adopted the name “Rhizaria” (proposed first by Cavalier-S mith, (2002), which refers to the root- like filose or reticulose pseudopodia) and included the Retaria, Cercozoa and Foraminifera within this supergroup. While additional protein data, and future molecular studies on Rhizaria, Retaria, Cercozoa and Foraminifera, are necessary to provide a better insight into the evolution of the pseudopodial divisions, the placement of the Foraminifera within the Rhizaria appears to be well supported (Pawlowski and Burki, 2009; Ruggiero et al., 2015; Burki et al., 2016) (see Fig. 1.5). Fig. 1.5. A consensus phylogeny of eukaryotes from Burki et al., (2016). 6 Evolution and Geological Significance of Larger Benthic Foraminifera Similarly, the higher taxonomy of the Foraminifera is still unsettled. Although pro- posed as the Class Foraminifères by d’Orbigny (1826), throughout most of the 20th century the group was considered as the Order Foraminiferida, and the major subdivi- sions were considered to be suborders. In 1992, Loeblich and Tappan recommended Lee’s (1990) re- elevation of the Order Foraminiferida to Class Foraminifera, thereby elevating the suborders to orders. Sen Gupta (1999), Platon et al. (2001) adopted the class- level designation with some modifications at the order-l evel that have been largely supported by molecular phylogenies (Mikhalevich, 2000, 2004; Pawlowski and Burki, 2009; Groussin et al., 2011). Most recently Ruggiero et al. (2015) suggested a subphy- lum status for the Foraminifera. Recognizing that the classification of Foraminifera is still in flux, in this edi- tion (in contrast to BouDagher- Fadel (2008)) we accept the elevation of the Order Foraminiferida to Class Foraminifera, and the concomitant elevating of the previously recognized suborders to the ordinal level. 1.1.2 Larger Benthic Foraminifera Foraminifera are separated into two groups following their life strategy, namely the planktonic and the benthic foraminifera. Fewer than 100 extant species of foraminifera are planktonic, though they occur worldwide in broad latitudinal and temperature belts. They drift in the pelagic waters of the open ocean as part of the marine zooplankton (see Fig.  1.6). Their wide distribution and rapid evo- lution reflect their successful colonization of the pelagic realm. When this wide geographical range, achieved through the Late Mesozoic and in the Cenozoic, is combined with a short stratigraphic time range due to their rapid evolutionary char- acteristic, they make excellent index fossils at family, generic and species levels (see BouDagher-F adel, 2013, 2015). The benthic foraminifera, however, are far more diverse, with estimates of roughly 10,000 extant species. Benthic foraminifera live, attached to a substrate or free of any attachment, at all ocean depths, and include an informal group of foraminifera with complicated internal structures known as “larger benthic foraminifera”. It is these forms that are the principle subject of this book. The larger benthic foraminifera are not necessarily morphologically bigger than other benthic foraminifera, although many are, but they are characterised by hav- ing internally complicated tests. While one can identify most small benthic foramin- ifera from their external morphology, one must study thin sections to identify many of the larger benthic foraminifera, using features of their internal test architecture (Fig. 1.7). Larger benthic foraminifera develop characteristically complicated endoskele- tons, which permit the taxa to be accurately identified, even when they are randomly thin-s ectioned. The tests of dead, larger foraminifera can dominate carbonate sediments, and foraminiferal- limestones are extensively developed in the Upper Paleozoic, the Mesozoic (especially the Upper Cretaceous) and in the Cenozoic (see Fig. 1.8). Biology and Evolutionary History of Larger Benthic Foraminifera 7 A B Fig. 1.6. (A) Globigerinoides sacculifer (Brady), a spinose species with symbionts carried out by rhizopo- dial streaming on the spines (courtesy of Dr Kate Darling); (B) Neogloboquadrina dutertrei (d’Orbigny), a non- spinose species (courtesy of Dr Kate Darling). See BouDagher-Fadel, 2015 for a detailed study of the planktonic foraminifera mode of life, classification and distribution. Following recent taxonomic studies and reassessments of classifications, we recog- nise 14 different large benthic foraminiferal orders (Fig. 1.9). The orders with larger foraminiferal lineages that are discussed in detail in this volume are the: • Parathuramminida • Moravamminida • Archaediscida • Earlandiida • Palaeotextulariida • Tetrataxida • Tournayellida • Endothyrida • Fusulinida • Lagenida • Involutinida • Miliolida • Textulariida • Rotaliida. 8 Evolution and Geological Significance of Larger Benthic Foraminifera A Equatorial Oblique section equatorial section Axial section Oblique axial section B Equatorial section Axial section Proloculus Piles Equatorial or main chamberlet cycles Lateral chamberlets Fig. 1.7. Examples of two dimensional sections through the three-d imensional test of a larger, three layered foraminifera. A) Sections through a milioline test (modified from Reichel, 1964). B) Three-d imensional view of Lepidocyclina sp., showing the distinction between equatorial or main chamberlet cycles and lateral cham- bers (modified from Vlerk and Umbgrove, 1927). Throughout this book standard nomenclature is followed, so orders are expressed via the suffix “– ida”, or generically as “– ides” (e.g. Miliolida or miliolides). The suffix of “– oidea” is used to denote superfamilies, rather than the older suffix “- acea” following the recommendation of the International Commission on Zoological Nomenclature (see the ‘International Code of Zoological Nomenclature’, 1999, p. 32, Article 29.2). Families are designated via the suffix “– idae”. In this book, the suffix “– ids” is used to indicate a generic superfamily or family member (e.g. Fusulinoidea/ Fusulinidae or fusulinids). The study of living larger foraminifera shows that they occur abundantly in the shelf regions of most tropical and subtropical shallow marine, especially in carbonate- rich, environments. Indeed, they seem to have done so ever since the first larger foraminifera emerged during the Carboniferous. Again, from the study of extant forms, it seems that many larger foraminifera enclose photosynthetic sym- bionts, which appear to be essential to the development of most of the lineages with morphologically larger forms (Hallock, 1985; BouDagher-F adel et al., 2000; BouDagher- Fadel, 2008). Biology and Evolutionary History of Larger Benthic Foraminifera 9 a c b A B Fig. 1.8. A. Eocene limestone containing fossil porcelaneous foraminifera; a) Alveolina sp., b) Orbitolites sp., c) Quinqueloculina sp., from France. B.  Miocene limestone dominated by Lepidocyclina spp. from Indonesia, courtesy of Peter Lunt. From their structural complexity, and because of the diversity of the shelf envi- ronments that they inhabited, fossil larger foraminifera provide unique information on palaeoenvironments and biostratigraphy of shelf limestones around the world. Generally, in such environments, calcareous nannofossils are unavailable because of the shallowness of the marine environment and because of the recrystallisation of the calcite in the limestone matrices. Furthermore, macrofossils are relatively rare in these habitats. By the late 1920s, the larger foraminifera had become the preferred fos- sil group for biostratigraphy in several oil-r ich regions including the Indonesian area, parts of Russia, and in the United States, especially western Texas. Larger foramin- ifera had the advantage that they were more abundant than molluscs, and additionally a scheme was developed that utilised assemblage zones, rather than percentages of forms to be found. Using molluscs to identify and correlate sections required exten- sive knowledge of both living and fossil species. The larger foraminiferal assemblage zones could be identified by the presence of a few key taxa, usually with use of a hand 10 Evolution and Geological Significance of Larger Benthic Foraminifera Fig. 1.9 The geological range of the larger foraminifera orders and some selected, important families. Biology and Evolutionary History of Larger Benthic Foraminifera 11 lens in the field. Some groups of larger foraminifera provide excellent biostratigraphic markers, sometimes the only ones which can be used to date carbonate successions (e.g. the fusulinids and schwagerinids in the Upper Paleozoic (Fig. 1.10A; 1.10B), orbitoi- doids in the Middle to Upper Cretaceous (Fig. 1.10C), nummulitids in the Paleogene (Fig. 1.10D), and lepidocyclinids (Fig. 1.10E) and miogypsinids in the Oligocene and Neogene (Fig. 1.10F)). Provincialism was often a problem in these groups, but this is now well understood, so that biozonal schemes applicable to certain time intervals in defined bioprovinces have recently been erected and successfully applied (BouDagher -Fadel and Price, 2010a, b, 2014; BouDagher et al., 2015). Larger foraminifera are an ideal “group” of organisms to use in the study of general evolutionary theory. Their fossil record is so rich in individual fossils that assemblage concepts can be used, and both horizontal and vertical variation can be studied in the stratigraphic record. Their preference for certain marine environments is well under- stood and documented. Because representatives of most of the orders are still extant, it is also possible to infer their reproductive strategy, which as will be seen later, is quite complex. This book does not attempt to present a comprehensive or extensive listing of all genera and species of larger foraminifera, but rather focuses on the taxonomy, phylogeny and biostratigraphic applications of the most important forms. For an almost comprehensive list, the reader can refer to Loeblich and Tappan (1988). In addi- tion, for brevity, the complete references to genera and species are not given and again A B C D E F Fig. 1.10 Examples of larger foraminifera which provide excellent biostratigraphic markers, A) Fusulina; B) Schwagerina; C) Lepidorbitoides; D) Nummulites; E) Lepidocyclina; F) Miogypsina. 12 Evolution and Geological Significance of Larger Benthic Foraminifera the reader can refer to Loeblich and Tappan (1988) and the contemporary, on- line lit- erature. Finally, the reader can refer to Hottinger (2006) for an exhaustive set of defini- tions of terms used in the taxonomic description of the larger foraminifera, many of which, but inevitably not all, are also explained below. 1.1.3 Trimorphic life cycle in larger benthic foraminifera Larger foraminifera may reproduce asexually by multiple fission, producing many hundreds of offspring, and at other times they reproduce sexually, many by broad- casting gametes. Röttger (1983) described the asexual reproduction of Heterostegina. He stated that the protoplasmic body leaves the test through an internally developed canal system. The spherical daughter cells are colourless and are without a calcareous test (Fig.  1.11; Plate 1.3, Fig.  3). A  small part of the symbiont- containing residual protoplasm is then apportioned to each daughter cell. At this stage the test is formed and consist of two chambers, which as the juvenile grows are followed by the addition of further chambers. The growth rate of the calcareous tests of foraminifera is light- dependent (Röttger, 1983). Fig. 1.11. Schematic figures (in the centre) showing a trimorphic life- cycle of the larger benthic foramin- ifera Amphistegina gibbosa from Dettmering et al. (1998). The upper part shows the dimorphic life cycle, con- sisting of an alternation between a haploid, megalospheric gamont with its gametes, and the microspheric diploid agamont with its offspring produced by multiple fission. The lower part represents the megalospheric generation in a trimorphic cycle reproduced by cyclic schizogony, inserted between agamont and schizont. The photographs of living Heterostegina depressa (courtesy of Prof R.  Röttger) show an alternation of generations in which a 2- 4mm sized gamont (megalospheric generation) (on the left) alternates with an 1- 2cm- sized agamont (microspheric generation) (on the right). During multiple fission of the agamont, the symbionts- containing protoplasm (top right) flows out of the calcareous test and then divides into 1000 to 3000 daughter individuals, the young gamonts. In addition to gamont and agamont forms another genera- tion, which looks like a gamont of Heterostegina depressa, but which reproduces asexually (Röttger, 1990). Biology and Evolutionary History of Larger Benthic Foraminifera 13 In other taxa, the process has some small differences. For example, in the spe- cies Amphistegina spp., the cytoplasm exits through the aperture (see Fig. 1.12). In the soritid foraminifera, the partitions of the final chambers are dissolved to form a brood chamber in which the daughter cells form (Röttger, 1984; 1990). After asexual multiple fission, the empty parent test becomes a lifeless grain of sediment. Larger foraminifera are dimorphic (having two forms), which is the result of the het- erophasic alternation of generations between a haploid, uninucleate gamont (the sexual generation which produces gametes) and a diploid, multinucleate agamont (the asexual generation which produces daughter individuals by multiple fission) (Schaudinn, 1895; Röttger, 1990). The dimorphic forms usually exhibit different morphological charac- ters; the two forms are called: • The asexual microspheric (or B-) form, which is larger, with numerous chambers, but with a small proloculus (first chamber, see Plate 1.3 and Fig. 1.11). It is this asexual generation which produces daughter individuals by multiple fission, and • the megalospheric (or A- ) form, which is smaller with fewer chambers, but with a large proloculus. It is this sexual generation which usually produces gametes (see Fig. 1.11). However, in addition to these two generations, a third generation is documented by many authors, where the agamont produces megalospheric schizonts instead of gam- onts (see Fig.  1.11). This life cycle was first discovered by Rhumbler (1909), and since then has been recognised by many authors (Leutenegger, 1977; Röttger, 1990; pseudopodia aperture umbo stellar chamberlet pustules stellar suture A B Septum of main chambers Hemiseptula Fig. 1.12. Amphistegina: A) an axial section of a fossil specimen of Amphistegina; B) A live specimen show- ing the cytoplasm exiting through the aperture. 14 Evolution and Geological Significance of Larger Benthic Foraminifera Dettmering et al., 1998; Harney et al., 1998). Röttger (1990) cultivated Heterostegina depressa (Plate 1.3, Fig. 1) in the laboratory and was able to confirm the trimorphic cycle. Dettmering et al. (1998) and Harney et al. (1998) suggested that the trimor- phic cycle can account for the abundance of the megalospheric generation in many populations. The schizonts, which are produced by asexual reproduction, in con- trast to zygotes, which are too small to carry symbionts, begin their ontogeny as large symbiont- bearing cells. Harney et al. (1998) also suggested that the trimorphic cycle provides tremendous colonization potential, allowing foraminifera to rapidly increase their population densities sufficient to successfully sexually reproduce by gamete broadcasting, while at the same time promoting genetic divergence by ampli- fying the colonizing genotypes, all of which could promote relatively rapid rates of evolution. 1.2 Morphological and Taxonomic Features Used in the Classification of Larger Foraminifera Larger foraminifera are subdivided into six groups according to the wall structure of their tests (see Fig. 1.13): • the agglutinated group, with walls composed of detrital particles held together by calcareous cement (as in the larger Textulariida), • the calcareous granular group, with compound, microgranular walls of low-M g cal- cite, in which the crystalline grains are without optical alignment (characteristic of the Fusulinida and related orders), • the porcelaneous group, composed of three- layered calcitic, imperforate, non- lamellar walls with a high percentage of rod- like magnesium calcite that have their axes ran- domly oriented in the embedding organic material and with an outer layer parallel to the outer walls, as shown by the Miliolida, • the hyaline calcareous group, a lamellar- perforate group, consisting of layers of cal- cite crystals, with the C- axis oriented perpendicular to the test surface (Haynes, 1981; Hallock, 1999). The magnesium ratio is low in some taxa and high in others. This wall structure is characteristic of the Rotaliida (e.g. Fig. 1.14). The pore canals in these perforate tests have proximal ends closed by organic membrane with micro- pores (Röttger, 1983). They do not, therefore, allow the passage of cytoplasm to the seawater, but they facilitate the transport of carbon dioxide, oxygen and nutrient salts in the symbiont-b earing larger benthic foraminifera. • The monolamellar group, with or without secondary laminations, with radiating cal- cite crystals which have the crystallographic c-a xis perpendicular to the surface, as shown by the Lagenida, and • The aragonitic group, commonly they are recrystallised to give a homogeneous microgranular structure. This wall is characteristic of the Involutinida. The wall structures of the larger foraminifera reflect the biological method used by their living cell to build its test. The microgranular walls developed by the fusulinides of the Paleozoic have a thin, dense outer layer forming the spirotheca (spiral outer Biology and Evolutionary History of Larger Benthic Foraminifera 15 a b c d Fig. 1.13. Wall structure of the larger foraminifera. A) Loftusia sp. (agglutinated); B) Alveolina sp. (calcar- eous imperforate); C) Quasifusulina sp. (calcareous microgranular); D) Rotalia sp. (calcareous perforate). Scale bars (1- 3) = 2mm; (4) 0.5mm. wall). In advanced fusulinides, the wall becomes alveolar (i.e., it develops small sacks), and has a honeycomb- like structure. The term “keriotheca” is restricted to structures with two layers of alveoles (Figs 1.13C and 1.15; see Chapter 2). The rotaliides test is made of perforate hyaline lamellar calcite, and many of the larger rotaliides are charac- terized by having a developed canal system (Fig. 1.14, and see Chapter 6), which gives rise to special laminations in the tests (Hottinger, 1977). The lamellar tests are formed during the process of chamber construction, where each chamber wall, consisting of secreted, Mg- calcite, covers the total test including all former chambers (see Figs. 1.14; Hohenegger et al., 2001). The basic structural element of the test is the chamber. Larger benthic foramin- ifera have multichambered (plurilocular) tests, which have attained large sizes, up to ~150mm in the case of Cycloclypeus carpenteri (See Chapter  7). The internal space between the chamber walls is called the chamber lumen. All cavities subdividing the chambers are called chamberlets. Hottinger (2006) divided the basic architectural com- ponents of the foraminiferal test into elements that do not modify the shape of the living cell, such as the wall, and those that do modify it. The elements that modify the shape of the living cell can be, according to Hottinger (2006), divided into three factors. The first factor is the shape of the first chamber (proloculus) and subsequently the growth of the second chamber (deuteroloculus). The chambers are separated by a wall (the septum) and connected by the intercameral 16 Evolution and Geological Significance of Larger Benthic Foraminifera Outer layer Inner layer Bilamellar c b canal system Mono-lamellar aperture initial chamber (proloculus) a foramen Non-lamellar d Fig. 1.14. The structure of A) non- lamellar, B) mono- lamellar and C) bilamellar test walls, where the sep- tum has an inner and outer primary lamella, separated by an organic layer, and is secondarily doubled distally by the “septal flap” formed from the inner lamella of the succeeding chamber, D) a rotaliine test showing that the open external spaces between juxtaposed chamber walls (intraseptal spaces), and between successive shell whorls as they become enclosed by the outer lamellae of newer chambers, thereby forming a canal system (modified after Haynes, 1981). B Equatorial section Septum A Proloculus Chomata Fluted septum Keriotheca Chomata Axial secton C Fig. 1.15. Views of a schematic fusulinide, A) 3- dimensional view of test; B) equatorial section of Fusulinella sp.; C) axial section of Triticites sp. illustrating the development of secondary deposits of calcite (chomata) on the chamber floor, and the development of contorted, “fluted” septa. Biology and Evolutionary History of Larger Benthic Foraminifera 17 lumen. Tubular foramen are called stolons, and if they are wide open they are called tunnels (Fig. 1.16). The two other factors that determine the shape of the living cells are the chamber shape and the arrangement of the chambers (Fig. 1.17). The arrangement of the cham- bers may form a compressed, planispiral, involute and flaring growth, e.g. Archaias (Fig.  1.18Aa), fusiform elongate test, e.g. Flosculinella (Fig.  1.18Ab), Alveolinella (Fig. 1.18B), annular concentric (in two dimensions), e.g., Cyclorbiculina (Fig. 1.18C), Marginopora (Fig.  1.18D), or spherical- concentric (in three dimensions) test, e.g., Sphaerogypsina (see Fig. 1.18F). The chambers can be developed in a serial arrange- ment, uniserial (chambers arranged in a single row), biserial (chambers arranged in two rows), etc., or in a spiral arrangement, such as the streptospiral arrangement, where coil- ing occurs in different planes, the planispiral arrangement where the spiral and umbilical A Stolons Proloculus Deuteroloculus Septa B Fig. 1.16. Two different forms of larger foraminifera: A) the elongated test of a miogypsinid, where the deuteroloculus is in alignment with the proloculus and the test is elongated; B) the lenticular test of a hetero- stegine where the position of the deuteroloculus makes it essential for the test to enrol. 18 Evolution and Geological Significance of Larger Benthic Foraminifera A B C D E a b c F a b c G Fig. 1.17. Chamber arrangement of A) equatorial section of Archaediscus, with a streptospiral test, with an undivided tubular second chamber; B) an axial section of Tournayella showing a planispiral test in the adult; C) an equatorial section of Endothyra showing an initially streptospiral to planispiral test with well developed septa and characterised by the development of secondary deposits of calcite (chomata) on the chamber floor; D) an equatorial section of a planispiral evolute test; E) an axial section of an evolute test; F) a, c) side views of an involute test, b), apertural view; G) a trochospiral test, a) umbilical side, b) vertical view, c) spiral side. b a A B C D E F Fig.  1.18. A) Thin section photomicrograph showing a) compressed, planispiral, involute and flaring growth, Archaias; b) a fusiform test, Flosculinella; B) fusiform elongate test, Alveolinella; C) annular con- centric growth, Cyclorbiculina; D) annular concentric growth Marginopora; E) annular concentric growth, Archaias; F) spherical concentric growth in Sphaerogypsina. Scale bars (A-B )  =  2mm; (C- E)  =  1mm; (F) = 0.5mm. Biology and Evolutionary History of Larger Benthic Foraminifera 19 sides are identical and symmetrical (Fig. 1.17F), and the trochospiral where spiral and umbilical sides are dissimilar (Fig. 1.17G). The trochospiral arrangement in the larger foraminifera exposes the umbilical region and creates a direct access to the ambient environment (Hottinger, 1978). In involute spiral forms, the lumina of the chambers in one coil cover laterally those of the preceding coil (e.g., Nummulites, Fig. 1.19A) and develop in some cases wing-l ike extensions from the lumen to the poles (alar prolon- gation). However, in a spirally coiled evolute form, the chamber lumina do not later- ally cover those of the preceding coil (e.g., Assilina, Fig.  1.19C). Thus, for example, Operculina has a planispiral, evolute lenticular, compressed and loosely coiled test (see Fig. 1.19B), while Heterostegina has a planispiral, involute to evolute test with chambers divided by secondary septa to form small chamberlets (Fig. 1.19E). In Spiroclypeus, the heterostegine chambers increase rapidly in height and project backwards (Fig. 1.19D). Alar A B C D E prolongation Fig. 1.19. Differing shapes of nummulitic tests. A) Involute test, axial section of Nummulites; B) Evolute test, axial section of Operculina; C) Evolute test, axial section of Assilina; D) Involute test, Spiroclypeus; E) Test initially involute, evolute in mature stage, Heterostegina (see Chapter 6). Scale bars = 0.5mm. 20 Evolution and Geological Significance of Larger Benthic Foraminifera Exoskeletal elements are developed that reflect protoplasmic flux (Hottinger, 2006). These include alveoles (honeycomb- like sacks), reticular subepidermal networks, etc., that produce multiple, small, blind- ended small chamberlets/ tiny compartments of the chamber cavity coated by organic lining. (Hottinger, 2000). Various agglutinated larger benthic foraminifera develop layers of alveoles, coating the lateral chamber wall, such as in the lituolids that have an exoskeletal layer of undivided, shallow alveoles., e.g., the early extinct representative Pseudocyclammina (Fig. 1.20A) and Everticyclammina sp. (Fig. 1.20B) and the still living Cyclammina (Fig. 1.20C). The alveoles in the por- celaneous Paleogene Alveolina (Fig.  20D) are blind recesses separated by septula. Alveolinids, such as Subalveolina or Bullalveolina (See Chapter 6) have alveoles in post- septal positions over supplementary apertures in the previous septal face. The Neogene genus of Textulariella has branching alveoles, while in the porcelaneous foraminifera, Austrotrillina alveoles (Fig. 1.20E) evolve from early forms with layers of shallow, undi- vided alveoles (see Chapter 6) to deep and branching alveoles (A. howchini, Fig. 20E) in order to harbour symbiotic algae. The early fusulinides had keriothecal cavities in their walls, which is described by Hottinger (2000) as an alveolar, honeycomblike structure with a spiral wall, not filled with living chamber plasma nor coated by the organic lining. In advanced fusulinides, the keriotheca may consist of an outer and an inner “layer” produced by a split of the alveoli into narrower subunits below the tectum, while others have both alveolar struc- tures and keriothecal wall texture (e.g. Verbeekina,see Chapter 2.) Some agglutinated larger benthic foraminifera have parapores (canaliculi), straight to tortuous tubular spaces, coated and closed off internally by organic lining (e.g., Chrysalidina, Fig. 1.20F). Others, have combined alveolar exoskeletons with a parapor- ous external wall (e.g., Dicyclina, see Chapter 5) or with a bilamellar perforate wall (e.g., Fabiania, see Chapter 6). Only agglutinated foraminifera possess exoskeletal polygonal structures called “sub- epidermal networks” (e.g. in Orbitolina and Pseudochoffatella, Fig. 1.21). The basal layer of the larger foraminiferal test, when thickened as in flosculinisation (Hottinger, 1960) or perforated by canalicular passages (Hottinger, 1978), exist only in the porce- laneous forms (Fig. 1.20D). Other exoskeletal features found in larger foraminifera are the partitions of the chamber lumen, by for example beams, which are perpendicular to the septum, or raf- ters which are parallel to the septum (Fig. 1.22). In many larger, recent larger foraminifera, the exoskeletal alveoles harbour photo- synthesising symbionts in internal pores. These pores are also seen in extinct Cenozoic species such as in Miogypsina Fig. 1.23A). However, exoskeletal structures also exist in species, such as Cyclammina (see Chapter 6), that live at depths too great to have pho- tosynthetic symbionts. Hottinger (2000) interpreted the exoskeletal structures of these deep- water larger foraminifera as providing a mechanism that permits control of gas exchange, by separating the gas diffusion from protoplasmic streaming. However, the endoskeleton of many larger foraminifera includes pillars (Fig. 1.23B), which fill the interior of the test, or continuous walls (septula), which subdivide the larger chamber lumen (see Fig. 1.23). Pillars may also be seen as providing mechanical strength to the test, so for example in discoidal forms, heavily pillared endoskeletons, as a rule, occur Biology and Evolutionary History of Larger Benthic Foraminifera 21 A B C D E F Fig.  1.20. Alveoles and parapores. A-D ) Wall covered by alveoles which are subepidermal blind cham- berlets/ recesses coated by the organic lining:  A- B) Cretaceous extinct agglutinated foraminifera, A) Pseudocyclammina; B) Everticyclammina; C) Cretaceous to Holocene Cyclammina sp.; D) an Eocene mili- oline foraminifera, Alveolina elliptica var. nuttalli, which shows also some degree of flosculinisation, (thick- ening of the basal layer of the early chambers). The “alveoles” are blind recesses separated by septula; E) Austrotrillina hochini with deep and branching alveoles which evolve from earlier forms with layers of shallow, undivided alveoles (see Chapter 7); F) Paraporous wall with tubular spaces, coated and closed off internally by the organic lining, Cretaceous Chrysalidina. Scale bars = 0.5mm. A B Fig. 1.21. Subepidermal networks: A) Pseudochoffatella; B) Orbitolina. Scale bars = 0.5mm. 22 Evolution and Geological Significance of Larger Benthic Foraminifera Beams Rafters Proloculus Chamber lumen Fig. 1.22. Schematic test of Alzonella, showing a planispiral test with a hypodermis consisting of a coarse lattice of beams and rafters. in forms living in very shallow, turbulent water (e.g., Archaias, Fig. 1.18A and E), while modestly pillared forms (e.g., Cyclorbiculina, Fig. 1.18C) inhabit deeper and quieter environments (Hottinger, 2000). The massive pillars in Lepidocyclina (see Fig. 1.23B) may also be thought to be associated with occupation of high-e nergy, marine environ- ments (BouDagher- Fadel et al., 2000); however, there are many exceptions to this rule. Larger foraminifera have many different overall adult shapes. Discoidal forms evolve progressively into flat tests, which can be generated by uniserial growth, such as in the orbitolinids (as in Orbitolina, Fig. 1.24A), spiral growth (as in Choffatella, Fig. 1.24B), and annular growth (as in Orbitopsella, Fig. 1.24C). An elongate form may be realized by a concentric growth pattern, such as in Lacazina (Fig. 1.24D), or in a planispiral- fusiform test, such as in Fusulina (Fig. 1.10A) and Alveolina (Fig. 1.13B). As larger foraminifera increased their sizes, their internal structures became more complicated. One of the most intriguing complication occurred in the fusulinides. They subdivided the inhabited space of the test by folding their test walls, thus creating sep- tal fluting (see Fig. 1.15). In very elongate forms the folded septa became disengaged from the chamber floor to create cunicular passages (see Chapter 2). This septal folding seems to be present in fusiform amphisteginids, such as Boreloides (Hottinger, 1978). In tightly coiled, elongate fusiform tests (e.g. in Alveolina), the function of the elongation of the fusiform test is related to motility, the test moving in the polar direction (but growing in equatorial direction) (see Chapter 6, and Hottinger, 2000). As the tests of the foraminifera become large, the protoplasmic body must inhabit all compartments and those compartments must be interconnected. Therefore, a sys- tem of apertures or stolons is necessary to shorten the distance between the first and final chambers, and to provide a communication route between the compartments Biology and Evolutionary History of Larger Benthic Foraminifera 23 a A- Miogypsina b Septulum c Internal pores Pillars B- Lepidocyclina Fig.  1.23. A) SEM image of Miogypsina, with enlargement of the lateral chamberlets to show internal pores that harbour symbionts; B) Lepidocyclina sp. with pillars embedded in the lateral parts of the test. Scale bars (A) 250 μm; (B) 1mm. (Hottinger, 1978, 2000). This can be provided by leaving a primary aperture or, in some cases, multiple apertures for extrusion of the rhizopods between the chambers during growth, and according to Hottinger, the linear nature of the rhizopodial protoplasm involved in wall building guarantees in some species that successive chambers are con- nected by a single foramen or aperture, the last formed chamber opening to the sur- rounding water via a terminal aperture. Some larger foraminifera enhanced the control of their chamberlet cycle growth by oblique, crossed-o ver stolon systems. Hottinger (2000) noted that in Mesozoic (see Chapters 4 and 5) non- perforate discoidal tests, the radial arrays (as in Orbitopsella) are more frequent than the oblique, crossed- over ones (Ilerdorbis), whereas, in conical forms, the latter pattern dominates without being exclu- sive (Orbitolina). During the Cenozoic, the discoidal tests with crossed- over stolon sys- tems prevail (Orbitolites, Marginopora, see Chapters 6 and 7), while in uniserial-c onical forms the crossover is less well developed (Dictyoconus, Chapmanina, see Chapters 6 and 7). Organic lining may cover the connections between the chamberlets, creating sealed compartments (Ferrandez-C anadell, 2002). In planispiral-f usiform to elongate 24 Evolution and Geological Significance of Larger Benthic Foraminifera A B C D Fig. 1.24. Examples of adult shapes A) uniserial growth as in Orbitolina; B) spiral growth, as in Choffatella; C) annular growth, as in Orbitopsella; D) concentric growth pattern, as in Lacazina. tests (as in the fusulinides and the alveolinids) or high-t rochospiral tests with a colu- mellar structure (kurnubiids, pfenderinids, see Chapter 5), the apertures are aligned around the centre and the columella are in the polar direction. As a result, the apertural face is enlarged to admit supplementary polar apertures (Hottinger, 2006, 2007). In the extinct fusulinides, it is believed that the rhizopods extruded from the sep- tal pores, which replaced the main aperture, in the apertural face (Hottinger, 2001). Similarly, many species of rotaliides do not have primary apertures (Hottinger, 1997, 2000). However, the chambers communicate instead by a canal system (Fig. 1.25) that replaces the true primary and secondary apertures (Röttger et al., 1984), and feeds the different cavities by opening into the ambient seawater. In Operculina, the canals allow communication between the chamber cavities and the lateral surface of the walls, while Heterostegina has in addition a three- dimensional network of canals within the mar- ginal cord (Murray, 1991). Nummulites also have a three-d imensional canal system within a thickened periph- eral keel (marginal cord). This canal system has multiple functions, such as locomo- tion, growth, excretion, reproduction and protection (Röttger, 1984). It permits the extrusion of the pseudopodia from any point of the marginal cord, provides the fora- minifera with radial symmetry and enables the disposal of waste products. During sexual reproduction, it enables the release of gametes, and during asexual reproduction it allows the extrusion of the cytoplasm and symbionts to the ambient seawater. Biology and Evolutionary History of Larger Benthic Foraminifera 25 Marginal cord Pores Chamber lumen Intraseptal canal system Fig. 1.25. The complicated canal system is visible within the chamber walls in an araldite cast of part of a shell of larger foraminifera (after Röttger, 1983). 1.3 Ecology of the Larger Foraminifera Most extant larger benthic Foraminifera are marine and neritic, living largely in warm, nutrient- poor, reef and carbonate shelf environments, where they are important pro- ducers of carbonate sediments (Fig. 1.26). It is inferred that larger benthic foraminifera had similar distributions in the Mesozoic and Cenozoic. Modern taxa have geographic ranges similar to that of hermatypic corals, although some larger foraminifera cer- tainly have a wider latitudinal distribution. Combining results from the study of the larger benthic and planktonic foraminifera provides an approximate guide to major changes in sea temperatures during the past 66  million years (McMillan, 2000). In general, the presence of larger benthic foraminifera in the fossil record indicates a warm environment, while their absence points to cooler or more nutrient-r ich environ- ments. Some extant larger foraminifera can tolerate water temperatures as low as 10- 11º C, including Amphistegina and Sorites in the Mediterranean (Hallock et al., 2011), and Amphisorus and Amphistegina on the southwest Australian shelf (Li et al., 1999). Depth distributions of larger foraminifera depend upon water transparency (Hallock, 1987; Mateu- Vicence et al., 2009), with some taxa such as Cycloclypeus found at depths exceeding 100 m (Hohenegger, 2004). It is inferred that the larger benthic foramin- ifera in the Cenozoic carbonates of Tethys occupied niches analogous to those filled by modern forms. Foraminifera typically gather food particles for extrathalamous digestion. Most larger foraminifera also have small chamberlets or cubiculae, which can act not only as a small convex lens for the focusing of sunlight, but also serve as “greenhouses” for the containment and development of symbiotic microalgae, which can provide the host foraminifera much more energy than they can consume as food (Hallock, 1981). 26 Evolution and Geological Significance of Larger Benthic Foraminifera Trochammina Ozawainellidae Haplophragmella Ammobaculites Schubertellidae Verbeekininae Neoschwagerininae Involutina Earlandia Shanita Globuligerina Triadodiscus Endotriada Trocholina Siphovalvulina Choffatella Everticyclammina Haurania Kilianina Globotruncana Miliolids Simplorbites PraealveolinaOrbitoides Orbitolina Trocholina Bolivinoides Heterohelix Orbitolites Globigerina Bolivina Nummulites Praerhapydionina Alveolina Discocyclina GloborotaliaRotalia Peneroplis Large Ammonia miliolids Elphidium Ophthalmids Miogypsina Spiroclypeus Sorites LepidocyclinaPeneroplis Borelis miliolids Bolivina Globigerina Pararotalia Operculina Ammobaculites Marginopora Archaias Alveolinella Baculogypsina Heterostegina Bulimina Globigerinoides Reef top mixed fauna Reef wall Talus bioherm Tidal swamps Back-reef detritus Bioherm of bioherm e.g. Mangrove or fringing reef coal bearing Patch reef and in the Palaeozoic fore-reef detritus Fore-reef Basin Fig.  1.26. The ecological distribution of larger and key smaller benthic and planktonic foraminifera through space and time. Most extant larger benthic foraminifera host endosymbiotic unicellular algae, such as rhodophytes, chlorophytes, diatoms or dinoflagellates, which enhance growth and calcification in much the same way as in zooxanthellate corals (Lee and Anderson, 1991). Algal symbiosis provides the foraminifera with a reliable source of energy in water that is poor in other food sources and allows them to recycle nutrients, a necessary strategy of life in environments where nutrients, not light energy, are the limiting factor for survival (Hottinger, 2000). These symbionts also determine the colour of their host foraminifera, which tend to be brown and yellow when hosting diatoms or dinoflagellates, red to violet with rhodophytes, and green with chloro- phytes (Röttger, 1983). PALEOGENE CRETACEOUS JURASSIC TRIASSIC PERMO -HOLOCENE NEOGENE CARB. Biology and Evolutionary History of Larger Benthic Foraminifera 27 Larger benthic foraminifera, which are discoidal and fusiform in shape, likely have achieved their large size because of such symbiotic associations. According to ter Kuile (1991) these endosymbionts release photosynthates into their hosts, and consume CO2 during photosynthesis, producing CO 2- 3 , which allows high rates of CaCO3 precipita- tion during test growth and calcification. It follows that larger benthic foraminifera are very sensitive to light levels. However, McConnaughey (1989) and McConnaughey and Whelan (1997) have proposed the reverse interpretation as a role for algal symbiosis. They suggest that lack of CO2 limits photosynthesis in warm, shallow environments and that calcification provides protons and make CO2 more readily available. These proposed benefits are not mutually exclusive. Because the symbionts tend to be located in the endoplasm, within the older chambers, photosynthesis does not occur in ecto- plasm, where calcification primarily occurs. Thus, the primary benefit of algal sym- biosis is likely the fixation of solar energy that can be used in all functions, including proton pumping and other cellular active- transport processes required for calcification (Pomar and Hallock, 2008). Although some benthic foraminifera are “r- strategists” (well adapted to an expo- nential increase in population size as they have the ability to produce large numbers of offspring), larger foraminifera have largely been considered “K-s trategists” (Hallock, 1985; Hottinger, 2007), where the terms r and K, come from standard ecological analy- ses such as the Verhulst model of population dynamics (Verhulst, 1838). K- strategists have limited ability to rapidly increase their population densities. They have relatively long life spans, large sizes, delayed reproduction, and invest a relatively large amount of energy into each of the offspring they produce. They also exhibit low reproductive effort in maturity. This leads to morphological adaptation and increase in complexity (Gould, 1977). The long life spans of larger foraminifera is documented by authors such as Purton and Brasier (1999), who, by using oxygen and carbon isotope var- iation in annular cycles in the test of Nummulites laevigatus (see Chapter 6), were able to deduce that this species lived at least 5 years and the largest Nummulites could be many years older than that. As a result of the prolonged adult stage and their relatively large sizes, larger foraminifera exhibit strong hypermorphic mutations which lead to complex morphological characters (McKinney and McNamara, 1991; Lunt and Allan, 2004). Gould (1977) and McKinney and McNamara (1991) have linked Cope’s (1896) rule (the increase in size of organisms during their evolutionary history) to K-s trategy and hyper- morphosis. The K-s trategy mode of life usually occurs in relatively stable environments, as it requires delayed maturity, fewer offspring, and therefore lower reproductive poten- tial. Gould (1977) notes that reaching sexual maturity usually marks the termination of growth and size increase of most organisms. This is certainly the case for larger foramin- ifera, where sexual (and asexual) reproduction usually coincides with death of the parent (Hallock, 1985). Harney et al. (1998), however, postulated that the trimorphic life cycle allows larger foraminifera to become r-s trategists when subjected to ecological stress, as successive asexual generations can more rapidly build up the population density than can strict alternation of generations. Fermont (1982) documented a 50% reduction in test size in two orbitoid species that survived the Cretaceous- Paleogene extinction. However, during other periods in the evolutionary history of larger foraminifera, such as in the Triassic (see Chapter 3), only opportunistic, small, short-l ived r- strategist foraminifera are believed to have survived after major, global extinctions events. 28 Evolution and Geological Significance of Larger Benthic Foraminifera In recent years attempts have been made to understand the palaeobiology of larger benthic foraminifera using the oxygen and carbon isotopic compositions of fossil tests (Purton and Brasier, 1999), and from living assemblages (Saraswati et al., 2003). The presence of algal endosymbionts leads to disequilibrium fractionation of isotopes (Hansen and Buchardt, 1977; Saraswati et al., 2003). Carbon isotopes show this effect more than oxygen isotopes (Saraswati et al., 2003). In reef-fl at environments, the size of the test of the symbiont-b earing foraminifera affects the oxygen and the carbon isotopic variations, while in deeper waters oxygen isotopic values vary little with the size of the test. These differences likely reflect differences in the variability of the envi- ronment during the life span of the foraminifera. Saraswati et  al. (2003) concluded from their observations of miliolides and rotaliides that the two groups differ distinctly in their carbon isotopic fractionation; the low and medium fractionated taxa belong to Miliolida and the highly fractionated taxa belong to Rotaliida. Although micro- habitat also appears to have a role in carbon isotope variation, the relative contri- butions of biomineralization, metabolism and microhabitat are difficult to estimate. Langer (1995) analysed five species from a lagoon in Papua New Guinea and examined isotopic composition with depth. These studies again showed a general trend in the depletion of heavier isotopes of oxygen and carbon with depth and intensity of light. Wefer and Berger (1980) recorded isotopic variation within an individual specimen of Marginopora vertebralis by sampling along the direction of growth (ontogeny), and they inferred that oxygen isotopes reflected seasonal variation in temperature. The car- bon isotopic composition is more variable between the specimens of the same species (Saraswati, 2004). Saraswati (2004) concluded that ontogenetic oxygen isotope varia- tion decreases progressively in deeper water species. Studies of living larger benthic foraminifera, in controlled laboratory environments, have provided some further information regarding life strategies (e.g., by the culture of Heterostegina depressa; Röttger, 1984), but much has been inferred by relating test morphology to habitat (Hallock, 1981; Murray, 1991). Predators such as bristle worms, crustacea, hermit crabs, snails, gastropods, echinoderms and fish, as well as microscopic predators (including other foraminifers (Hallock and Talge, 1994), some nematodes (roundworms) and flatworms), selectively feed upon foraminifera. Such predation pressure, of course, will depress the foraminiferal populations and, in most cases, the observed thanatocoenosis (death assemblage) is not fully representative of the living population (biocoenosis). Living distribution patterns of the symbiont-b earing larger foraminifera are con- fined to tropical, subtropical and warm-t emperate photic marine environments, as their distribution is determined by a complex set of inter- related parameters such as temperature, nutrient availability, water transparency and light intensity (Renema, 2002). Water depth is a secondary factor related to the distribution of larger benthic foraminifera, because light intensity, temperature and hydrodynamic energy decrease with depth. Some larger foraminifera, such as Amphistegina (Plates 1.3 and 1.4), are known to become flatter, with thinner outer walls, with increasing water depth and decreasing light intensity (Hallock et al., 1986). Imperforate foraminifera, such as the miliolides, are generally restricted to shallower depths than perforate forms (Hallock, 1988; Hottinger, 2000). However, both perforate and imperforate larger foramin- ifera house symbionts, and the dependence on light for their symbionts limits their Biology and Evolutionary History of Larger Benthic Foraminifera 29 distribution to the photic zone. The depth distribution of living larger benthic forami- niferal taxa is related to the transparency of their test walls (porcelaneous versus hya- line) and the light wavelengths required by their symbionts (see Renema, 2002), e.g., Archaias (0– 20 m, chlorophytes, red light), Peneroplis (0– 70 m, rhodophytes, yellow light) and Amphistegina (0– 130 m, diatoms, blue light) (Hallock, 1985). This has led many authors to use calcareous algae and larger foraminifera assemblages as proxy water- depth indicators in carbonate sediments (Banner and Simmons, 1994; Mateu- Vicens et al., 2009). The interpretation of use of sunlight by many fossil larger benthic foraminifera, such as Miogypsina, Miolepidocyclina, is shown by the common occurrence of miogyp- sinids only in shallow water marine limestones where fossil algae also occur (Fig. 1.26). The irregular shape of many species of Miogypsina, often revealing apparent divi- sion of their flanges (e.g., Miogypsina bifida, see Chapter 7), shows they could become concavo- – convex and were not growing on a flat surface. Such a shape would be devel- oped if the individual was attached to a strongly curved substrate, such as fronds of macroalgae or the stems or leaves of seagrass (the substrate would be biodegradable and seagrass is not seen preserved in thin sections). Therefore the sedentary, attached miogypsinids likely grew to accommodate the shape of the vegetable substrate to which they adhered. Only in strong ambient sunlight, which would benefit both the miogyp- sinids and their vegetable substrate, could true Miogypsina flourish (BouDagher- Fadel and Wilson, 2000). On the other hand, elongate forms, such as Alveolinella, can hide or shelter under shallow layers of coral sand, in order to regulate the illumination that they require. The ectoplasm of foraminifera forms the template on which new chamberlets are secreted (Röttger, 1984). Larger foraminifera can modify their shape and structure to some degree depending upon environmental conditions. Moreover, morphologic trends indicate both adaptation and acclimation (Hallock et al., 1986). Wide varia- tion in test structure and morphology has resulted from such adaptation. According to Larsen and Drooger (1977) and many subsequent researchers, the diameter- thickness (D/ T) ratio of many larger benthic foraminifera varies inversely with depth. The near-s hore samples have a higher D/ T than the more offshore species. Mateu- Vicens et  al. (2009) further quantified this relationship for Amphistegina, which can be used to estimate paleodepth and water transparency. Thus, the mor- phology of larger benthic foraminifera has been used for depth-e stimation in geo- logical facies interpretations, usually based on comparison with homeomorphs of living occurrences. Robust and fusiform tests, as seen in Fusulina and Alveolina, and conical (e.g. orbi- tolinids) or strongly biconvex (Amphistegina; Plates 1.3 and 1.4) forms are adapted to a life in environments of high hydrodynamic energy, characteristic of water depths less than 10m, but usually not in mobile sands, but rather on phytal or hard sub- strates like reef rubble (Hallock, 1985). Those with very thin and flat tests, however, can only live in very calm waters that tend to have lower levels of light intensity, and some of them such as Discospirina are able to live in deeper and cooler waters. The larger foraminifera found attached on sediments in deeper waters tend to be flat and discoidal in shape (e.g. Spiroclypeus). Forms adapted for adherence to seagrass or algae tend to be flat (Amphisorus and Cyclorbiculina), and sometimes relatively small 30 Evolution and Geological Significance of Larger Benthic Foraminifera (e.g., Peneroplis; Plate 1.4, Fig. 6). Some foraminifera develop some kind of anchor- age (e.g. the spines in calcarinids; Plate 1.4, Figs. 2– 3), or an ectoplasmic sheath (Heterostegina depressa; Plate 1.3, Fig. 1). The numbers of H. depressa tend to be low in shallow waters and restricted to shaded locations (Haynes, 1981). In some taxa, such as Amphistegina, the morphology of the apertural face can change to increase their potential to cling to surfaces in turbulent waters (see see Chapter 7). According to Hottinger (2000), in Alveolina and Fusulina the function of the elongated fusiform test is related to motility, the test growing in equatorial direction but moving in the polar direction. So, ever since the Carboniferous, larger foraminifera have thrived in the shallow, warm, marine environments (see Fig. 1.26). Their remarkable abundance and diversity is due to their ability to exploit a range of ecological niches by having their tests utilised as greenhouses for symbionts. However, attaining large size made some forms very spe- cialised and vulnerable to rapid ecological changes. For this reason K- strategist, larger foraminifera show a tendency to suffer periodic major extinctions when environmental conditions change rapidly and/ or substantially. This makes them valuable biostrati- graphic zone fossils, and, as will be explored in the following section, also provides a valuable insight into the general process of biological evolution, including parallel and convergent evolutionary trends. 1.4 Palaeontological and Evolutionary History of the Larger Foraminifera 1.4.1 Evolution Like all fossils, from dinoflagellates to dinosaurs, the larger benthic foraminifera are biofacies bound, and often regionally constrained. They have biotopes closely asso- ciated with carbonate environments. Large-s cale changes in these biotopes occur in response to, for example, eustatic sea-l evel fluctuations and climate change, that cause stress to the associated fauna and flora. Palaeoecological studies have demonstrated that feeding mechanisms and reproductive strategies are key traits that affect survival rates (Twitchett, 2006). Small unspecialized and opportunistic taxa fare better than large and advanced forms during times of stress, and after a major event, the surviving primitive forms thrive in the new environment, which initially is one of low diversity and limited competition for food resources. Typically, this state is associated with the predominance of small forms (the “Lilliput effect”) with very high turnover rates and low biomass. These “disaster species” can quickly take advantage of the relatively high food supplies and lack of competition (Pomar and Hallock, 2008). As environmental conditions stabilize, the survivors diversify into the new environments and eventually new larger forms evolve and colonise more specialised niches. Larger foraminifera always preserve the juvenile stage, at least in the microspheric form, before growing into an adult stage. They show considerable similarities in evo- lutionary trends, and often follow parallel lines of evolution. These similarities were first recognised by van der Vlerk (1922) and Tan Sin Hok (1932), and were later sum- marised by Haynes (1981). Larger foraminifera, as they evolve, tend to increase the size of their test, together with increasing the size of the macrospheric embryont, Biology and Evolutionary History of Larger Benthic Foraminifera 31 especially the second embryonic chamber (the deuteroconch). The increase in size is also accompanied by a tendency to increase the number of chambers, or subdivision of the chambers. However, there is also a tendency towards the reduction in embryont size with radial growth. Hallock (1985) proposed biological interpretations for these trends that include optimizing the benefits of algal symbiosis, relatively long life span, and minimizing mortality of asexually produced juveniles by increasing the size of the embryont. Advanced forms tend to have ogival (curved diamond shape) or hexagonal equatorial chamberlets. There is also a progressive development of lateral chamberlets in the rotaliides. During the evolutionary history of larger foraminifera many morphological features have shown convergent evolutionary trends. Identical gross morphologies, shapes or structures appear again and again within the same lineage or in parallel lineages from different stocks. An example of such convergence is the fusiform test, which appeared several times in different lineages. The elongate fusiform fusulinides made their first appearances in the shallow warm waters of the Carboniferous-P ermian. They evolved from simple foraminifera with calcareous granular walls, acquiring as they did so alve- olar compound walls (Chapter 2). Quite independently, and from a miliolide ancestor, fusiform elongate alveolinids made their first appearance in the Paleogene (Chapter 6), and have re-e volved many times (but showing different detailed internal features) up to the present day (Fig. 1.27). This morphological evolution occurred as the foraminifera occupied specific envi- ronmental or ecological niches. Genes are the fundamental units of life, and determine the genotypical properties of any life form. However, environment, ontogeny and con- ditions during growth determine the phenotypical character of a species, determining to some degree, for example, the shape of the test. The interaction between the gene in the embryont and the selection of the external features that the foraminifera develop during growth is an example of this selection process discussed by Gould (2002) in his book The Structure of Evolutionary Theory. This process acts on features that emerge from complex gene interaction during ontogeny and not from individual genes. One need only look at the range of sizes and shapes of domestic dogs to see how mor- phological variation can be generated by selected or controlled conditions for repro- duction, to realise how, despite the variety of genetic components, certain ecological conditions will favour certain morphological developments. The many iterative and convergent evolutionary trends found in larger foraminiferal lineages provide wonder- ful examples of the fundamental restrictions on genetic possibilities within a higher taxon, which is the basis of the concept of “bauplan” (Hintzsche, 1947) that is widely recognized in evolutionary biology and palaeontology. Many characteristics in the evolution of foraminifera are gradual and linked to the timing of phases in growth. They could have been started by DNA mutation at one stage of foraminiferal life and amplified as a consequence of having to cope with stress and adverse environments. A good example is the gradual evolution of the lateral cubiculae/c hamberlets that are created by intra-l amellar spaces and linked by pores in Miogypsinoides, which is accompanied with nepionic reduction that results in forms such as Miogypsina (see Chapters 6 and 7). Other characteristics appear suddenly, such as alveoles in the miliolide Austrotrillina, which appear to have evolved directly from 32 Evolution and Geological Significance of Larger Benthic Foraminifera e d b c a PERMO- CARB. TRIASSIC JURASSIC CRETACEOUS PALEOCENE EOCENE OLIGOCENE MIOCENE PLIOC. PLEIS. HOLO. Fig. 1.27. The convergence of similar shapes of test, but with different internal structures, shown by the fusiform test in a) the calcareous granular fusu- linines and b) the porcelaneous alveolinids; or the rotaliides which have same shape and the same internal three-l ayered structures but different shapes of embryonic apparatus and shape of chamberlets, in forms such as c) the orbitoids, d) the discocyclinids, e) the lepidocyclinids. Biology and Evolutionary History of Larger Benthic Foraminifera 33 Quinqueloculina, a small, simple benthic foraminifer. However, these characteristics had also appeared in larger miliolides earlier on in the stratigraphic column, such as the alveoles in Alveolina. Evolving from a quinqueloculine ancestor, Alveolina, did not acquire the alveoles in its ontogeny but later on in the adult stage. This suggests that the propensity to develop a mutation that allows for the development of alveoles is common to most or all orders of larger foraminifera, and that when the conditions are suitable, this feature will be selected. Larger foraminifera, such as Eulepidina and Discocyclina, achieve large sizes and radial symmetry early in their evolutionary history. Such forms having reached the optimum size with a small nepionic stage generally do not exhibit any more changes for millions of years unless subjected to environmental change. Some features seem to be essential to establish morphological trends in larger fora- minifera. An example is the appearance of a complicated system of stolons in the lat- eral layers, which allow the cytoplasm to extrude from these apertures in order to build small chamberlets. The stolon system must have resulted from an adventitious gene mutation which enabled the cytoplasm to extrude and build more chamberlets. This would also have allowed the foraminiferal test to spread and have more space to house symbionts. Tan Sin Hok (1932) and Smout (1954), whose work preceded recognition that algal symbiosis was nearly ubiquitous among extant larger foraminifera, proposed that large tests are more efficient as they provide shorter lines of communication for internal processes, thus providing a competitive advantage. The larger foraminifera, by developing alveoles, chamberlets, systems of stolons, pillars and other structures that strengthen their tests, are one of the best examples of how genetic mutation allows a wide range of stable environmental niches to be occupied. Proliferation is the reason for a species’ existence, but survival is the short- term goal. However, as for all creatures, larger more specialised forms cannot as eas- ily survive the development of adverse conditions, or adapt as readily to a different mode of life. They are therefore the first ones to become extinct. However, the nascent characteristics still exist in the genes of more primitive forms, and when a niche is re- colonized the old morphological features may reappear. Genotypical convergence gives rise to forms similar to the extinct ones, but with slight or major structural vari- ations, such as the folded septa in the fusiform fusulinides compared with the alveoles in the fusiform alveolinids (see Fig.  1.20). However, the fusulinides and miliolides are not closely related, as the first ones have calcareous granular walls and the sec- ond ones are porcelaneous with calcitic non- perforate walls. This parallel evolution from two different stocks was nonetheless driven by comparable forces, for the same purpose and benefit for the foraminifera, as both variations likely were used to house symbionts. Since Carboniferous times, changes in local or global conditions have caused the highly specialised forms of foraminifera to become extinct. Later, when environmental conditions that can support larger foraminifera are re- established, new taxa evolve and fill the restored niches. Based on the evolutionary records of larger foraminifera and other carbonate- producing organisms, Hallock (1987) proposed the concept of “geo- logically longevous” versus “geologically ephemeral” niches. Moreover, the disappear- ance and reappearance of larger foraminiferal taxa through geological time illustrates 34 Evolution and Geological Significance of Larger Benthic Foraminifera the converging trends highlighted so clearly by Conway Morris (2003). These conver- gent trends are consistent with developing similar characteristics at different times and probably under similar conditions. Larger foraminifera are, therefore, good examples of the creative powers of gene mutation and gene interaction, and can provide case studies for the important role of both genotypical and phenotypical processes. 1.4.2 Palaeontology As a consequence of their ability to evolve rapidly and fill a range of ecological niches, larger foraminifera are very valuable guides to changes in geological environments (Marriner et al., 2005; Morhange et al., 2005; Stefaniuk et al., 2005; Jones et al., 2003; Jones et al., 2002; Lord et al., 2009; Meinhold et al., 2009; Henderson et al., 2010; BouDagher- Fadel et  al., 2010c). Their study provides excellent insight into palaeo- ecology but also into the factors that give rise to both local and global extinction events. Eustatic sea level fluctuations and climate change have been two of the main mechanisms invoked by geologists to explain past extinction events. More recently, processes like volcanic eruption, anoxia or changes in atmospheric chemistry, and still more famously catastrophic meteorite impacts have been invoked. Hallock (1987) used a simple conceptual model to illustrate how significant perturbations to ocean circu- lation, whether gradual (such as climate change) or nearly instantaneous (such as a meteor impact), could eliminate high proportions of the niches occupied by taxa spe- cialized to shallow, warm, nutrient- poor habitats within the photic zone. It is certainly the case that the ecological sensitivity, or vulnerability, of the larger foraminifera cou- pled with the ability of smaller primitive forms to survive and then evolve to repopulate the niches that were left vacant, means that larger foraminifera are very good tools to study global extinction processes (Fig. 1.28). The story of the evolution and periodic extinction of larger foraminifera will be told in detail in the subsequent chapters, but is summarised briefly below. The first larger foraminifera seem to have evolved from the agglutinated foramin- ifera in the Carboniferous (see Chapter 2) by developing compound walls and a com- plicated internal structure. These lineages developed fusiform tests and gave rise to the Fusulinida and several related orders (such as the Staffellida, etc) that became very abundant rock- forming fossils. However, these Palaeozoic orders did not survive the end Permian, but some small related orders, such as the Earlandiida and the Endothyrida survived to the Early Triassic and end Triassic respectively (see Chapters 2 and 3). The Triassic was a period of recovery, larger foraminifera re-e merged slowly. During the Early Triassic, foraminifera were simple and rare. They were initially dominated by small arenaceous forms, the “disaster forms”, which are characteristic of the survival phase following a mass extinction (see Chapter 3). The recovery forms were dominated by the miliolides and later by the involutinides in the Eastern Tethys area. The end of the Triassic saw another crisis, which again killed most of the larger foraminifera, and all of the small Permian survivors were completely wiped out (see Chapter 4). Following this mass extinction, the early Jurassic witnessed the steady evo- lution of the agglutinated forms from the small simple textulariides to those with an internally complicated form, which became abundant from the Pliensbachian onwards, Biology and Evolutionary History of Larger Benthic Foraminifera 35 Ma Stages Extinction events Possible causes 5.3 End of Messinian Extinction of Heterostegina (Vlerkina ) Dessication of the Mediterranean and tectonic events. (Late Miocene) 11.6 End of Serravallian Extinction of lepidocyclinids and Tectonic events. miogypsinids in the Tethyan province. (Middle Miocene) 15.9 End of Burdigalian Extinction of lepidocyclinids and Eruption of the Columbia River miogypsinids in the American province. basalts. (Early Miocene) 23.0 End of Oligocene Extinction of Orbitolinoidea and Plate tectonic events. Orbitoidoidea 33.9 End of Eocene Extinction of the nummulitd Pellatispira, Global climate changes attributed Biplanispira to the expansion of the Antarctic ice cap and multiple bolide impact events. 37.8 End of Bartonian Extinction of Coskinolinoidea and the The move into an “ice house” nummulitid Assilina climate may have been triggered (Middle Eocene) at this time by the opening of the Tasmanian gateway. 66.0 End of Maastrichtian Extinction of Cretaceous alveolinids, Multiple impact events, orbitoids, pfenderinids, large lituolids, etc. e.g.Chixculub crater impact. (Late Cretaceous) 93.9 End of Cenomanian Extinction of the Orbitolina The Wallaby eruption and a near- peak Mesozoic eustatic sea-level (Early Cretaceous) highstand. 201.3 End of Triassic Extinction of all small fusulinids and large Central Atlantic Magmatic lagenids Province. 251.9 End Permian Extinction of all large fusulinids Siberian traps flood basalts. 259.1 End of Guadalupian Extinction of all large Schwagerinidae The Emersham basalts outcrop. (Late Permian) Fig. 1.28. Summary of the main extinction events affecting the larger benthic foraminifera through the geologic column. 36 Evolution and Geological Significance of Larger Benthic Foraminifera thereby giving the carbonate facies of the Jurassic a characteristic that is recogniza- ble throughout all of Tethys. By the end of the Jurassic some forms became extinct, but many robust forms from shallow, clear waters survived the Tithonian and crossed over to the Cretaceous. The Early Cretaceous biota as a whole was dominated by large agglutinated foraminifera with complicated structures such as alveoles, internal partitions and pillars (see Chapter 5). The complex and partitioned lituolids (of the order Textulariida, such as Orbitolina) dominated the biota of inner carbonate plat- form environments that were widespread along the western and eastern margins of the Cretaceous Tethys. They were joined by the alveolinids (of the order Miliolida), which showed spectacular expansion in the Middle Cretaceous, proliferating in mid-l atitudes, and often becoming annular or discoid and subdivided by partitions. Many of them resembled the planispiral-f usiform fusulinides of the Permian, attaining approximately the same range of sizes, but differing fundamentally in their imperforate, porcelaneous wall structure. In the Late Cretaceous new simple rotaliides forms evolved into forms with complicated three-l ayered textures, the orbitoids. While the previous two groups had their main breeding ground in the Tethyan realm, the orbitoids showed provincial- ism and some were only found in the Caribbean. As will be seen in Chapter 5, the Cretaceous– Tertiary crisis wiped out most of the Maastrichtian larger benthic foraminifera. The Early Paleocene was a recov- ery period and only by the Late Paleocene had larger miliolides, nummulitids and orthophragminids appeared and spread throughout Tethys. The miliolides included large fusiform alveolinids, which showed morphological convergence with the extinct Cretaceous alveolinids, and the discoid soritids, which became prominent throughout the Eocene. Parallel to these lines of evolution, the agglutinated fora- minifera developed into forms imitating their Cretaceous ancestors by developing internal pillars (as in the textulariides), or with more complicated partitions (as in the orbitolinids). However, in the American province the rotaliides evolved into three- layered lepidocyclinids. Towards the end of the Paleogene, the extinction of larger foraminifera was not as pronounced, however fluctuations in climate, sea-l evel and/ or oceanic currents influ- enced the geographic distribution of the larger benthic foraminifera, causing biogeo- graphical provincialism. Although most of the Miocene superfamilies are still extant, provincialism was prominent at generic and specific levels. The Late Oligocene and Early Miocene show the least provincialism, with miogypsinids and lepidocyclinids spreading from the Americas to the west (see Chapter 6). By the Middle Miocene, with the closure of Tethys, provincialism was re-e stablished. Lepidocyclinids and miogy- psinids completely disappeared from America in the latest Early Miocene and from the Mediterranean in the Serravallian (late Middle Miocene). Deep- water textulariides made their first appearance in America, while new genera of alveolinids appeared in the Indo- Pacific. The development of the Indo- Pacific as a separate province continued in the Late Miocene with the closure of the Tethyan seaway. Neogene sedimentary sequences (Chapter 7) of this province are dominated by warm-w ater, shallow- marine carbonates of crucial importance as the product and record of climatic/o ceanic conditions and as hydrocarbon reservoirs. Throughout the past century, larger foraminifera have been Biology and Evolutionary History of Larger Benthic Foraminifera 37 used extensively in the Far East from Japan southwards to Australia, and through the Tethyan region westward to the Caribbean and America for biostratigraphical, palae- oceanographical and palaeoclimatological analysis. In the Middle East, larger fora- minifera have been used to correlate the many biostratigraphical stages, which have been proposed for Europe. Their role as markers for biostratigraphical zonation and correlation underpins most of the drilling of marine sedimentary sequences that is cen- tral to hydrocarbon exploration. By the Holocene, new forms of larger benthic foraminifera appeared through the tropical realm, especially in the Indo- Pacific. Present day larger foraminifera play anal- ogous roles in the ecosystem, where the tropical belt is divided into two parts, and differ- ent assemblages colonise different environments. In particular, the Indo- Pacific larger soritids (Amphisorus and Marginopora) are substituted in the Caribbean by Archaias and Cyclorbiculina as porcelaneous discoidal epiphytes on tropical seagrasses (Langer and Hottinger, 2000). Modern tropical western- Pacific sandy shoals and beaches can be composed of nearly pure concentrations of Calcarina, Amphistegina, or Baculogypsina tests (see Fig. 1.26), though only Amphistegina spp., which are directly descended from a Paleogene ancestor, are abundant across the subtropical- tropical regions worldwide. 1.5 Conclusion From this introductory chapter, it can be seen that larger foraminifera are biologi- cally complex and highly versatile. They have repeatedly evolved from simple ances- tors since the Carboniferous, becoming highly specialised and therefore highly sensitive to environmental changes. Their study provides, as a result, considerable insight into evolutionary process as well as into the major geological mechanisms associated with extinction and recovery. Larger foraminifera also have, and continue to occupy, the very important ecolog- ical niche of being a reef-f orming group. The worldwide distribution of carbonate biota, especially reef biota, contains important information on environmental condi- tions, including oceanographic parameters, that control this most sensitive of habitats. The study of the distribution patterns of this biome, over different time slices, provides valuable information on how the climate of the Earth has evolved in the past 350 Ma. Finally, the carbonate rich shallow marine and reef environments favoured by larger foraminifera are also those which may give rise to economically vital deposits of oil and gas. So larger foraminifera are now central to our ability to date, correlate and analyse the sedimentary basis that are currently key to the economic wellbeing of the world. A detailed understanding of the taxonomy of the larger foraminifera is essen- tial, therefore, for any applied biostratigraphic analysis. In this book, each of the following chapters outlines the palaeobiological and the geological significance of the larger foraminifera through time. Specifically, the tax- onomy, phylogenetic evolution, palaeoecology and biogeography of the larger fora- minifera are outlined and discussed, relative to the biostratigraphical time scale (as defined by Cohen et al., 2013) of the middle and late Phanerozoic. In establishing the most suitable markers for the biozonal boundaries, new (bio)chronostratigraphic units 38 Evolution and Geological Significance of Larger Benthic Foraminifera are established. The larger benthic foraminiferal of the ‘letter stages’ of the Far East, as defined by BouDagher- Fadel and Banner (1999) and BouDagher-F adel (2008), and the shallow benthic zones (SBZ) for the Paleocene–E ocene epochs as proposed by Serra- Kiel et al. (1998) are revised and correlated to the planktonic foraminiferal zonal scheme of BouDagher- Fadel (2015). Many of the figures presented here are type figures from the Natural History Museum, London (referred to as NHM), while other types, such as those from the Permian are from the Senckenberg- Forschungsinstitut und Naturmuseum, Germany (referred to as SFN), others are deposited in the UCL Collections, while some are taken from referenced literature sources. 1 2 G G 3 4 5 Plate 1.1 Scale bars: Fig. 1 = 0.5mm; Fig. 2 = 50μm; Fig. 3 = 0.170μm; Fig. 4 = 0.038μm Fig. 1. A fig- ure showing the nuclear envelope (Ne), of the nucleus which is situated in the sixth youngest chamber of Elphidium excavatum (Terquem), enclosing the nucleoplasm (Nu) which contain nucleoli (n) and dense chro- matin granules (cg) (after Alexander, 1985). Fig. 2. Enlargement of the nucleus in Fig. 1, showing the nuclear envelope to consist of a perinuclear space (Pn) separating the outer envelope (oe) from the inner envelope. Abbreviations: pc = peripheral chromatin, nc = nucleus- associated chromatin (after Alexander, 1985). Fig. 3. Longitudinal section of fine pseudopodia of Haynesina germanise (Ehrenberg). Microtubules are visible in the central region. G = granules. Fig. 4. High magnification detail of a pseudopod from Elphidium wil- liamsoni Haynes showing the membrane-l ined canal system (cs). Fig. 5. Enlargement of the cytoplasm of Elphidium williamsoni Haynes to show close association of the golgi (G) with vacuoles containing chloro- plasts (C), and vacuolar lumen (L). 2 1 V Va M Va F 4 3 U S Sc B A 5 6 Plate 1.2 Scale bars: Figs 1, 2, 5-6  = 50μm; Figs 3-4  = 25μm Fig. 1. Ammonia batava (Hofker). SEM pho- tograph of dorsal side a partial resin cast of an embedded specimen. Resin has filled the umbilical fissure (Uf) and also vertical canals (arrow) extending from the umbilical region towards the dorsal test surface. L = chamber lumina of outer whorl; LI = lumen of inner whorl chamber (after Alexander, 1985). Fig. 2. Equatorial section of Ammonia batava (Hofker). Y = youngest chamber; P = proloculus; sa = septal aper- ture (after Alexander, 1985). Figs 3-4 . Detail of penultimate chamber in two different planes of section of Ammonia batava (Hofker, 1951). Large voids (V) occupy much of the cytoplasm and small spherical vacuoles (Va) are present. Diatoms (arrows), mineral- like fragments (M) and other ingested material such as filament structures (F) are also present (after Alexander, 1985). Fig. 5. Equatorial apertural view of a megalospheric form of Elphidium williamsoni Haynes. Last formed chamber removed. A = apertural openings located on a retral process of the previous whorl; B = apertural opening formed above a fosette (the opening to the exte- rior of an intraseptal interlocular space) of the previous whorl; S = “septal (rotaliine) flap”; Sc = intraseptal canal; U  =  umbilical lobe of chamber lumen, opening to the umbilicus (after Alexandra, 1985). Fig.  6. Equatorial section of Elphidium excavatum (Terquem) showing an increased density of cytoplasm within the proloculus (Pr) when compared to the sixth youngest chamber (CH6), which contains the nucleus (N), Pa = primary aperture of youngest chamber (y), Sa = septal aperture between youngest chamber and penul- timate chamber (Pc) (after Alexander, 1985). 1 2 3 4 5 6 Plate 1.3 All photographs are from living specimens. Scale bars:  Fig.  1  =  1mm; Figs 2-4   =  2mm. Fig.  1. Heterostegina depressa d’Orbigny, the large individual is an agamont (diameter 14.50 mm), the small individual a gamont (diameter 3.75 mm). In many cases, the youngest chambers of agamonts are partly broken off, and older parts of the test show signs of former damage, now healed and overgrown by later chambers. The pseudopodia protrude from the canal system of the marginal cord. The living laboratory specimens from a dredge haul from 40 m water depth off Kekaa Point, Island of Maui, Hawaii (courtesy of Prof R. Röttger). Fig. 2. Larger fora- minifera in their natural habitat, the protective algal tangle of a rock-p ool at Makapuu Point, Island of Oahu, Hawaii. In the centre, Heterostegina depressa d’Orbigny (size 2.4 mm); top left Amphistegina lessonii d’Orbigny, behind it Aphistegina lobifera Larsen. In the lower right corner another Heterostegina depressa d’Orbigny can be seen. These species of larger foraminifera obtain their characteristic yellowish coloration from symbiotic diatoms which contribute to the nutrition of their hosts, (courtesy of Prof R. Röttger). Fig. 3. Amphistegina lessonii d’Orbigny (test size 1.57 mm), two days after the reproductive process. Specimen from the natural hab- itat, Hawaii, photographed in the laboratory. The now colourless inanimate mother cell is surrounded by its 670 daughter individuals formed by multiple fission of the parent protoplasm. Most of the offspring are in the 3-c hamber stage. The majority of symbiotic diatoms transferred from the mother individual to the juveniles are located in the second and third chamber, (courtesy of Prof R. Röttger). Fig. 4. Heterostegina depressa d’Orbigny (test size 3.5 mm), specimen from off Pearl Harbour, Hawaii. Transmitted light reveals the planispiral test con- struction (courtesy of Prof R. Röttger). Figs 5, 6. Heterostegina depressa d’Orbigny. Microspheric agamonts after multiple fission. 5). The protoplasm of the specimen depicted in Fig. 1 has divided into ca. 1300 gamonts now in the 2-c hambered embryonic stage (size 155µm). 6). Some hours later, the juveniles being transported off the inanimate mother test by streams of residual protoplasm, (courtesy of Prof R. Röttger). Plate 1.4 All photographs are from living specimens. Scale bars: Fig. 1 = 2mm. Figs 2-6  = 1mm. Fig. 1. Amphistegina radiata Fichtel and Moll (test size 1.93 mm). Specimen from Kudaka Island, Ryukyu Islands, Japan (courtesy of Prof R. Röttger). Fig. 2. Calcarina gaudichaudii d’Orbigny, Agamont (diameter 3.5 mm). From the trochospiral test branched processes (spines) with a protoplasm-c ontaining internal canal system radiate. On the side facing the observer the spiral of chambers is covered by a thin- walled brood chamber within which the many juveniles (the young gamonts) are formed. They are released by dissolution of the roof of this chamber. Specimen from Belau, Micronesia (courtesy of Prof R. Röttger). Fig. 3. Calcarina gaudichau- dii d’Orbigny, Gamont (test size 2.8 mm). Gamonts are smaller than agamonts. Also the number of spines, now unbranched, is much smaller than in agamonts. Specimen from Kudaka Island, Ryukyu Islands, Japan, (cour- tesy of Prof R. Röttger). Fig. 4. Baculogypsina sphaerulata Parker and Jones (test size 2.2 mm). The subspheri- cal test differs from other calcarinids; only the innermost part, not visible from outside, forms a spiral. All other chambers are inserted between the axes of the spines according to a different pattern of construction. The net- work of chambers is interrupted by transparent pillars (visible as dark spots) oriented perpendicular to the test surface. Specimen from Kudaka Island, Ryukyu Islands, Japan (courtesy of Prof R. Röttger). Fig. 5. Calcarina gaudichaudii d’Orbigny and one Baculogypsina sphaerulata Parker and Jones (lower left corner) attached to a red algal thallus from their natural habitat, the littoral algal zone of Kudaka Island (Ryukyu Islands, Japan). Attachment occurs with an organic cement secreted from the tip of the canaliferous spines. This device helps to counteract transport by wave action (size of largest specimen 3.8 mm) (courtesy of Prof R. Röttger). Fig. 6. Peneroplis sp., Peneroplidae (order Miliolida) (test size 1.5 mm) has formed daughter individuals by multiple fission. The test was partly dissolved during this process to release the juveniles. Peneroplis harbours symbiotic unicellular red algae which were transferred to the offspring (courtesy of Prof R. Röttger). 2 6 1 3 4 5 6 45 Chapter 2 The Palaeozoic Larger Benthic Foraminifera 2.1 Introduction The first foraminifera with a hard test (i.e. one that was biomineralised and there- fore had preservation potential) to appear in the fossil record are the unilocular, sim- ple agglutinated Allogromiida. From these the calcareous agglutinated foraminifera, Textulariida, evolved during the Cambrian. The members of the Textulariida remained the dominant (but morphologically small) group in the Early Palaeozoic, but mor- phologically larger foraminifera with compound, microgranular walls and a com- plicated internal structure, exemplified by the order Fusulinida (which are generally referred to here as fusulinides), became abundant and ecologically dominant in the Late Palaeozoic. These larger fusulinides and related forms are the major fossil group in many Late Palaeozoic shallow- marine limestones. In addition to the textulariides and the fusulinides related forms, the lagenides fora- minifera evolved in the Silurian, but did not show significant evolutionary diversity until the late Gzhelian in the Carboniferous, becoming more abundant in the Permian. It appears likely that their defining morphological characteristic (having walls made of orientated calcite crystals) evolved independently several times in the Palaeozoic, and that the superfamilies Robuloidoidea and Nodosinelloidea had different phylogenetic roots. However, these forms always remained morphologically relatively small, and it was not until the Triassic that significant, larger lagenides appeared. The morphologi- cal evolution of this group is discussed in detail in Chapter 3, and here the Palaeozoic genera are only briefly discussed to enable the continuity between these forms and those of the later Triassic to be understood. The miliolides (with porcelaneous test walls) were also present in the Palaeozoic, but again (as with the lagenides) they remained morphologically small and primi- tive throughout the Palaeozoic. It was not before the Triassic that they became rela- tively large and more prominent as rock-f orming fossils (see Chapter 3). Whereas, true Involutinida really first appeared in the Triassic:  their only suggested occurrences in the Permian were erroneous, based on the misidentification of the porcelaneous genera Neohemigordius and Pseudovidalina (Loeblich and Tappan, 1988). As stated, the Palaeozoic Fusulinida and related forms first evolved from the Allogromiida in the Silurian (see Fig. 2.1), with the appearance of members of the Parathuramminida, followed directly by the Moravamminida (a disputed forami- niferal order, see Vachard, 1994; Vachard and Cozar, 2010), the Archaediscida, and the Lagenida referred to above. From the Devonian onwards fusulinides and related forms became established and included, the Earlandiida, the Palaeotextulariida, the Tetrataxida, the Tournayellida, the Endothyrida, and the Fusulinida. 46 Evolution and Geological Significance of Larger Benthic Foraminifera Fig. 2.1. Evolution of the Palaeozoic larger foraminifera. During the latter part of the Palaeozoic, the fusulinides and related forms evolved from being simple to highly- specialized, and are characterised by a history of rapid evolution that gave rise to diverse lineages. They were sensitive to their physical envi- ronment, and were the first foraminifera to develop a test with a complex internal architecture, combined with, for a single cell organism, a relatively gigantic, centimetric size. The fusulinides and related forms became cosmopolitan, colonising most shallow, warm waters of the Carboniferous and Permian. All (except members of the superfam- ily Fusulinoidea) occurred in temperate waters, and in shallow to deeper conditions. They are well preserved, but their preservation mode depends on post- sedimentary processes. In many areas, during the Late Carboniferous and Permian, the fusulinides far outnumber any other single marine invertebrate group. The Palaeozoic foraminifera (excluding the lagenides) comprises more than 667 distinct genera, not including synonyms (see Loeblich and Tappan, 1988; Vdovenko et al., 1993; Rauser-C hernousova et al., 1996). In China alone, more than 3395 species belonging to 92 genera of the Fusulinida have so far been described (Jin-Z hang, 1990). The lagenides and members of the other Palaeozoic orders were minute, mea- suring only 0.4 to 1.3 mm in diameter. Many of the advanced fusulinides, however, became morphologically large, reaching 1 to 2 cm in length, while some of the Permian forms reached 15  cm in length. These are amongst the largest foraminifera ever to have existed (Douglass, 1977). The larger forms are an exclusively Late Palaeozoic group that achieved their peak diversities during the Visean stage of the Carboniferous and the Cisuralian epoch of the Permian. However, at the end of the Palaeozoic, just as they began to achieve their largest sizes, with extremely complicated internal struc- tures, these forms became extinct. The Palaeozoic Larger Benthic Foraminifera 47 The taxonomic ranking and division of the fusulinides is controversial, and has been subjected to considerable change over time. In 1937, Wedekind first introduced the name superfamily Fusulinacea in order to represent all the widely spread unique Paleozoic forms. Fursenko (1958) subsequently proposed ranking the fusulinides as a separate order, which was later accepted in Fundamentals of Paleontology (Rauser- Chernousova and Fursenko, 1959; Haynes, 1981) and other works of Soviet micropaleontologists. At the same time, in the best known and widely used classification developed by Loeblich and Tappan (1964; 1980; 1987), the fusulinides were regarded as a suborder, and only in 1992 was their rank raised to that of order (Loeblich and Tappan, 1992). However, in the meantime Mikhalevich (1980) had raised the fusulinides to the rank of a superorder, including within it the orders Endothyrida and Fusulinida. The system of superorder was further developed by a team of Russian micropaleontologists headed by Rauser- Chernousova (Rauser- Chernousova et al., 1996), and additionally included the order Tournayellida. Most recently, Vachard et al. (2010) assigned the early foraminifera to the class Fusulinata, and then subdivided this class into six orders: Parathuramminida, Archaediscida and Earlandiida (forming together the subclass Afusulinana n. subcl.), and Tournayellida, Endothyrida and Fusulinida (subclass Fusulinana nom. translat.). In this chapter, various views are referenced, and an approach is presented which attempts to provide a self-c onsistent and evidenced based taxonomic and phylogenetic classification. The approach presented here is intended to resolve some of the confusion and complexity surrounding the larger benthic Palaeozoic foraminiferal taxonomy. To elucidate their evolutionally trajectory and to guide their classification, previous work (e.g. Delage and Hérouard, 1896; Hohenegger and Piller, 1973; Mikhalevich, 1980; Leven, 2009; Davydov, 2011; Mikhalevich, 2013)  is built upon and synthesized. The wall composition and ultrastructure, the presence or absence of the chomata, and the degree of septal folding in the test are used to define phylogenetic relationships and relative ranking (see Fig. 2.1). It is acknowledged, however, that this is an active area of research which may well develop alternative interpretations in the future as the evi- dence base grows. In this the revised approach six orders are separated from the Fusulinida, namely the Parathuramminida, the Palaeotextulariida, the Archaediscida, the Tetrataxida, the Tournayellida and the Endothyrida (see below for phylogenetic description). Thus, in the next section of this chapter the main features of the taxonomy of the following orders are presented: • Parathuramminida • Moravamminida • Archaediscida • Earlandiida • Palaeotextulariida • Tetrataxida • Tournayellida • Endothyrida • Fusulinida • Lagenida • Miliolida 48 Evolution and Geological Significance of Larger Benthic Foraminifera This is followed by a discussion of the biostratigraphic significance and phylogenetic evolution of the geologically most important Palaeozoic forms, and the chapter con- cludes with a review of their palaeoecological significance and their palaeogeographic distribution during the Palaeozoic. 2.2 Morphology and Taxonomy Of Palaeozoic Larger Benthic Foraminifera ORDER PARATHURAMMINIDA MIKHALEVICH 1980 In this order all forms are unilocular, large globular or tubular, and occurred both as free and as attached forms. The wall is thin, calcareous microgranular, simple to bila- mellar with an inner hyaline pseudofibrous layer. Apertures are terminal at the top of a hollow neck. Silurian to Permian. Superfamily PARATHURAMMINOIDEA Bykova, 1955 In this superfamily all forms are unilocular, globular or tubular, and occurred both as free and as attached forms. The wall structure is simple calcareous microgranular or with granulo-fi brous layers (see Fig. 2.2). They range from Early Silurian to Permian. Family Archaesphaeridae Malakhova, 1966 The representatives of this family mainly have one or more globular to elongate chamber(s), with a test with no apparent aperture. They range from Late Silurian to Early Permian. • Archaesphaera Suleymanov, 1945 (Type species:  Archaesphaera minima Suleymanov, 1945). The test is globular and smooth. Devonian to Carboniferous (Tournaisian), (Fig. 2.2). • Diplosphaerina Derville, 1952 (Type species: Diplosphaera inaequalis Derville, 1931). The two-c hambered test, has a small proloculus enveloped within a larger chamber. The wall is dark, granular, nonperforated and single layered. Middle Devonian to Carboniferous (Tournaisian), (Plate 2.1, figs 15, 16; Plate 2.2, Fig. 10; Fig. 2.3). Family Parathuramminidae Bykova, 1955 The representatives of the Parathuramminidae include globular, or irregular forms with a simple calcareous agglutinated wall and multiple apertures at the end of a tubu- lar neck. They range from Silurian to Carboniferous (Mississippian) and they are the ancestral group of all the fusulinids. • Parathurammina Suleymanov, 1945 (Type species:  Parathurammina dagmarae Suleymanov, 1945). The test is globular, with apertures at the ends of numerous tubu- lar protuberances. Late Silurian to Carboniferous (Tournaisian) (Plate 2.2, fig.19). Family Chrysothuramminidae Loeblich and Tappan, 1988 Members of this family have a globular to irregular test in outline, with the aperture at the end of tubular projections. They range from Middle Devonian to Carboniferous (Tournaisian). The Palaeozoic Larger Benthic Foraminifera 49 • Chrysothurammina Neumann, Pozaryska and Vachard, 1975. (Type species: Chrysothurammina tenuis Neumann, Pozaryska and Vachard, 1975). The test is com- posed of a single subspherical chamber with an aperture on a neck- like protrusion. Middle Devonian to Carboniferous (Tournaisian) (Plate 2.3, Fig. 1B). Free forms Attached forms Free forms Semitextulariidae Dev. - Carb. Colaniella FUSULINOIDEA Colaniellidae Ptychocladia L. Carb. - Perm. L..Dev. - Perm. Ptychocladiidae L. Dev. - Carb. Palaeotexulariidae Carb. - Perm. Geinitzinidae Biseriamminidae L.Dev.-Permian Carb. - Perm. Tetrataxidae Endothyridae Carb. Tuberitinidae Dev. - L. Carb.. Sil.-Perm. ? Semitextulariidae LAGENIDA Dev. - Carb. Howchinia Lasiodiscus Nodosinella Permian Lasiodiscidae Carb.-Perm. Archaediscidae Tournayellidaegranular Carb. L. Dev-Carb. fibrous Nodosinellidae or granulo- Umbellinidae L.Sil.-Perm. fibrous Sil. Dev. Earlandinita Wall compound Wall simple Earlandinitidae calcareous L.Dev.- Carb.MORAVAMMININOIDEA granular L.. Sil. - Tourn. NODOSINELLOIDEA ? Parathuramminidae Paracaligella agglutinated E. Sil.-E. Carb. layer CaligellidaePARATHURAMMINOIDEA L. Sil. - E. Carb. Earlandiidae L. Sil. - E. Triass. Allogromiida Archaesphaera EARLANDIOIDEA Fig.  2.2. Schematic morphological evolution of the fusulinides and lagenides from the Allogromiida. Figures are not to scale. Not all superfamilies and families discussed in the text are shown on this figure and not all superfamilies names are given (for reasons of simplicity and clarity). Chamber lumina complex, NON- SEPTATE SUBSEPTATE SEPTATE divided 50 Evolution and Geological Significance of Larger Benthic Foraminifera Middle Devonian to Permian Visean to Moscovian Bashkirian to Lopingian Eotuberina microgranular and finely perforate wall Paratuberina Tuberina microgranular, coarsely perforate wall microgranular, thick, finely perforate wall Middle Devonian to Carboniferous (Visean) Middle Devonian to Carboniferous (Tournasian) Diplosphaerina Tubeporina Dark, granular, nonperforated and single layered wall coarsely perforate three-layered wall Fig. 2.3. Schematic morphological evolution of the tuberitinids. Family Ivanovellidae Chuvashov and Yuferev, 1984 The representatives of this family are spherical with a thick calcareous wall with two layers, a thin dark compact granular inner layer and a thick grey radially fibrous outer layer. This family ranges from Late Silurian to Early Carboniferous (Tournaisian). • Elenella Pronina, 1969 (Type species:  Neoarchaesphaera (Elenella) multispinosa Pronina, 1969). The outer surface has tiny projections. There is no aperture. Late Silurian to Carboniferous (Tournaisian) (Plate 2.3, Fig. 2B). Family Marginaridae Loeblich and Tappan, 1986 The members of this family are globular in shape, with a wall made of three layers, the inner and outer ones are dark and the median one light grey, with canals running through it. These canals open into the chamber cavity and make conical projections outside the wall. This family ranges from the Middle to Late Devonian. • Marginara Pertova, 1984 (Type species:  Parathurammina tamarae Petrova, 1981). The test is spherical, with some projections from the surface with canals opening through the projections. Middle Devonian. Family Uralinellidae Chuvashov, Yuferev and Zadorozhnyy, 1984 The Uralinellidae are small, unilocular forms with a three- layered wall, the median layer is thick and clear and the surrounding layers dark. A multiple apertural neck is projected from the test. They range from Middle Devonian (Givetian) to Early Carboniferous (Visean). • Sogdianina Saltovskaya, 1973 (Type species: Sogdianina angulata Saltovskaya, 1973). Similar to Elenella, but with a thicker wall and a very prominent apertural necks. Carboniferous (Visean) (Plate 2.3, Fig. 3). The Palaeozoic Larger Benthic Foraminifera 51 • Uralinella Bykova, 1952 (Type species: Uralinella bicamerata Bykova, 1952). The test is subglobular with neck-l ike tubular projections. Middle Devonian to Carboniferous (Givetian to Tournaisian). Family Auroriidae Loeblich and Tappan, 1986 The Auroriidae have an ovate test with a two-l ayered wall, a thin inner finely porous dark layer and a thick outer canaliculated dark layer. They range from Middle to Late Devonian. • Auroria Poyarkov, 1969 (Type species: Auroria singularis Poyarkov, 1969). The test is irregularly globular. Middle to Late Devonian (Plate 2.3, Fig. 4B). Family Usloniidae Miklukho- Maklay, 1963 Members of this family have a globular test with no distinct aperture. They range from Middle Devonian to Early Carboniferous. • Bisphaera Birina, 1948 (Type species: Bisphaera malevkensis Birina, 1948). The test is subglobular, with a constriction. Middle Devonian to Carboniferous (Givetian to Tournaisian). • Parphia Miklukho- Maklay, 1965 (Type species: Cribrosphaeroides (Parphia) robusta Miklukho- Maklay, 1965). The test is spherical to ovate with a wall that is dark and evenly porous. Late Devonian (Plate 2.3, Fig. 5). Family Eovolutinidae Loeblich and Tappan, 1986 The representatives of this family have a globular test with a micro- granular wall and a distinct single aperture. They range from Late Silurian to Late Devonian. • Eovolutina Antropov, 1950 (Type species: Eovolutina elementa Antropov, 1950). The test is tiny, with the proloculus completely surrounded by the second chamber. Late Silurian to Late Devonian (Plate 2.3, fig.6). Family Tuberitinidae Miklukho- Maklay, 1958 Members of the Tuberitinidae have a simple attached test, made of one or more glob- ular chamber with calcareous microgranular walls (see Figs 2.2, 2.3). They occur between the Silurian and the Permian. • Draffania Cummings, 1957 (Type species: Draffania biloba Cummings, 1957). The test is flask- shaped with an aperture at the end of an elongate neck. Carboniferous (Tournaisian to Visean) (Plate 2.1, figs 6, 10, 13, 17, 18; Plate 2.2, Fig. 16; Plate 2.3, Fig. 5B). • Eotuberitina Miklukho-M aklay, 1958 (Type species:  Eotuberitina reitlingerae Miklukho- Maklay, 1958). The test is hemispherical with a basal disc, and with a wall that is microgranular and finely perforate. Middle Devonian to Permian (Lopingian) (Plate 2.2, figs 6, 8, 9; Fig. 2.3). • Paratuberitina Miklukho- Maklay, 1957 (Type species: Tuberitina collosa Reytlinger, 1950). The test is highly hemispherical with a basal disk, with a microgranular, coarsely perforate wall. Carboniferous (Visean to Moscovian) (Plate 2.2, Fig.  17; Fig. 2.3). 52 Evolution and Geological Significance of Larger Benthic Foraminifera • Tubeporina Pronina, 1960 (Type species: Tubeporina gloriosa Pronina, 1960). The test is hemispherical with a basal disc, and with a coarsely perforate three- layered wall, a hyaline layer between two microgranular layers. Middle Devonian to Carboniferous (Visean) (Plate 2.2, fig.7; Fig. 2.3). • Tuberitina Galloway and Harlton, 1928 (Type species: Tuberitina bulbacea Galloway and Harlton, 1928). Shows bulbous chambers in an arcuate or straight series. The wall is thick and finely perforate. Late Carboniferous to Late Permian (Bashkirian to Lopingian) (Plate 2.2, Fig. 11; Plate 2.3, Fig. 8). ORDER MORAVAMMINIDA POKORNY, 1951 Members of this order have an attached tubular test, irregularly septate. They may also have closer affinity to algae than foraminifera (Vachard, 1994; Vachard and Cozar, 2010). Late Silurian to Early Carboniferous (Tournaisian). Superfamily MORAVAMMINOIDEA Pokorny, 1951 This superfamily (Fig. 2.2) had simple free or attached tests, consisting of a proloculus and a rectilinear second chamber that was subseptate (i.e. with partial septa). They had simple walls and evolved from the Parathuramminoidea in the Silurian, but died out in the Early Carboniferous (Tournaisian). Family Moravamminidae Pokorny, 1951 In this possible foraminiferal family, the test is attached and irregularly septate with an enrolled rectilinear part. They occurred in the Middle to Late Devonian. ORDER ARCHAEDISCIDA POJARKOV AND SKVORTSOV 1979 The members of this order have free discoidal, lenticular to conical involute or rarely evolute test. It is composed of a proloculus followed directly by a planispiral tro- chospirally, or streptospirally tubular second chamber. The wall is calcareous, bilay- ered formed of an inner dark microgranular layer and a hyaline and radially fibrous outer layer. They range from the Early Carboniferous to the Late Permian (Visean to Lopingian). Superfamily ARCHAEDISCOIDEA Cushman, 1928 This superfamily is characterized by having a small, free test, consisting of a proloculus followed by an enrolled chamber. The wall is formed of one or more layers. They range from the Early Carboniferous to the Late Carboniferous (Visean to Moscovian), or unilayered and pseudofibrous. Family Archaediscidae Cushman, 1928 The Archaediscidae (Fig. 2.2) have a discoidal test, composed of a proloculus followed by a streptospirally enrolled second chamber. The wall is formed of a dark inner gran- ular layer and a clear radially fibrous or granulo-fi brous outer layer. They range from the Early Carboniferous (Visean) to the Late Carboniferous (Moscovian). • Archaediscus Brady, 1873 (Type species: Archaediscus karreri Brady, 1873). The test is free, streptospiral, with an undivided tubular second chamber and thickened wall. The Palaeozoic Larger Benthic Foraminifera 53 Carboniferous (Visean to Moscovian) (Plate 2.3, Fig. 6; Plate 2.4, Fig. 17C; Fig. 2.5; Plate 2.5, Fig. 18B; Plate 2.6, figs 3- 16; Plate 2.7, Fig. 6). • Asteroachaediscus Miklukho- Maklay, 1956 (Type species: Archaediscus baschkiricus Krestovnikov and Teodorovitch, 1936). The test is small, lenticular with sigmoi- dal coiling, stellate and with occluded chamber cavity. Carboniferous (Visean to Moscovian) (Plate 2.8, Fig. 12A; Plate 2.7, Fig. 3). • Glomodiscus Malakhova, 1973 (Type species:  Glomodiscus biarmicus Malakhova, 1973). The test is discoidal, with a globular proloculus followed by an involutely coiled, undivided tubular second chamber. Carboniferous (Visean). • Hemiarchaediscus Miklukho-M aklay, 1957 (Type species: Hemiarchaediscus planus Miklukho- Maklay, 1957). The test is small, discoidal with a glomospiral early whorl, later becoming planispiral; chamber lumen are always open. Carboniferous (late Visean to early Serpukhovian) (Plate 2.7, figs 1, 2). • Permodiscus Dutkevich, 1948 (Type species:  Permodiscus vetustus Dutkevich, in Chernysheva, 1948). The test is planispirally enrolled with a tubular undivided sec- ond chamber, later whorls have small nodosities partially filling the chamber lumen. Carboniferous (Visean to Bashkirian). • Planoarchaediscus Miklukho-M aklay, 1957 (Type species: Archaediscus spirillinoides Rauzer- Chernousova, 1948). The test is small, discoidal with a glomospiral early stage, followed by a planispiral adult. Chamber lumen remain open. Carboniferous (Visean to Serpukhovian) (Plate 2.7, Fig. 5). • Propermodiscus Miklukho-M aklay, 1953 (Type species:  Hemigordius ulmeri Mikhaylov, 1939). The test is small, lenticular with glomospiral early whorls and a planispiral, involute adult. Carboniferous (Visean) (Plate 2.7, Fig. 4). Superfamily LASIODISCOIDEA Reitlinger in Vdovenko et al., 1993 The members of this superfamily have a bilayered wall with finely granular dark inner layer and a radially fibrous outer layer. Tubercles of pseudofibrous fillings or pillars fill the umbilical region. The aperture is simple and terminal, but additional supplemen- tary apertures may occur along the spiral sutures of the successive whorls. The range is between the Early Carboniferous (Visean) and the Permian. Family Lasiodiscidae Reytlinger, 1956 In this family, the test is discoidal to conical, composed of a proloculus and an undi- vided enrolled tubular second chamber. The radial fibrous outer layer is mainly con- centrated in the umbilical region, where it may form pillars, or a series of tubercles on the surface of one side of the test. The range is between the Early Carboniferous (Visean) and the Permian. • Eolasiodiscus Reytlinger, 1956 (Type species: Eolasiodiscus donbassicus Reytlinger, 1956). The test is discoidal and concavo- convex, with the umbilical filling on the concave side. Middle to Late Carboniferous. • Glomotrocholina Nikitina, 1977 (Type species: Glomotrocholina pojarkovi Nikitina, 1977). The test is conical. Second chamber is initially streptospirally enrolled, later forming an irregular trochospiral filled with shell material. Late Permian (Lopingian). 54 Evolution and Geological Significance of Larger Benthic Foraminifera • Howchinia Cushman, 1927 (Type species: Patellina bradyana Howchin, 1888). The test is conical with septal bridges. The umbilical region is filled with fibrous calcite forming pillars (Fig. 2.2). Carboniferous (Visean to Moscovian) (Fig. 2.2; Fig. 2.22; Plate 2.1, Fig. 11A; Plate 2.5, Fig. 1; Plate 2.10, figs 1-7 , 13- 14; Plate 2.14, figs 1- 7, 13-1 4). • Lasiodiscus Reichel, 1946 (Type species: Lasiodiscus granifer Reichel, 1946). The test is planispiral with tubular extensions. Middle Carboniferous to Late Permian (Fig. 2.2). • Lasiotrochus Reichel, 1946 (Type species: Lasiotrochus tatoiensis Reichel, 1946). The test is conical with tubular chamberlets curved towards the proloculus, with distinct pillars filling the centre of the test. Permian (Cisuralian to Lopingian). • Monotaxinoides Brazhnikova and Yartseva, 1956 (Type species:  Monotaxinoides transitorius Brazhnikova and Yartseva, 1956). The test is low conical to nearly pla- nispiral, with a spherical proloculus followed by a semi- cylindrical second chamber. Carboniferous. ORDER EARLANDIIDA SABIROV IN VDOVENKO ET AL., 1993 EMEND. VACHARD ET AL., 2010 This order had simple free or attached tests, consisting of a proloculus and a rectilinear second chamber. Aperture terminal, simple. Late Silurian to Early Triassic. Superfamily EARLANDIOIDEA Cummings, 1955 This superfamily is characterised by having a free, non- septate test with a globular first chamber and a straight tubular second chamber. Members range from the Late Silurian to Early Triassic. Family Earlandiidae Cummings, 1955 The Earlandiidae have a single free chamber (Fig. 2.2) and range from the Late Silurian to Early Triassic. • Aeolisaccus Elliott, 1958 (Type species: Aeolisaccus dunningtoni Elliott, 1958). Free test, elongate and slightly tapering. Late Permian (Lopingian). • Earlandia Plummer, 1930 (Type species: Earlandia perparva Plummer, 1930). Free, elongate test, composed of a globular proloculus followed by an undivided straight tubular chamber. The wall is calcareous microgranular. Late Silurian to Early Triassic (Plate 2.1, figs 1- 5; Plate 2.3, Fig. 1A; Plate 3.2, fig.3). • Gigasbia Strank, 1983 (Type species: Gigasbia gigas Strank, 1983). The test is free and elongate, consisting of a globular proloculus and an undivided tubular chamber. Early Carboniferous (Tournaisian). Family Endotebidae Vachard, Martini, Rettori and Zaninetti, 1994 The test is free, planispiral, in early stages, but later uniserial to biserial. The walls are calcareous, grey, thick, calcareous agglutinated, and the apertures are simple. Late Permian to Triassic. • Endoteba Vachard and Razgallah, 1988 emend. Vachard et  al. 1994 (Type spe- cies:  Endoteba controversa Vachard et  al., 1994). Axial view compressed. Late Permian to Late Triassic (Late Kungurian to Rhaetian) (Fig. 3.5). The Palaeozoic Larger Benthic Foraminifera 55 Family Pseudoammodiscidae Conil and Lys, 1970 The Pseudoammodiscidae are small, with a simple aperture and globular prolocu- lus followed by a planispiral, trochospiral or streptospirally coiled, undivided second chamber. They occur from the Devonian to the Permian. • Brunsia Mikhaylov, 1935 (Type species: Spirillina irregularis von Möller, 1879). The proloculus is followed by an early streptospiral coil, and later a planispirally enrolled tubular chamber. Devonian to Carboniferous (Tournaisian) (Plate 2.3, Fig. 9). • Brunsiella Reytlinger, 1950 (Type species:  Glomospira ammodiscoidea Rauzer- Chernousova, 1938). The globular proloculus is followed by an undivided sec- ond chamber. Early whorls are streptospiral, later ones planispiral and evolute. Carboniferous (Visean) to Early Permian (Plate 2.2, Fig. 15B; Plate 2.3, figs 9,11,12; Plate 2.11, figs 7- 9). Family Pseudolituotubidae Conil and Longerstaey, 1980 The Pseudolituotubidae have attached tests, consisting of a single enrolled chamber with a simple terminal aperture and a compound calcareous microgranular wall. They range from the Early Carboniferous (Visean) to the Late Carboniferous (Moscovian). • Pseudolituotuba Vdovenko, 1971 (Type species: Lituotuba? gravata Conil and Lys, 1965). The proloculus is followed by an undivided streptospirally coiled tube. Carboniferous (Visean to Moscovian). Superfamily CALIGELLOIDEA Reytlinger 1959 This superfamily (Fig. 2.2) had simple free or attached tests, consisting of a proloculus and a rectilinear second chamber that was subseptate (i.e. with partial septa). They had simple walls and evolved from the Parathuramminoidea in the Late Silurian, but died out in the Early Carboniferous (Tournaisian). Family Caligellidae Reytlinger, 1959 In this Family, the test is attached and subseptate. They ranged from the Late Silurian to the Early Carboniferous (Tournaisian). • Paracaligella Lipina, 1955 (Type species:  Paracaligella antropovi Lipina, 1955). The test is composed of a sub-s pherical proloculus followed by an irregular tubu- lar chamber partially divided by incipient septa. Devonian to Carboniferous (Late Tournaisian) (see Fig. 2.2). Family Paratikhinellidae Loeblich and Tappan, 1984 This family was free with a elongate, sub- septate test and with a simple microgranular wall. They ranged from Middle Devonian to Carboniferous (Tournaisian). • Paratikhinella Reytlinger, 1954 (Type species: Tikhinella cannula Bykova, 1952). The test is elongate and composed of a subglobular proloculus followed by a cylindrical tubular chamber with incipient septation. Late Devonian to Early Carboniferous (Plate 2.4, Fig. 5). 56 Evolution and Geological Significance of Larger Benthic Foraminifera • Saccaminopsis Sollas, 1921 (Type species:  Saccammina carteri Brady, 1871). The test is free, uniserial with ovate chambers, a thin wall and a terminal aperture. Late Devonian to Carboniferous (Tournaisian). (Plate 2.1, Fig. 11B; Plate 2.2, figs 12-1 5). Superfamily PTYCHOCLADIOIDEA Elias, 1950 These were attached forms with no distinct apertures. Their walls were microgranular, banded with transverse tubuli. They range from Late Devonian to Late Carboniferous. Family Ptychocladiidae Elias, 1950 Here the test is uniserial and attached. They ranged from the Late Devonian to Late Carboniferous (Pennsylvanian). • Ptychocladia Ulrich and Bassler, 1904 (Type species: Ptychocladia agellus Ulrich and Bassler, 1904). The wall is calcareous microgranular, with two layers with no distinct aperture. Carboniferous (Late Pennsylvanian) (see Fig. 2.2). ORDER PALAEOTEXTULARIIDA HOHENEGGER AND PILLER, 1975 Members of this order have biserial or uniserial tests, with a microgranular calcareous wall, commonly with an inner radial fibrous layer and a finely granular outer layer. The aperture is generally single but may be multiple in later stages. Their range is from Late Devonian to Permian. Superfamily PALAEOTEXTULARIOIDEA Galloway, 1933 This superfamily has a test that is biserial or uniserial, with a single or two- layered microgranular calcareous wall. The aperture is generally single but may be multiple in later stages. Their range is from Late Devonian to Permian. Family Semitextulariidae Pokorny, 1956 Here the test is biserial and flattened, but often becoming monoserial with broad chambers, and is fully septate (see Fig.  2.2). They range from the Devonian to the Carboniferous (Pennsylvanian). • Koskinobigenerina Eickhoff, 1968 (Type species:  Koskinobigenerina breviseptata Eickhoff, 1968). The test is elongate, biserial in the early stage, later becoming unise- rial. Early Carboniferous to Late Carboniferous (Visean to Gzhelian). • Koskinotextularia Eickhoff, 1968 (Type species:  Koskinotextularia cribriformis Eickhoff, 1968). The test is elongate, biserial throughout with a single layered wall. The aperture is single in the early stage but later becoming cribrate. Early Carboniferous to Late Carboniferous (Visean to Gzhelian) (Plate 2.4, Fig. 6). Family Palaeotextulariidae Galloway, 1933 This family (see Figs 2.2 and 2.3) has biserial to uniserial genera that closely resembles the Textulariidae, but they have a dark granular calcareous outer wall and an inner clear to yellowish “fibrous” layer with stacks of granules perpendicular to the surface. Cummings (1956) demonstrated that these forms evolved from simple agglutinated forms in the Devonian (see Fig. 2.4). Early Carboniferous (Tournaisian) to Permian (Lopingian). The Palaeozoic Larger Benthic Foraminifera 57 • Climacammina Brady, 1873 (Type species: Textularia antiqua Brady, in Young and Armstrong, 1871). The test has a biserial early stage followed by a uniserial stage. The aperture in the early stage is at the base of the last chamber, later becoming areal, multiple and cribrate. Carboniferous to Permian (Visean to Early Lopingian) (Plate 2.4, Fig. 7; Plate 2.5, figs 7, 9- 10). 251.9 298.9 323.2 ? 358.9 Semitextulariidae Palaeotextulariidae Nondosinellidae PALAEOTEXTULARIOIDEA NODOSINELLOIDEA Fig. 2.4. Evolution of the Palaeotextulariida (modified from Cummings (1956). Devonian Carboniferous Permian Tournaisian Visean Serpukhovian Bashkirian Moscovian Kasimovian Gzhelian Cisuralian Guadalupian Lopingian Monogenerina Palaeobigenerina Deckerella Deckerellina Palaeotextularia Cribrostomum Climacammina Cribrogenerina Earlandinitidae Nodosinella 58 Evolution and Geological Significance of Larger Benthic Foraminifera • Cribrogenerina Schubert 1908 (Type species: Bigenerina sumatrana Volz, 1904). The test has a short early biserial stage followed by a uniserial stage. The aperture is a slit in the early stage, later becoming areal, multiple and cribrate in the uniserial stage. Carboniferous to Permian (latest Gzhelian to Lopingian) (see Fig. 2.4). • Cribrostomum von Möller, 1879 (Type species: Cribrostomum textulariforme von MöIler, 1879). The test is biserial throughout. The aperture in the early part is basal, but in the last two whorls becomes areal, multiple and cribrate. Carboniferous (late Tournaisian to Gzhelian) (Plate 2.4, figs 1, 5, 13, 15; see Figs 2.2; 2.4). • Deckerella Cushman and Waters, 1928 (Type species: Deckerella clavata Cushman and Waters, 1928). The test is biserial in the early stage, later becoming uniserial and rectilinear. The aperture is a low slit in the early stage, later becoming terminal with two parallel slits separated by a partition. Carboniferous to Permian (late Visean to early Lopingian) (Plate 2.4, Fig. 8; see Fig. 2.4). • Deckerellina Reytlinger, 1950 (Type species: Deckerellina istiensis Reytlinger, 1950). The test is biserial throughout. The aperture, in the early stage, is a low slit, but the adult aperture is two parallel slits. Carboniferous (Visean to Moscovian) (Plate 2.4, Fig. 11, 16; see Fig. 2.4). • Monogenerina Spandel, 1901 (Type species: Type species: Monogenerina atava Spandel, 1901). The test is biserial in the early stage, later becoming uniserial. The aperture extends throughout the test. Permian (Cisuralian to Lopingian) (see Fig. 2.4). • Palaeobigenerina Galloway, 1933 (Type species: Bigenerina geyeri Schellwien, 1898). The test is biserial in the early stage, but later uniserial, with a slit opening in the biserial stage, followed by a single rounded aperture. Carboniferous to Permian (late Serpukhovian to early Lopingian) (see Fig. 2.4). • Palaeotextularia Schubert, 1921 (Type species: Palaeotextularia schellwieni Galloway and Ryniker, 1930). The test is biserial throughout. The wall is thick, and the a perture is a low slit opening at the base of the final chamber. Carboniferous to Permian (Tournaisian to Cisuralian) (Fig. 2.4; Plate 2.4, figs 2-4 , 9-1 0, 12, 14, 17-1 8; Plate 2.5, Fig. 6). Family Biseriamminidae Chernysheva, 1941 In this family (Fig.  2.2), the test is biserial becoming planar in later stages, with a microgranular wall with one or more layers. They range from the Carboniferous (Tournaisian) to Permian (Lopingian). • Biseriammina Cherhysheva, 1941 (Type species: Biseriammina uralica Cherhysheva, 1941). The test is planispirally enrolled. Early Carboniferous (Tournaisian). • Biseriella Mamet, 1974 (Type species:  Globivalvulina parva Chernysheva, 1948). The early stage is tightly coiled, while later it has an open helicoid spire. Late Carboniferous (Bashkirian to Moscovian) (Plate 2.3, Fig. 10). • Dagmarita Reytlinger, 1965 (Type species:  Dagmarita chanakchiensis Reyltinger, 1965). The test is flattened with spine- like projections at the outer corners of the angular chambers. Late Permian (Lopingian). • Globispiroplectammina Vachard, 1977 (Type species:  Globispiroplectammina mam- eti Vachard, 1977). The test is biserial with the early stage enrolled, later becoming uncoiled. Carboniferous (Visean). The Palaeozoic Larger Benthic Foraminifera 59 • Globivalvulina Schubert, 1921 (Type species: Valvulina bulloides Brady, 1876). The test is planispirally coiled. Middle Carboniferous (Bashkirian) to Late Permian (Wuchiapingian). • Lipinella Malakhova, 1975 (Type species: Lipinella notata Malakhova, 1975). The test is planispirally enrolled following a globular proloculus. Middle Carboniferous to Late Permian (Bashkirian to Lopingian). • Louisettita Altiner and Brönnimann, 1980 (Type species: Louisettita elegantissima Altiner and Brönnimann, 1980). The test is trochospirally enrolled in the early stage. In the biserial stage, chambers are subdivided by vertical partitions perpendicular to the septa and have spine-l ike projections. Late Permian (Wuchiapingian). • Paraglobivalvulina Reytlinger, 1965 (Type species: Paraglobivalvulina mira Reytlinger, 1965). The test is spherical with the biserial chambers enrolled trochospirally. Late Permian (Wuchiapingian). • Paradagmarita Lys, 1978 (Type species: Paradagmarita monodi Lys, 1978). The test is small with the early stage completely enrolled, but the later stage is uncoiled. Late Permian (Wuchiapingian). • Paraglobivalvulinoides Zaninetti and Jenny- Deshusses, 1985 (Type spe- cies: Paraglobivalvulina? septulifera Zaninetti and Jenny-D eshusses, 1981). The test is globular with biserially, enrolled, strongly enveloping chambers. The wall is cal- careous, microgranular, a single layer. The aperture is rimmed with a well- developed tongue that bends inward. Late Permian (Wuchiapingian). ORDER TETRATAXIDA MIKHALEVICH 1981 Members of this order have a conical, trochospiral test with an evolute spiral side and involute umbilical side. The wall is microgranular calcareous with one or two dis- tinct layers. Early Carboniferous to Late Permian (Tournaisian to Wuchiapingian). Superfamily TETRATAXOIDEA Galloway, 1933 The Tetrataxoidea have a conical to spreading test with secondary internal parti- tions. The walls are microgranular with one or two layers. They range from the Early Carboniferous to Late Permian (Tournaisian to Wuchiapingian). Family Tetrataxidae Galloway, 1933 The test of the Tetrataxidae is conical leaving an open central umbilicus where the chambers are partially overlapping (Fig. 2.2). The wall is calcareous, microgranular, and in two layers. They occur from the Early Carboniferous to the Late Carboniferous (Tournaisian to Moscovian). • Globotetrataxis Brazhnikova, 1983 (Type species:  Tetrataxis (Globotetrataxis) ele- gantula Brazhnikova, in Brazhnikova and Vdovenko, 1983). The test enlarges grad- ually with a last large hemispherical umbilical chamber that forms a convex base. Early Carboniferous (late Visean). • Polytaxis Cushman and Waters, 1928 (Type species: Polytaxis laheei Cushman and Waters, 1928). The test is low conical. Later chambers form many whorls and result in a spreading test. Late Carboniferous (Moscovian). 60 Evolution and Geological Significance of Larger Benthic Foraminifera Early spiral chambers A strongly overlapping chambers B Fig. 2.5. Main morphological features of Tetrataxis. A) The early spires of a solid specimen, B) An oblique thin section through the axis of the test. • Tetrataxis Ehrenberg, 1854 (Type species: Tetrataxis conica Ehrenberg, 1854). The test is circular in plan with chambers strongly overlapping on the umbilical side. The outer layer is dark and microgranular and the inner layer is light and fibrous. Carboniferous (late Tournaisian to Visean) (Fig. 2.5; Plate 2.2, figs 1, 2, 5; Plate 2.4, Fig. 17B; Plate 2.7, figs 8- 18). Family Pseudotaxidae Mamet, 1974 In this family, the wall is calcareous, microgranular and single layered. They range from Early Carboniferous to Late Carboniferous (Late Tournaisian to Early Bashkirian). • Pseudotaxis Mamet, 1974 (Type species:  Tetrataxis eominima Rauzer- Chernousova, 1948). The test is conical with the proloculus followed by irregular trochospiral coil. Early Carboniferous to Late Carboniferous (late Tournaisian to Early Bashkirian). • Vissariotaxis Mamet, 1970 (Type species: Monotaxis exilis Vissarionova, 1948). The test is conical with trochospirally coiled chambers and a wide aperture. Carboniferous (early Bashkirian) (Plate 2.10, figs 8- 12). Family Valvulinellidae Loeblich and Tappan, 1984 Here the test is conical with subdivided chambers and a single layer microgran- ular wall. The range is from the Early to Late Carboniferous (Visean to early Bashkirian). The Palaeozoic Larger Benthic Foraminifera 61 • Valvulinella Schubert, 1908 (Type species: Valvulina youngi Brady, 1876). Chambers are subdivided by numerous vertical pillars and one or two horizontal ones. Carboniferous (Visean to Early Bashkirian). (Plate 2.2, figs 3-4 ; Plate 2.9, fig. 4; Fig. 2.22C). Family Abadehellidae Loeblich and Tappan, 1984 Representatives of this family differ from the Valvulinellidae in having two double- layered walls instead of a single layered wall. They occur in the Late Permian (Wuchiapingian). • Abadehella Okimura and Ishii, 1975 (Type species: Abadehella tarazi Okimura and Ishi, in Okimura et al., 1975). The test is conical, with numerous whorls. Chambers are subdivided by regularly spaced pillars. Late Permian (Wuchiapingian) (Plate 2.3, Fig. 12). ORDER TOURNAYELLIDA HOHENEGGER AND PILLER 1973 The representatives of this order have walls that range from homogeneously calcare- ous microgranular to those that are differentiated into two or more layers in the most advanced forms. Late Devonian to Carboniferous. Superfamily TOURNAYELLOIDEA Dain, 1953 In this superfamily, the proloculus is followed by a planispiral or streptospiral non- septate to fully septate test. The wall is made of microgranular calcite evolving into forms with a thick outer layer, called a tectum, and an inner granulo- fibrous layer. Late Devonian to Carboniferous (Frasnian to Early Bashkirian). Family Tournayellidae Glaessner, 1945 The Tournayellidae are small, with early streptospiral or trochospiral coiling, tending to become planispiral in the adult. The aperture is simple and open at the end of a tubular chamber. The Tournayellidae evolved in the Devonian from simple planispiral, non- septate forms (e.g. Eotournayella, see Fig. 2.6) with an initial proloculus and an aperture at the end of an uncoiled tube, to forms with a compound wall and rudimen- tary septa in the Late Tournaisian (e.g. Tournayella Dain, 1953, see Fig. 2.6; Carbonella Dain, 1953, see Fig. 2.6; Plate 2.9, Fig. 14) to almost complete septa in the late Visean (Mstinia Dain, Plate 2.9, figs 8, 10). The streptospiral tubular forms (Glomospiranella Lipina, 1951, see Fig.  2.6) gradually also developed rudimentary septa in the Late Tournaisian (e.g. Brunsiina Lipina, 1953, see Fig. 2.6). The Tournayellidae range from the Devonian to the Carboniferous. • Bogushella Conil and Lys, 1977 (Type species: Mistinia ziganensis Grozdilova and Lebedeva, 1960). The test is streptospirally enrolled in the early stage, later becom- ing rectilinear with no true septa. The aperture is cribrate. Early Carboniferous (Visean). • Brunsiina Lipina, 1953 (Type species: Brunsiina uralica Lipina, 1953). The test is streptospirally enrolled in early stage, later becoming planispiral, without rudi- mentary constrictions. Late Devonian to Early Carboniferous (Fammenian to Visean) (Fig. 2.6). 62 Evolution and Geological Significance of Larger Benthic Foraminifera L. Devonian L. Devonian - Tournaisian L. Tounaisian Eotournayella Tournayella Carbonella Tectum Inner granulo-fibrous layer Glomospiranella Septaglomospiranella Tournayellidae Brunsiina Septabrunsiina Fig. 2.6. Schematic evolution of the Tournayellidae. • Carbonella Dain, 1953 (Type species Carbonella spectabilis Dain, 1953). The test is planispiral with septa that are well developed only in the final whorl. Early Carboniferous (Late Tournaisian to Early Visean) (Fig. 2.6; Plate 2.9, Fig. 14). • Conilites Vdovenko, 1970 (Type species:  Ammobaculites? Dinantii Conil and Lys, 1964). The test is planispirally enrolled and undivided in the early undivided part, but the later part is uncoiled rectilinear with distinct septa. The aperture is multiple and cribrate. Late Carboniferous (Late Tournaisian to Early Visean). • Costayella Conil and Lys, 1977 (Type species:  Tournayella costata Lipina, 1955). The test has an early part that is streptospiral, but is later planispiral, with later whorls having slight constrictions opposite basal supplementary deposits. Early Carboniferous (Tournaisian to Visean). • Chernobaculites Conil and Lys, 1977 (Type species:  Ammobaculites sarbaicus Malakhova subsp. beschevensis Brazhnikova, in Brazhnikova et al., 1967). The test is streptospirally enrolled in the early part, later becoming rectilinear with straight and horizontal septa. Late Carboniferous (Bashkirian). • Chernyshinella Lipina, 1955 (Type species: Endothyra glomiformis Lipina, 1948). The test is streptospirally enrolled throughout with a tubular undivided early portion, and a final whorl with strongly asymmetrical and teardrop-s haped chambers. Late Devonian to Early Carboniferous (Late Famennian to Visean) • Chernyshinellina Reytlinger, 1959 (Type species:  Ammobaculites? pygmaeus Malakhova, 1954). The test is streptospirally enrolled in the early stage and has teardrop- like chambers in the later whorl, which are separated by distinct horizontal septa. Carboniferous (Tournaisian) (Plate 2.3, Fig. 16). • Condrustella Conil and Longerstaey, 1977 (Type species: “Mstinia” modavensis Conil and Lys, 1967). The test is streptospirally enrolled in the early part, later becoming planispiral with teardrop-s haped chambers. Early Carboniferous (Late Tournaisian to Visean). The Palaeozoic Larger Benthic Foraminifera 63 • Elbanaia Conil and Marchant, 1977 (Type species: Plectogyra michoti Conil and Lys, 1964). The test is streptospirally enrolled in the early part, but later evolute and sep- tate. Early Carboniferous (Late Tournaisian to Early Visean). • Endochernella Conil and Lys, 1980 (Type species: Plectogyra (Latiendothyra) quae- sita Ganelina, 1966). The test is streptospirally enrolled in the early part, but later with radial septa and numerous chambers in the whorl. Early Carboniferous (Early Tournaisian to Visean). • Eoforschia Mamet, 1970 (Type species:  Tournayella moelleri Malakhova, in Dain and Grozdilova, 1953). The test is planispirally enrolled lacking definite septa. Early Carboniferous (Tournaisian to Visean). • Eotournayella Lipina and Pronina, 1964 (Type species Tournayella (Eotournayella) jubra Lipina and Pronina, 1964). The test is streptospiral to planispiral in later stages, with slight growth constrictions. Late Devonian (Fig. 2.8). • Forschia Mikhaylov, 1935 (Type species: Spirillina subangulata von Möller, 1879). The test is evolutely coiled without true septa. The aperture is cribrate. Early Carboniferous (Visean). • Forschiella Mikhaylov, 1935 (Type species: Forschiella prisca Mikhaylov, 1935). The test is evolutely coiled without true septa in the early stage, later it is uncoiled. The aperture is cribrate. Early Carboniferous (Visean). • Glomospiranella Lipina, 1951 (Type species: Glomospiranella Asiatica Lipina, 1951). The test is streptospiral in the early stage, later becoming streptospiral, without dis- tinct septation, but with slight constrictions. Late Devonian to Early Carboniferous (Visean) (Fig. 2.6). • Glomospiroides Reytlinger, 1950 (Type species: Glomospiroides fursenkoi Reytlinger, 1950). The test is streptospirally enrolled, being undivided in the early stage, but becoming rectilinear and irregularly divided with thin pseudosepta. Middle to Late Carboniferous. • Laxoseptabrunsiina Vachard, 1977 (Type species:  Laxoseptabrunsiina valuzierensis Vachard,1977). The test is streptospirally enrolled and undivided, later becoming planispiral with true septa. Early Carboniferous (Visean). • Lituotubella Rauzer- Chernousova, 1948 (Type species: Lituotubella glomospiroides Rauzer- Chernousova, 1948). The test is streptospirally coiled in the early stage, later becoming rectilinear with no true septa. Early Carboniferous (Visean). • Mstinia Dain, 1953 (Type species: Mstinia bulloides Dain, in Dain and Grozdilova, 1953). The test is involute, with teardrop-l ike chambers. The aperture is cribrate. Early Carboniferous (Late Visean). • Mstiniella Conil and Lys, 1977 (Type species: Mstinia fursenkoi Dain, in Dain and Grozdilova, 1953). The test is streptospirally, undivided in the early part, but later planispiral with septa. The aperture is cribrate. Early Carboniferous (Visean) (Plate 2.9, figs 8, 10). • Neobrunsiina Lipina, 1965 (Type species: Glomospiranella finitima Grozdilova and Lebedeva, 1954). The test is streptospirally enrolled in the early stage, but later pla- nispiral with constrictions in the wall. Early Carboniferous (Tournaisian to Visean). • Nevillea Conil and Lys, 1980 (Type species: Georgella dytica Conil and Lys, 1977). The test is streptospirally enrolled in the early stage, but later rectilinear, divided by arched septa. The aperture is multiple with a cribrate aperture occupying the entire terminal chamber surface. Early Carboniferous (Visean). 64 Evolution and Geological Significance of Larger Benthic Foraminifera • Nodochernyshinella Conil and Lys, 1977 (Type species:  Chernyshinella tumulosa Lipina, 1955). The test is streptospirally enrolled in the early part, but later planispi- ral with teardrop- like chambers. Early Carboniferous (Tournaisian). • Pohlia Conil and Lys, 1977 (type species: Septatournayella henbesti Skipp, in Skipp et al., 1966). The test is planispirally enrolled with distinct septa in the final whorl. Early Carboniferous (Visean). • Pseudolituotubella Vdovenko, 1967 (Type species: Pseudolituotubella multicamerata Vdovenko, 1967). The test is streptospirally enrolled in the early stage, but later is planispiral becoming rectilinear with true septa. The aperture is cribrate. Early Carboniferous (Late Tournaisian to Early Visean). • Rectoseptatournayella Brazhnikova and Rostovceva, 1963 (Type spe- cies: Rectoseptatournayella stylaensis Brazhnikova and Rostovceva, 1963). The test is planispirally coiled in the early stage, but uncoiled in the final stage with septa and distinct chambers. Early Carboniferous (Tournaisian). • Rectotournayellina Lipina, 1965 (Type species:  Tournayellina (Rectotournayellina) Lipina, 1965). The test is streptospirally enrolled in the early part, but later rectilin- ear. Late Devonian to Early Carboniferous (Famennian to Tournaisian). • Septabrunsiina Lipina, 1955 (Type species: Endothyra? Krainica Lipina, 1948). The test is streptospirally enrolled in the early nonseptate stage, but later it is planispiral with distinct septa. Early Carboniferous (Tournaisian) (Fig. 2.6). • Septaforschia Conil and Lys, 1977 (Type species: Tournayella questita Malakhova, in Dain and Grozdilova, 1953). The test is enrolled, with later stages having true septa. Early Carboniferous (Tournaisian). • Septaglomospiranella Lipina, 1955 (Type species:  Endothyra primaeva Rauzer- Chernousova, 1948) Early Carboniferous (Tournaisian) (Fig. 2.6). • Septatournayella Lipina, 1955 (Type species: Tournayella segmentata Dain, in Dain and Grozdilova, 1953). The test is a planispirally enrolled tube, with slight constric- tions in the early stage, but later with complete septa forming distinct chambers. Late Devonian (Famennian) to Early Carboniferous (Visean). • Spinobrunsiina Conil and Longerstaey, 1980 (Type species:  Septabrunsiina (Spinobrunsiina) ramsbottomi Conil and Longerstaey, 1980). The test is strepto- spirally enrolled in the early stage, but later becoming planispiral and septate. Early Carboniferous (Tournaisian to Visean). • Spinolaxina Conil and Naum, 1977 (Type species: Plectogyra pauli Conil and Lys, 1964). The test is streptospirally coiled in the early stage, later becoming nearly pla- nispiral. Early Carboniferous (Visean). • Spinotournayella Mamet, 1970 (Type species: Plectogyra tumula Zeller, 1957). The test is involute with early parts being streptospiral, but later becoming planispiral with endothyroid chambers. Early Carboniferous (Late Tournaisian). • Tournayella Dain, 1953 (Type species: Tournayella discoidea Dain, 1953). The test is enrolled without definite septa and supplementary deposits. Early Carboniferous (Tournaisian to Visean) (Fig. 2.5). • Tournayellina Lipina, 1955 (Type species:  Tournayellina vulgaris Lipina, 1955). The test is enrolled with few chambers enlarging rapidly. Late Devonian to Early Carboniferous (Famennian to Tournaisian). The Palaeozoic Larger Benthic Foraminifera 65 • Uviella Ganelina, 1966 (Type species:  Uviella aborigena Ganelina, 1966). The test is streptospirally enrolled in the early undivided stage, but later is planispiral with constrictions in the early part followed in the final whorl by true septa. Early Carboniferous (Tournaisian to Visean). • Viseina Conil and Lys, 1977 (Type species: Septatournayella? Conspecta Conil and Lys, 1977). The test is enrolled with later chambers divided by true septa. The aper- ture is cribrate in the final stage. Early Carboniferous (Visean). Family Palaeospiroplectamminidae Loeblich and Tappan, 1984 The members of this family represent a biserial development from the Tournayellidae, and may be ancestral to the Palaeotextularioidea. They are initially streptospiral with teardrop- like or endothyroid-l ike chambers, then become planispiral and in later stages the test become biserial. The range is Late Devonian to Early Carboniferous. • Endospiroplectammina Lippina, 1970 (Type species:  Spiroplectammina venusta Vdovenko, 1954). The test is streptospirally enrolled, with distinctly endothyroid chambers, later becoming planispiral, uncoiled with a later biserial stage with a sin- gle aperture. Early Carboniferous (Late Tournaisian to Middle Visean). • Eotextularia Mamet, 1970 (Type species: Palaeotextularia diversa Chernysheva, 1948). The test is irregularly coiled in the early part, later becoming uncoiled with a few pairs of biserial chambers. Early Carboniferous (Late Tournaisian to Middle Visean). • Halenia Conil, 1980 (Type species: Halenia legrandi Conil, 1980). The test has a bise- rial elongate stage, followed by a broad uniserial stage with horizontal septa. The aperture is cribrate in the adult. Early Carboniferous (Visean). • Palaeospiroplectammina Lipina, 1965 (Type species:  Spiroplectammina tchernyshi- nensis Lipina, 1948). The test has a biserial uncoiled stage with curved septa, but may become uniserial with last chambers. Late Devonian to Carboniferous (Early Visean) (Plate 2.3, Fig. 11). • Rectochernyshinella Lipina, 1960 (Type species: Spiroplectammina mirabilis Lipina, 1948). The test is streptospirally enrolled in a large early stage, the later stage uncoiled with few biserially arranged chambers. Late Devonian to Early Carboniferous (Late Famennian to Early Tournaisian). ORDER ENDOTHYRIDA FURSENKO 1958 Members of this order have a lenticular test, planispirally coiled. The wall is dark microgranular, but sometimes can be bilayered or multilayered. Aperture simple, basal or cribrate. Late Devonian to Triassic. Superfamily ENDOTHYROIDEA Brady, 1884 nom. translat. Fursenko, 1958 Members of this superfamily have typical endothyroid coiling with streptospiral to planispiral tests and constant deviation of the axis of coiling, with many chambers fol- lowed by a rectilinear stage in some forms (see Fig. 2.9). The walls are microgranular and calcareous, but some forms evolved two to three distinct layers, others may develop an inner perforate or keriothecal layer. Notably, they ranged from the Late Devonian to the Triassic (Famennian to Rhaetian). 66 Evolution and Geological Significance of Larger Benthic Foraminifera Family Endotebidae Vachard, Martini, Rettori and Zaninetti, 1994 Test: free planispiral, in early stages, later uniserial to biserial. Wall: calcareous, grey, thick, calcareous agglutinated. Aperture: simple. Late Permian to Triassic. • Endoteba Vachard and Razgallah, 1988 emend. Vachard, Martini, Rettori and Zaninetti, 1994 (Type species: Endoteba controversa Vachard and Razgallah, 1988 emend. Vachard, Martini, Rettori and Zaninetti, 1994). The axial view is com- pressed. Late Permian to Late Triassic (Late Kungurian to Rhaetian) (Fig. 3.4). Family Endothyridae Brady, 1884 Here the test is small, multilocular, enrolled, with evolute planispiral coiling and well developed septa in early whorls (e.g. Loeblichia Cummings in the Visean), or initially streptospiral to planispiral characterised by the development of secondary depos- its of calcite (chomata) on the chamber floor (e.g. Endothyra Phillips, 1846, Early Carboniferous to Permian). They exhibit a low equatorial aperture. Advanced forms may become uniserial (e.g. Haplophragmella Rauzer- Chernousova, 1936, Visean, see Fig. 2.7) with a cribrate terminal aperture. Their range is Late Devonian to Permian. • Andrejella Malakhova, 1975 (Type species: Andrejella laxiformis Malakhova, 1975). The test is streptospirally coiled in the early part, with later chambers rapidly increas- ing in height. Early Carboniferous (Visean). • Banffella Mamet, 1970 (Type species:  Endothyra? banffensis McKay and Green, 1963). The test is streptospirally coiled in early whorl, later it is nearly planispiral and evolute, with straight and slightly oblique septa well developed only in early whorls. Early Carboniferous (Visean). • Corrigotubella Ganelina, 1966 (Type species: Corrigotubella posneri Ganelina, 1966). The test is streptospirally enrolled in the early stage, later it is planispiral and recti- linear, with a few short and broad chambers. Septa in the enrolled stage are short, while those of the rectilinear portion are horizontal. Early Carboniferous (Late Tournaisian to Early Visean) • Cribranopsis Conil and Longerstaey, 1980 (Type species: Cribranopsis fossa Conil and Longerstaey, in Conil et al., 1980). The test is enrolled, with a streptospiral early stage and a planispiral late coil, expanding rapidly. The septa are short and thick. Early Carboniferous (Visean). • Cribrospira von Möller, 1878 (Type species: Cribrospira panderi von Möller, 1878). The test is streptospirally enrolled in the early stage, but later coiling is planispiral with numerous chambers per whorl. Septa are nearly radial, short and thick. Early Carboniferous (Visean). • Elergella Conil, 1984 (Type species: Elergella simakoyi Conil, 1984). The test is strep- tospirally enrolled in the early stage, later becoming planispiral and evolute with short, slightly oblique septa. Early Carboniferous (Tournaisian) • Endostaffella Rozovskaya, 1961 (Type species: Endothyra parya von Möller, 1879). The test is streptospirally enrolled in the early stage, later becoming planispiral and evolute. Early Carboniferous (Late Tournaisian to Visean). • Endothyra Phillips, 1846 (Type species: Endothyra bowmani Phillips, 1846). The text is partially involute, planispiral with an early part streptospirally enrolled. It is formed The Palaeozoic Larger Benthic Foraminifera 67 by few whorls, with strong development of secondary deposits, such as nodes, ridges, or hooks on the chamber floor. The wall is calcareous, microgranular, with two or three layers, a thin dark outer layer or tectum and a thicker, fibrous to alveolar inner layer or diaphanotheca. The aperture is simple, basal. Early Carboniferous to Permian (Plate 2.5, Fig. 20; Plate 2.12, Fig. 8; Plate 2.6, figs 1- 2; Plate 2.13, figs 1- 3; see Figs 2.7, 2.8). • Endothyranella Galloway and Harlton, 1930 (Type species:  Ammobaculites pow- ersi Harlton, 1927). Streptospiral test becoming planispiral and then rectilinear. The aperture is simple and rounded. Late Carboniferous (Moscovian to Ladinian) (Plate 2.9, Fig. 1; Plate 3.5, Fig. 10). • Endothyranopsis Cummings, 1955 (Type species: Involutina crassa Brady, in Moore, 1870). The test is subglobular involute, being almost planispiral with not more than 3 and a half whorls. They have a thick wall with no secondary deposits. Early coiling Loeblichia Visean Tectum Endothyra Inner “ganulo-fibrous” L. Devonian ? - Lower Carboniferous-Permian H aplophragmella Bradyina E. - L. Carboniferous E. - L. Carboniferoust Fig. 2.7 Evolution of the Endothyroidea. 68 Evolution and Geological Significance of Larger Benthic Foraminifera plectogyroid with a large proloculus. Sutures are perpendicular to the spirotheca. The aperture is a low basal equatorial arch. Carboniferous (Visean) (Plate 2.9, figs 3, 6, 7; Plate 2.12, figs 7, 10, 11; Plate 2.14, figs 1, 4, 5). • Eoendothyra Miklukho- Maklay, 1960 (Type species: Endothyra communis Rauzer- Chernousova, 1948). The test is involute and streptospirally enrolled in the early stage, later becoming nearly planispiral, with numerous chambers. Late Devonian to Early Carboniferous (Famennian to Visean). • Eoendothyranopsis Reytlinger and Rostovzeva, 1966 (Type species:  Parastaffella pressa Grozdilova, in Lebedeva, 1954). The test is planispirally enrolled, involute, with oblique septa and secondary deposits at the base of the chambers that appear as a forward projecting hook or spine in the final chamber. Early Carboniferous (Early Visean). • Eoquasiendothyra Durkina, 1963 (Type species: Endothyra bella Chernysheva, 1952). The test is streptospirally enrolled in the early stage, but later becoming planispiral, with inflated chambers and septa slightly oblique to the outer wall. Late Devonian to Early Carboniferous (Tournaisian). • Euxinita Conil and Dil, 1980 (Type species:  Dainella? efremoyi Vdovenko and Rostovtseva, in Brazhnikova et al., 1967). The test is streptospirally enrolled in the Sub-rectangular chamberlets A B Early streptospiral whorls Tectum plus C Keriothecal structure Fig. 2.8 Main morphological features of A) Loeblichia, B) Endothyra and C) Bradyina. The Palaeozoic Larger Benthic Foraminifera 69 early stage with numerous chambers per whorl, later it becomes nearly planispiral. Early Carboniferous (Visean to Bashkirian). • Globochernella Hance, 1983 (Type species: Globochernella braibanti Hance, 1983). The test is enrolled, with chambers enlarging rapidly as added. The septa are thick at the base and tapering rapidly to a very thin at the inner edge. Early Carboniferous (Visean). • Globoendothyra Bogush and Yuferev, 1962 (Type species:  Globoendothyra pseudo- globulus Bogush and Yuferev, 1962). The test is streptospirally enrolled in the early stage, but later is nearly planispiral with oblique septa. Secondary deposits on the chamber floors. Early Carboniferous (Visean to Moscovian). • Granuliferella Zeller, 1957 (Type species: Granuliferella granulosa Zeller, 1957). The test is involute, streptospirally enrolled in the early stage, but later nearly planispiral, with relatively few chambers per whorl. The septa are short and slightly oblique. Late Devonian (Famennian) to Early Carboniferous (Visean). • Granuliferelloides McKay and Green, 1963 (Type species: Granuliferelloides jasperen- sis McKay and Green, 1963). The test is streptospirally enrolled in the early stage, but later are nearly planispiral, with a few slightly inflated chambers and oblique septa, becoming uncoiled, with short cylindrical chambers and nearly horizontal septa. Early Carboniferous (Late Tournaisian). • Haplophragmella Rauzer- Chernousova and Reytlinger, 1936 (Type species: Endothyra panderi von Möller,1879). The test is streptospirally enrolled in the early stage, with few chambers per whorl, later chambers become uncoiled and rectilinear. The aper- ture is simple in the early stage, becoming multiple and cribrate in the later stages. Early Carboniferous (Visean) (Fig. 2.7). • Haplophragmina Reytlinger, 1950 (Type species:  Haplophragmina kashkirica Reytlinger, 1950). The test is involute, planispirally enrolled in the early stage, becom- ing uncoiled in the late stage. Late Carboniferous (Moscovian). • Holkeria Strank, 1982 (Type species: Rhodesina avonensis Conil and Longerstaey, in Conil et al., 1980). The test is streptospirally enrolled with the plane of coiling oscil- lating in the early stage, but later it is nearly planispiral and evolute with short septa short, which follow the curvature of the chambers. Early Carboniferous (Middle Visean). • Klubonibelia Conil, 1980 (Type species: Klubonibelia immanis Conil, 1980). The test is streptospirally enrolled in the early stage, later becoming planispiral, with numer- ous sub-q uadrate chambers per whorl. The final stage is uncoiled and rectilinear. Early Carboniferous (Middle to Late Visean). • Latiendothyra Lipina, 1963 (Type species:  Endothyra latispiralis Lipina, in Grozdilova and Lebedeva, 1954). The test is inflated, with a broadly rounded periphery and flattened sides. The early stage is streptospiral, later becoming pla- nispiral, with rapidly enlarging whorls, with a moderate number of chambers per whorl (about six to eight in the final whorl). The septa are short, thick, and slightly oblique, projecting toward the aperture. The wall is calcareous, thick, microgranu- lar, dark, single layered, with secondary deposits that result in septal thickening. The aperture is simple at the base of the apertural face. Early Carboniferous (Late Tournaisian to Early Visean). 70 Evolution and Geological Significance of Larger Benthic Foraminifera • Latiendothyranopsis Lipina, 1977 (Type species:  Endothyra latispiralis Lipina, in Grozdilova and Lebedeva, 1954). The test is streptospiral and enrolled in the early stage, later becoming planispiral, with rapidly enlarging whorl and short, thick, slightly oblique septa. Early Carboniferous (Late Tournaisian to Early Visean). • Laxoendothyra Brazhnikova and Vdovenko, 1972 (Type species:  Endothyra para- kosyensis Lipina, 1955). The test is streptospirally enrolled in the early stage, later is planispiral, with elongate chambers in rapidly enlarging whorls and short septa. Early Carboniferous (Tournaisian to Visean). • Mediopsis Bogush, 1984 (Type species: Planoendothyra? kharaulakhensis Bogush and Yuferev, 1966). The test is involute, streptospirally enrolled in the first half to one and a half whorls, later being planispiral. Early Carboniferous (Late Tournaisian to Visean). • Melatolla Strank, 1983 (Type species: Melatolla whitfieldensis Strank, 1983). The test is streptospirally enrolled in the early stage, later becoming planispiral, and finally tending to uncoil, with massive and robust chomatal deposits in the last chambers. The aperture is areal and cribrate in the uncoiled part. Early Carboniferous (Visean). • Mikhailovella Ganelina, 1956 (Type species:  Endothyrina? gracilis Rauzer- Chernousova, 1948). The test is streptospiral to rectilinear with a short rectilinear part. The aperture is a low basal slit in the early enrolled part, but cribrate and in the uncoiled stage. Carboniferous (Middle Visean) (Plate 2.3, Fig. 19). • Mirifica Shlykova, 1969 (Type species:  Endothyra mirifica Rauzer- Chernousova, 1948. The test is involute, streptospirally enrolled in the early stage, later nearly pla- nispiral with numerous chambers per whorl and short, hooked sutures. The aperture is cribrate in the final whorl. Early Carboniferous (Late Visean). • Omphalotis Shlykova, 1969 (Type species: Endothyra omphalota Rauzer- Chernousova and Reytlinger, in Rauzer- Chernousova and Fursenko, 1937). The test is involute, with early streptospiral coiling, later becoming planispiral. The sutures are radial to slightly oblique. Carboniferous (Middle Visean to Early Bashkirian). • Paradainella Brazhnikova, 1971 (Type species:  Paradainella dainelliformis Brazhnikova and Vdovenko, 1971). The test is streptospirally enrolled in the early stage, but the final whorls are nearly planispiral with secondary deposits forming massive chomata that may cover the inner surface of the whorls. The aperture is sim- ple, basal. Early Carboniferous (Late Tournaisian). • Paraendothyra Chernysheva, 1940 (Type species:  Paraendothyra nalivkini Chernysheva, 1940). The test is enrolled and biumbilicate. The early stage is slightly streptospiral, later becoming nearly completely planispiral. The sutures are radial with a slight serrate appearance, and the final few septa are hook- like in section. The aperture is in a median septal concavity. Early Carboniferous (Middle Tournaisian). • Paraplectogyra Okimura, 1958 (Type species:  Paraplectogyra masanae Okimura, 1958). The test is streptospirally enrolled in the early stage, involute, later becom- ing planispiral with whorls expanding rapidly with straight, radial septa. Early Carboniferous (Late Tournaisian to Early Visean). • Planoendothyra Reytlinger, 1959 (Type species:  Endothyra aljutovica Reytlinger, 1950). The test is compressed, streptospiral in the early stages, later becoming The Palaeozoic Larger Benthic Foraminifera 71 planispiral and evolute, and slightly asymmetrical with supplementary deposits. Carboniferous (Visean to Bashkirian) (Plate 2.9, Fig. 18; Plate 2.12, Fig. 12). • Plectogyra Zeller, 1950 (Type species: Plectogyra plectogyra Zeller, 1950). A strepto- spiral test with a two- layered wall. There are 4 to 12 chambers in the final whorl. Late Devonian/ Early Carboniferous to Permian (Plate 2.3, Fig. 2C; Plate 2.12, figs 1, 3- 5; Plate 2.13, Fig. 10B; Plate 2.14, figs 3, 6- 9, 11, 12). • Plectogyranopsis Vachard, 1977 (Type species:  Endothyra convexa Rauzer- Chernousova, 1948). The test is planispiral, biumbilicate with short, straight, thick septa. Early Carboniferous (Early Visean) to Late Carboniferous (Early Bashkirian). • Pojarkovella Simonova and Zub, 1975 (Type species: Pojarkovella honesta Simonova and Zub, 1975). The test is streptospirally enrolled in the early stage, biumbilicate, tightly coiled and involute, later becoming planispiral with radiate and straight septa and massive secondary deposits consisting of angular chomata. Early Carboniferous (Visean). • Priscella Mamet, 1974 (Type species:  Endothyra prisca Rauzer- Chernousova and Reytlinger, in Rauzer- Chernousova et al., 1936). The test is streptospirally enrolled in the early stage, later becoming nearly planispiral with long strongly oblique septa and no distinct basal secondary deposits. Early Carboniferous (Tournaisian to Bashkirian). • Quasiendothyra Rauzer- Chernousova, 1948 (Type species:  Endothyra kobeitusana Rauzer- Chernousova, 1948). The test is discoidal with concave sides and slightly streptospiral early whorls, with septa as thick as the outer wall. Secondary deposits are chomata-l ike, at each side of the median line against the previous whorl. Late Devonian (Famennian) to Early Carboniferous (Tournaisian). • Rectoendothyra Brazhnikova, 1983 (Type species: Endothyra (Rectoendothyra) don- bassica Brazhnikova, 1983). The test is streptospirally coiled in the early stage, becoming planispiral with the final whorl enlarging rapidly and may tend to uncoil. Septa are straight and radial. Chomata-l ike deposits occur on chamber floor. The aperture is multiple in the final chamber. Early Carboniferous Tournaisian. • Rhodesinella Conil and Longerstaey, 1980 (Type species:  Cribrospira pansa Conil and Lys, 1965). The test is streptospiral enrolled in the early part, later becoming pla- nispiral with a tendency to uncoil in the later stage. Septa moderately short, straight, and inclined toward the aperture. The aperture is multiple a in the final chamber. Early Carboniferous (Tournaisian). • Semiendothyra Reytlinger, 1980 (Type species:  Semiendothyra surenica Reytlinger, 1980). The test is streptospirally enrolled with early whorls, involute, but later slightly evolute, planispiral with slightly arched septa. Secondary deposits are exten- sive, small chomata-l ike mounds near the aperture, spine-l ike in the final chamber. The aperture is simple, basal, and low. Middle Carboniferous (Bashkirian). • Spinoendothyra Lipina, 1963 (Type species: Endothyra costifera Lipina, in Grozdilova and Lebedeva, 1954). The test is closely coiled in the early stage, with later whorls pla- nispiral. Sutures are slightly curved without any thickenings at the ends. Secondary deposits are spine-l ike, well developed on chamber floors, anteriorly directed. The aperture is simple and basal. Carboniferous (Late Tournaisian to Early Visean) (Plate 2.9, Fig. 15). 72 Evolution and Geological Significance of Larger Benthic Foraminifera • Spinothyra Mamet, 1976 (Type species: Endothyra pauciseptata Rauzer- Chernousova, 1948). The test is streptospirally enrolled, involute, biumbilicate, with short thick septa, whorls enlarge rapidly and secondary chomata-l ike deposits occur on the floor of the chambers. Carboniferous (Middle Visean to Early Bashkirian). • Timanella Reytlinger, 1981 (Type species: Endothyra eostaffelloides Reytlinger, 1950). The test is planispirally enrolled with rapidly enlarging whorls and straight oblique septa, with deposits covering the chamber floor and filling the lateral areas. Late Carboniferous (Moscovian). • Tuberendothyra Skipp, 1969 (Type species: Endothyra tuberculata Lipina, 1948). The test is streptospirally enrolled in the early stage, later becoming planispiral with long, curved septa, long and secondary spine-l ike deposits on the floor of all chambers. Early Carboniferous (Late Tournaisian to Early Visean). • Urbanella Malakhova, 1963 (Type species:  Quasiendothyra urbana Malakhova, 1954). The test is streptospirally enrolled in the early whorls, with later whorls increasing in height, becoming planispiral and evolute, with secondary small rounded chomata at the margins of the aperture. Early Carboniferous (Tournaisian to Early Visean). • Zellerinella Mamet, 1981 (Type species: Endothyra discoidea Girty, 1915. The test is streptospirally enrolled in early stage, later becoming planispiral and evolute with pseudochomata. Early Carboniferous (Late Visean to Early Bashkirian). Family Bradyinidae Reytlinger, 1950 The Endothyridae evolved into forms such as Bradyina von Möller, 1878 (see Fig. 2.9). These forms have nautiloid planispiral coiling with few chambers. The microgranular calcareous wall is perforate, with distinct radial lamellae with supplementary septal pores opening into the septal aperture. The primary aperture is also cribrate. They occur in the Carboniferous (Visean) to the Permian (Sakmarian). • Bibradya Strank, 1983 (Type species Bibradya inflata Strank, 1983). The test is strep- tospirally enrolled with short, thick septa which bifurcate close to the outer wall. Early Carboniferous (Late Visean) • Bradyina von Möller, 1878 (Type species: Bradyina nautiliformis Möller, 1878). The test is planispiral and involute with supplementary pores and septal partitions. The wall has a microgranular tectum and keriothecal structure. Carboniferous (Tournaisian to Gzhelian) (Figs 2.7, 2.8; Plate 2.8, Fig. 13; Plate 2.9, Fig. 20; Plate 2.12, Fig. 2; Plate 2.13, Fig. 10A). • Glyphostomella Cushman and Waters, 1928 (Type species: Ammochilostoma? trilocu- lina Cushman and Waters, 1927). The test is planispirally enrolled, involute, with a complex wall, with plates and partitions and large and branching pores in the wall. Late Carboniferous to Permian (Kasimovian to Changhsingian). • Janischewskina Mikhaylov, 1935 (Type species: Janischewskina typica Mikhaylov, 1935). The test is planispirally enrolled, and involute with septal chamberlets as in Bradyina, but without the alveolar structure of Bradyina. Early Carboniferous (Late Visean). • Postendothyra Lin, 1984 (Type species Postendothyra scabra J. X. Lin, 1984). The test is planispirally enrolled and involute, with outer tectum and inner coarsely alveolar keriotheca. Early Permian (Asselian). The Palaeozoic Larger Benthic Foraminifera 73 Family Loeblichiidae Cummings, 1955 Members of this family have atypical endothyroid planispiral tests with numerous chambers, pseudochomata, a basal aperture, two-l ayered walls, an inner clear fibrous light layer (called early tectorium) and a dark granular outer layer (called tectum) (Fig. 2.8A). They range from Middle Devonian to Late Carboniferous (Moscovian). • Dainella Brazhnikova, 1962 (Type species:  Endothyra? chomatica Dain, in Brazhnikova, 1962). The test is streptospiral and involute throughout, with second- ary deposits in the form of massive chomata. Early Carboniferous (Early Visean) • Loeblichia Cummings, 1955 (Type species:  Endothyra ammonoides Brady, 1873). The test is small, flattened, discoidal and is planispiral throughout, with up to 20 sub-r ectangular chambers in the final whorl, but lacks secondary deposits. Early Carboniferous (Visean) (Figs 2.7, 2.8; Plate 2.7, Fig. 7; Plate 2.8, Fig. 12B; Plate 2.13, Fig. 11B). • Lysella Bozorgnia, 1973 (Type species:  Lysella gadukensis Bozorgnia, 1973). The test is streptospirally enrolled in the early stage, later becoming planispiral, invo- lute with many chambers per whorl that appear sub- quadrate in section. Septa are slightly inclined toward the aperture. The wall has chomata-l ike structures. Early Carboniferous (Visean). • Novella Grozdilova and Lebedeva, 1950 (Type species: Novella evoluta Grozdilova and Lebedeva, 1950). The test is planispirally enrolled and evolute with straight septa and well- developed chomata in later whorls. Late Carboniferous (Bashkirian to Early Moscovian). • Seminovella Rauzer- Chernousova, 1951 (Type species:  Eostaffella (Seminovella) elegantula Rauzer- Chernousova, in Rauzer-C hernousova et  al., 1951). The test is planispiral, involute becoming later evolute, with weakly developed pseudochomata. Late Carboniferous (Late Bashkirian to Moscovian). ORDER FUSULINIDA WEDEKIND, 1937 This order includes all larger benthic foraminifera with a homogeneously microgranu- lar primary test made of low-M g calcite, in which the crystal units have no optical align- ment and various foreign particles might be incorporated (see Rauzer- Chernousova, 1948; Loeblich and Tappan 1964; Brazhnikova and Vdovenko 1973; Tappan and Loeblich 1988; Rigaud et al., 2014). Advanced forms have two or more differentiated layers in the wall. They range from the Carboniferous to the Permian. The Fusulinida have special morphological diagnostic characters, which are unique to this order (see Haynes, 1981; Loeblich and Tappan, 1988; and Figs 2.9, 2.10). The traces of the septa on the external surface are called “septal furrows”. These furrows extend from pole to pole of the test and marks the early part of the partitions between the chambers, called “septa”. The apertural face of the test is the “antetheca” while the external wall of the test is the “spirotheca”. Both spirotheca and antetheca are finely perforate by numerous small openings “septal pores” and the test lacks primary apertures. Communication between chambers occurs at the base of septa and aided by resorption of a tunnel in the central part of the test in many fusulinides or of several small circular tunnels, foramina throughout the length of the test in others. In some advanced fusulinides, the antetheca is corrugated into uniformly or irregular spaced 74 Evolution and Geological Significance of Larger Benthic Foraminifera waved, called fluting. Thin sections of the fusulinides reveal highly complicated inter- nal structures which are essential for identifying and classifying the genera and spe- cies. The first chamber is spherical to sub- spherical, called proloculus. Dense calcite, called chomata, were deposited along the margins of the tunnel in many fusulinides, and ridges of dense calcite, parachomata were developed between adjacent foramina in those with multiple foramina). Deposition of dense calcite and simultaneous forma- tions of chomata and parachomata, axial fillings occurred in axial regions of some fusulinides. The spirotheca consists of a thin, dense, primary layer, tectum, which is sometimes covered by layers of tectoria. In more advanced forms, the tectum is aug- mented by a transparent layer, the diaphanotheca, or a thick layer of honeycomb-l ike structure, keriotheca. Ridges, called septula spread down from the lower surface of the spirotheca to subdivide the chambers in the neoschwagerinoids. Since the studies made by Loeblich and Tappan (1964, 1980) and Haynes (1981), many authors have tried to subdivide the fusulinides by considering the highly com- plicated internal structures, such as the wall structure, the presence or absence of the chomata, and the degree of septal folding in the test. On the basis of these main mor- phological features, the Fusulinida are divided here into six superfamilies (see Fig. 2.1 and 2.2), the Fusulinoidea, the Ozawainelloidea, the Staffelloidea, the Schubertelloidea, the Fusulinoidea, the Schwagerinoidea, and the Neoschwagerinoidea (= Verbeekinoidea). These superfamilies are considered as orders by Rauzer-C hernousova et al. (1996) and Mikhalevich (2013). The structure of the spirotheca plays an important role in differentiating the fusulinids. Five types of wall structures (see Fig. 2.10) characterise these superfamilies: 1. Forms with two layered walls, an inner clear fibrous light layer (called early tectorium) and a dark granular outer layer (called tectum) (Fig. 2.10A) Tectum Wall coarsely alveolar = (dark with organic matter) Keriotheca Inner “granulo-fibrous” layer Fig. 2.9. Schematic figures showing a microgranular wall and a Bradyina-t ype of wall with keriotheca, mul- tiple tunnels and parachomata in more advanced forms. The Palaeozoic Larger Benthic Foraminifera 75 Types A-D, e.g. Ozawainellidae, Staffellidae, Fusulinidae, Schubertellidae Tectum A Upper tectorium Lower tectorium Tectum B Lower tectorium Upper tectorium Tectum C Tectum Diaphanotheca D Lower tectorium Diaphanotheca Lower tectorium E E.g. Schwagernidae, Verbeekinidae, Neoschwagerinidae Tectum keriotheca Spirotheca Septal furrows Antetheca “Septal pores” = antethecal pores Septum Tunnel “Chamberlets” Septum (fluted) Tectum Chomata Lower tectorium Tunnel Upper tectorium Fig. 2.10. Schematic figures showing important features of the fusulinides. 76 Evolution and Geological Significance of Larger Benthic Foraminifera 2. Forms with three layered walls, the late and early tectorium surrounding the dark layer of the tectum (Fig. 2.10B); 3. Forms with four layered walls, with the late and early tectorium surrounding the tectum and a glossy layer called the diaphanotheca (Fig. 2.10C); 4. Forms with three layered walls, where the late tectorium disappears and they are left with the tectum, diaphanotheca and the early tectorium (Fig. 2.10D); 5. Forms composed of a tectum and glossy layer with alveoles running through it (the keriotheca) (Fig. 2.10E). The first four types of wall characterise the ozawainelloids, staffelloids, and the fusulinoids and the most advanced final type of wall belong to the schwagerinoids and neoschwagerinoids (discussed below). Over 100 fusulinides genera are recognised. Below are the main taxonomic classifi- cations of the fusulinides, but for an even more detailed taxonomic description at the generic and specific levels see Loeblich and Tappan (1988) and Rauser- Chernousova et al. (1996). Superfamily OZAWAINELLOIDEA Slovieva, 1978 The test is small, lenticular to rhombic to discoid and involute. The early coiling may be streptospiral. The wall is simple in the early forms, composed of tectum with upper and lower tectoria (Fig. 2.10B), but with diaphanotheca occurring between tectum and lower tectorium (Fig. 2.10C) in advanced forms. The proloculus is small and spheri- cal. The pseudochomata develop into weak to strong chomata. The aperture is simple and basal. This superfamily ranges from the Early Carboniferous to the Late Permian (Visean to Capitanian). Family Ozawainellidae Thompson and Foster, 1937 Here, the tests are discoidal to elongate in shape. They are involute to slightly evo- lute with moderate to strong secondary deposits of calcite (chomata) (Fig. 2.11). The wall is of a primitive fusulinid type, with a thin dense, primary layer (tectum) dark with organic matter, surrounded by a thicker, less dense layer of upper and lower tectoria (Fig. 2.10B). Carboniferous to the Late Permian (Late Tournaisian to Capitanian). This family includes genera such as: • Chenella Miklukho- Maklay, 1959 (Type species:  Orobias kueichihensis S. Chen, 1934). The test is planispirally enrolled, with the final whorl increasing abruptly in height. Late Permian (Capitanian). • Eoparastaffella Vdovenko, 1954 (Type species. Parastaffella (Eoparastaffella) sim- plex Vdovenko, 1954). Test is almost planispiral. The wall is of two very thin, and dark microgranular tecta surrounded by thick, pale brown-g ray zone. Early Caroboniferous (Late Tournaisian to Early Visean). • Eostaffella Rauzer- Chernousova, 1948 (Type species:  Staffella (Eostaffella) para- struvei Rauzer- Chernousova, 1948). The test is lenticular, involute and planispiral with a sub- acute periphery. Septa are thick, curved, and often with truncated edges. The Palaeozoic Larger Benthic Foraminifera 77 The wall is layered with tectum and upper and lower tectoria. There are discontin- uous knobs on either side of the tunnel but no continuous chomata. Carboniferous to Permian (Visean to Capitanian) (Plate 2.3, Fig. 4A; Plate 2.5, Fig. 19; Plate 2.9, Fig. 5; Plate 2.12, figs 4, 9, 13, 14; Plate 2.14, figs 14, 15; Plate 2.15, Fig. 4). • Eostaffelloides Miklukho- Maklay, 1959 (Type species:  Eostaffelloides orientalis Miklukho- Maklay, 1959). The test is lenticular, planispirally enrolled with second- ary deposits forming triangular chomata. Late Permian (Capitanian). • Millerella Thompson, 1942 (Type species: Millerella marblensis Thompson, 1942; see Fig. 2.9). Minute discoidal and partially evolute test. The septa are slightly arched forward. Carboniferous (Serpukhovian to Gzhelian) (Fig. 2.13; Plate 2.3, Fig. 2A; Plate 2.12, Fig. 6; Plate 2.14, Fig. 10; Plate 2.8, Fig. 11). • Ozawainella Thompson, 1935 (Type species: Fusulinella angulata Colani, 1924). The test is angular with massive chomata. Late Carboniferous to Permian (Serpukhovian to Capitanian). • Pamirina Leven, 1970 (Type species:  Pamirina darvasica Leven, 1970). The test is small and varies in shape from nautiloid to subspherical. The spirotheca is thin, one- layered in inner whorls and two- layered consisting of tectum and protheca in later whorls. The septa are straight and flat. The chomata vary from rudimentary to well developed. Early Permian (Sakmarian to Kungurian). • Pseudonovella Kireeva, 1949 (Type species: Pseudonovella irregularis Kireeva, 1949). The test is lenticular with early whorls evolute and final whorl increasing rapidly in height and becoming partially evolute. The pseudochomata are weakly developed. Late Carboniferous (Late Bashkirian to Moscovian). chomata Lower tectorium Upper Tectum tectorium Diaphanotheca Septa not fluted Millerella Ozawainellidae Staffella Staffellidae Diaphanotheca Schubertella Schubertellidae Fig.  2.11. Schematic figures showing the features of the Ozawainellidae, Staffellidae, Schubertellidae. Drawings not to scale. 78 Evolution and Geological Significance of Larger Benthic Foraminifera • Quasireichelina Ueno 1992 (Type species: Quasireichelina expansa Ueno, 1992). The test is involute in the early part, but later is uncoiled with numerous unfluted septa. Secondary deposits form small chomata. Middle Permian (Guadalupian). • Reichelina Erk, 1942 (Type species:  Reichelina cribroseptata Erk, 1942). The test is planispirally enrolled in the early stage, but the final whorl uncoils to appear almost peneropliform. Secondary deposits consist of broad chomata, extending and thickening poleward from the tunnel or the small circular foramina. Late Permian (Capitanian). • Sichotenella Tumanskaya, 1953 (Type species: Sichotenella sutschanica Tumanskaya, 1953). The test is lenticular, resembling Chenella, but the final whorl increases rap- idly in height and uncoils. The uncoiled stage is relatively large. The wall is three layered, with diaphanotheca. Late Permian (Capitanian). • Zarodella Sosnina, 1981 (Type species: Zarodella zhamoidai Sosnina, 1981). The test is small, slightly biumbilicate. The proloculus is large in size compare to the adult test which comprises three to five whorls. The septa and flat and flat and the spi- rotheca is weakly differentiated. Early Permian (Sakmarian). Family Pseudostaffellidae Putrya 1956 The Pseudostaffellidae have an involute test which may be a streptospiral in the early stage. Early walls are weakly differentiated, later ones have a tectum, thin diaphano- theca and two tectoria (Fig. 2.10D) with well- developed chomata. Carboniferous to Permian (Serpukhovian to Capitanian). • Hubeiella Lin, 1977 (Type species:  Hubeiella simplex J.  X. Lin, 1977). The test is planispirally enrolled and involute in the early stage, later stages may be evolute, and the final whorl is greatly expanded in height and flares. The septa are straight, and the wall shows a tectum and fine keriotheca with small but distinct chomata present only in the inner whorls. Early Permian (Serpukhovian). • Kangvarella Saurin, 1962 (Type species: Kangvarella irregularis Saurin, 1962). The test is involute in the early part, and later whorls increase rapidly in height. The wall has a tectum, diaphanotheca, and inner and outer tectoria. Irregular and asymmet- rical parachomata may be present in the second and third whorls. Late Permian (Capitanian) • Mediocris Rozovskaya, 1961 (Type species:  Eostaffella mediocris Vissarionova, 1948). The test is streptospirally enrolled in the early part, later becoming planispiral and involute. The wall has a dark tectum and a poorly defined tectoria, with strong secondary deposits in the axial region. Carboniferous (Middle Visean to Early Moscovian). • Neostaffella Miklukho- Maklay, 1959 (Type species:  Pseudostaffella sphaeroidea Ehrenberg, in Rauzer-C hernousova et al., 1951). Chomata are very broad, extend- ing from the edge of the narrow tunnel outward to the umbilical region. Late Carboniferous (Moscovian) (Plate 2.16, Fig. 9). • Ninella Malakhova, 1975 (Type species:  Endothyra staffelliformis Chernysheva, 1948). The test is streptospirally enrolled in the early part, later becoming nearly planispiral, with straight septa and widely spaced rounded pseudochomata in all whorls. Early Carboniferous (Visean). The Palaeozoic Larger Benthic Foraminifera 79 • Pseudostaffella Thompson, 1942 (Type species:  Pseudostaffella needhami Thompson, 1942). The test is streptospirally coiled in early whorls, later becoming planispiral with straight, long, unfluted and perpendicular to the outer wall septa, slightly curved in the polar regions. Chomata are well developed. Late Carboniferous (Bashkirian to Moscovian) • Rauserella Dunbar, 1944 (Type species: Rauserella erratica Dunbar, 1944). The test is planispirally enrolled in the early whorls. The whorls become much wider and the test fusiform with irregular septa in outer whorls. Septal pores are present in the outer whorls. Middle Permian (Kungurian) • Toriyamaia Kanmera, 1956 (Type species:  Toriyamaia latiseptata Kanmera, 1956). The test is fusiform, the first whorl is evolute, later whorls are sub-c ylindrical, involute, increasing rapidly in height. Septa are sparse and unfluted. Early Permian (Cisuralian). Superfamily STAFFELLOIDEA Slovieva, 1978 emend. Vachard et al., 2013 In this order, the tests are small, spherical to discoidal and strong chomata are present. The wall structure of this order has been discussed by many authors, but is still ill- resolved (see Jin-Z hang, 1990). The wall is usually composed of tectum and diaphano- theca (Fig. 2.11) and may have tectoria, but it is mainly found recrystallized and altered by secondary mineralisation, suggesting that they secreted an aragonite or high-M g calcite wall (Groves et  al., 2005). Early Carboniferous to Late Permian (Visean to Wuchiapingian). Family Staffellidae Miklukho- Maklay, 1949 Staffellidae have a primitive wall structure, consisting of a tectum followed by a trans- parent layer (diaphanotheca) which is formed of clear calcite (Fig. 2.11). They occur from the Early Carboniferous to Late Permian (Visean to Wuchiapingian). • Chenia Sheng, 1963 (Type species: Chenia kwangsiensis Sheng, 1963). The test is len- ticular with plane, unfluted septa. Chomata are well developed in all whorls with poorly developed and discontinuous parachomata present in the outer four or five whorls. Middle Permian (Kungurian). • Nankinella Lee, 1934 (Type species: Staffella discoides Lee, 1931). The test is discoi- dal with plane septa and distinct chomata. Early to Late Permian (Sakmarian to Wuchiapingian). • Pisolina Lee, 1934 (Type species: Pisolina excessa Lee, 1934). The test is spherical with thick, plane septa and well defined, low, asymmetrical chomata. Early Permian (Asselian to Kungurian). • Pseudoendothyra Mikhaylov, 1939 (Type species:  Fusulinella struvii von Möller, 1879). The test is lenticular, compressed, planispiral and involute, with plain septa and supplementary deposits adjacent to the aperture but not forming continuous chomata. The aperture is low and broad. Early Carboniferous to Early Permian (Visean to Sakmarian). (Plate 2.9, figs 2, 13; Plate 2.12, Fig. 15A; Plate 2.14, figs 2, 13). • Staffella Ozawa, 1925 (Type species: Fusulina sphaerica Abich, 1859). It is the main representative of this group. It has a small lenticular test with a wall composed of a tectum and diaphanotheca. Permian (Artinskian to Kungurian) (Fig. 2.11). 80 Evolution and Geological Significance of Larger Benthic Foraminifera Family Kahlerinidae Leven, 1963 This family is monotypic; the test is globose with minor chomata and rarely with small parachomata. Middle Permian. • Kahlerina Kochansky- Devidé and Ramovš, 1955 (Type species: Kahlerina pachytheca Kochansky-D evidé and Ramovš, 1955). Septa are long, planar and unfluted. Chomata are only rudimentarily present in the initial whorls. Middle Permian (Guadalupian). Superfamily SCHUBERTELLOIDEA Vachard in Vachard et al., 1993 The shell is minute, fusiform or irregular in shape. The wall is composed of tectum with lower tectorium only (Fig. 2.10A), or tectum surrounded by an upper and lower tectoria (Fig.  2.7B), or tectum with diaphanotheca and tectoria (Fig.  2.10C). Late Carboniferous to Late Permian (Bashkirian to Changhsingian). Family Schubertellidae Skinner, 1931 Schubertellidae have fusiform tests with a streptospiral early coiling. Early septa are flat but evolve into fluted septa in advanced forms. The wall varies, as in some forms the diaphanotheca is present (Fig. 2.10B), while in others only the tectum and early tectorium are present (see Fig.  2.10A; Fig.  2.11). This family includes spirothecal structures of the post- keriotheca phase. Carboniferous to Permian (Bashkirian to Changhsingian) Significant genera include: • Biwaella Morikawa and Isomi, 1960 (Type species: Biwaella omiensis Morikawa and Isomi, 1960). The test is large, fusiform to sub-c ylindrical, with subglobose and tightly coiled initial whorls followed by ovoid whorls that increase rapidly in size. The wall in first whorls is composed of a thin, dark tectum and a thicker but lighter tectorium. The wall in the final whorl is perforated with coarse mural pores, which do not develop into keriotheca. The septa are nearly straight and widely separated. The chomata are small and prominent in all the test, except for the final whorl. (Davydov, 2007; 2010). Late Carboniferous to Early Permian (Early Gzhelian to Kungurian). • Boultonia Lee, 1927 (Type species: Boultonia willsi Lee, 1927). The test is fusiform, with strongly fluted septa throughout and asymmetric chomata. Early Permian (Asselian to Sakmarian). • Codonofusiella Dunbar and Skinner, 1937 (Type species: Codonofusiella paradoxica Dunbar and Skinner, 1937). The test is irregularly coiled, becoming uncoiled in later stages, small, fusiform, with strongly fluted septa. Chomata are present but tunnels are not clearly defined. Late Permian (Lopingian). • Dunbarula Ciry, 1948 (Type species: Dunbarula mathieui Ciry, 1948). The test is an elongate ellipsoid, irregularly coiled, with septa strongly fluted throughout. Late Permian (Capitanian). • Dutkevichites Daydov, 1984 (Type species: Dutkevichites darvasica Davydov, 1984). The test is cylindrical. Septa are few, but strongly fluted with an increase in intensity towards the poles. Minute chomata are present in outer whorls. Late Carboniferous to Early Permian (Gzhelian to Asselian) (Fig. 2.14). The Palaeozoic Larger Benthic Foraminifera 81 • Eoschubertella Thompson, 1937 (Type species: Schubertella lata Lee and Chen, in Lee et al., 1930). The test is elongate with early whorls close coiled and showing a broad tunnel, bordered by low chomata. The wall is unilayered. Late Carboniferous (Moscovian). • Fusiella Lee and Chen, 1930 (Type species:  Fusiella typica Lee and Chen, in Lee et al., 1930). The test is elongate, early whorls are discoidal and endothyroid, later becoming fusiform with axial fillings prominent. Late Carboniferous (Moscovian) (Plate 2.15, Fig. 6). • Gallowaiina Chen, 1934 (Type species: Gallowaiina meitiensis Chen, 1934). The test is fusiform with very thin septa, closely fluted throughout, with folds extending over most of the chamber height. No chomata are present, fillings only found in the axial region. Late Permian (Changhsingian). • Grovesella Davydov and Arefifard, 2007 (Type species: Grovesella tabasensis Davydov and Arefifard, 2007). Test: Discoidal, nearly planispirally coiled with a large proloc- ulus and two-l ayered poorly developed wall lacking chomata. Late Carboniferous to Permian (middle Bashkirian up to Wordian). • Kwantoella Sakagami and Omata, 1957 (Type species: Kwantoella fujimotoi Sakagami and Omata, 1957). The test is fusiform with numerous, straight and unfluted septa. Chomata are poorly developed, and axial fillings are more extensive in later whorls. Early Permian (Sakmarian). • Lantschichites Tumanskaya, 1952 (Type species:  Codonofusiella (Lantschichites) maslennikovi Tumanskaya, 1953). The test is cylindrical with strongly developed and fluted septa. Late Permian (Lopingian). • Mesoschubertella Kanuma and Sakagami, 1957 (Type species:  Mesoschubertella thompsoni Sakagami, in Kanuma and Sakagami, 1957). The test is fusiform with straight to slightly fluted septa in the polar ends. The wall is thick and composed of a thin, dark tectum and a diaphanotheca between a well- developed upper tectorium and a lower tectorium. Chomata are small, but well developed and asymmetrical. Early Permian (Cisuralian to Guadalupian). • Minojapanella Fujimoto and Kanuma, 1953 (Type species: Minojapanella elongata Fujimoto and Kanuma, 1953). The test is subcylindrical, with intensely fluted septa, but without developing cuniculi (i.e. multiple tunnels, see Fig.  2.9). Chomata are massive. Late Permian (Capitanian to Changhsingian) (see Fig. 2.14). • Nanlingella Rui and Sheng, 1981 (Type species:  Nanlingella meridionalis Rui and Sheng, 1981). The test has a small proloculus. Septa are strongly fluted except in the median part. The wall is thin and composed of a tectum and diaphanotheca. Chomata are present only in the first two whorls. Late Permian (see Fig. 2.14). • Neofusulinella Deprat, 1912 (Type species: Neofusulinella praecursor Deprat, 1913). The test is ovoid in the early stage, later becoming fusiform, planispirally coiled throughout with flat, slightly fluted septa. Chomata are large and asymmetrical. Late Carboniferous (Early Moscovian). • Neoschubertella Saurin, 1962 (Type species:  Neoschubertella sisophonensis Saurin, 1962). The test is fusiform, with slightly curved but not fluted septa and continuous pseudochomata. Early Permian to Late Permian (Sakmarian to Capitanian). • Palaeofusulina Deprat, 1912 (Type species: Palaeofusulina prisca Deprat, 1913). The test has the spirotheca composed of a tectum and an early transparent diaphanotheca. 82 Evolution and Geological Significance of Larger Benthic Foraminifera The Palaeofusulina- type fusulinids became abundant after most of the Schwagerina- Verbeekina-N eoschwagerina types had disappeared (Jin- Zhang, 1990). Lattermost Permian (Changhsingian) (see Fig. 2.14). • Paradoxiella Skinner and Wilde, 1955 (Type species: Paradoxiella pratti Skinner and Wilde, 1955). The test is discoidal, flaring and uncoiling with intensely fluted septa, and with well- developed cuniculi and low chomata. Late Permian (Capitanian). • Paradunbarula Skinner, 1969 (Type species: Paradunbarula dallyi Skinner, 1969). The test is fusiform with intensely fluted septa and moderately wide, weak chomata. Late Permian (Capitanian) • Russiella Miklukho- Maklay, 1957 (Type species:  Russiella pulchra Mildukho- Maklay, 1957). The test is fusiform, and the first two whorls endothyroid. Septa are strongly fluted, with heavy axial fillings. Chomata are weakly developed except in the last two revolutions. Late Permian (Capitanian to Changhsingian). • Schubertella Staff and Wedekind, 1910 (Type species: Schubertella transitoria Staff and Wedekind, 1910). The test is small becoming fusiform in later stages, with straight to weakly fluted septa at the polar ends. The wall is differentiated into three layers that are penetrated by relatively coarse pores. Chomata are prominent, low and asymmetrical. Late Carboniferous to Permian (Moscovian to Wordian) (see Fig. 2.11; Plate 2.17, Fig. 1). • Schubertina Marshall, 1969 (Type species: Schubertina circuli Marshall, 1969). The test is subglobose to ovoid in outline, with a thin wall clearly differentiated into two layers, poorly developed secondary deposits and straight septa. Late Carboniferous to Permian (Bashkirian to Lopingian). • Sphaeroschwagerina Miklukho- Maklay, 1959 (Type species: Schwagerina sphaerica var. karnica Shcherbovich, in Rauzer-C hernousova and Shcherbovich, 1949). The test is almost spherical with weakly fluted septa at the poles and small chomata. Permian (Plate 2.18, figs 6, 7, 9). • Yangchienia Lee, 1934 (Type species: Yangchienia iniqua Lee, 1934). The test is fusi- form with plane unfluted septa and massive, asymmetrical chomata that extend nearly to the poles. Late Permian (Asselian to Capitanian). Superfamily FUSULINOIDEA von Möller, 1878 The test of the Fusulinoidea is large and varies from spherical to fusiform. The devel- opment of secondary deposits of calcite (chomata) is prominent in most forms, but the numerous chambers are subdivided by folds or septula (Figs 2.10, 2.12). Late Carboniferous to Permian (Bashkirian to Kungurian). Family Fusulinidae von Möller, 1878 In the Fusulinidae, the test is spherical or elongate fusiform, and mostly planispiral. The Fusulinidae family belongs to the pre- keriotheca stage of fusulinides development. It includes genera with a wall composed of a tectum surrounded by a late and an early tectorium (Fig. 2.10B) (e.g., the type of wall typical for Profusulinella, see Fig. 2.12). However, it also includes more evolved genera such as the cylindrical Fusulina and the fusiform Beedeina (see Fig. 2.12) with four layered walls, with the late and early tecto- rium surrounding the tectum and a glossy layer called the diaphanotheca (Fig. 2.10C). Carboniferous to Permian (Bashkirian to Kungurian). The Palaeozoic Larger Benthic Foraminifera 83 Upper and lower Chomata Tectoria Fluted septa at poles Profusulinella Fusulinella Four layers with Diaphanotheca Beedeina Fig. 2.12. Schematic drawings showing the features of the fusulinides (not to scale). Significant genera include: • Beedeina Galloway, 1933 (Type species: Fusulinella girtyi Dunbar and Condra, 1928). The test is fusiform with plane to weakly fluted septa in the early whorls, later septa are strongly fluted. Secondary fillings also coat the septa. Chomata are massive, high and broad in the early stage, but lower in the adult. Late Carboniferous (Moscovian) (Fig. 2.12). • Eofusulina Rauzer- Chernousova, 1951 (Type species:  Fusulina triangula Rauzer- Chernousova and Belyaev, in Rauzer-C hernousova et a1.,1936). The test is fusiform with septa strongly fluted throughout, with high and narrow folds. Carboniferous (Early Moscovian). • Eowedekindellina Ektova, 1977 (Type species: Eowedekindellina fusiformis Ektova, 1977). The test is fusiform with fluted septa, mainly in final whorl. Chomata are extended poleward, forming a basal layer. Late Carboniferous (Bashkirian). • Fusulina Fisher de Waldheim, 1829 (Type species:  Fusulina cylindrica Fischer de Waldheim, 1830). The test is cylindrical with a spirotheca composed of a tectum, diaphanotheca and thick late and early tectoria. The septa are fluted throughout the test but more intensely near the polar ends of the test and the chomata are strong. Late Carboniferous (Moscovian) (Plate 2.8, figs 6, 9; Plate 2.19, Fig. 2). • Fusulinella von Möller, 1877 (Type species: Fusulinella bocki von Möller, 1878). The test is fusiform with the septa being fluted in the polar region. Chomata are promi- nent and asymmetrical. Late Carboniferous (Moscovian) (Figs 2.12, 2.13). • Hemifusulina von Möller, 1877 (Type species: Hemifusulina bocki von Möller, 1878). The test has its early whorls closely coiled, later they become more loosely coiled. Septa are slightly to moderately fluted. Chomata are rounded and well developed. Middle Carboniferous (Moscovian). 84 Evolution and Geological Significance of Larger Benthic Foraminifera Period, Epoch, Fusulinindes phylogenies Events Stage Palaeofusulina Conofusiella Post - Keriotheca Phase Nanlingella Chusenella Minojapanella Taiyuanella Stalactotheca stage Parafusulina Triticites Chusenella Anthotheca stage Dutkevichites Pseudofusulina Montiparus Quasifusulina Keriotheca sensu stricto stage Obsoletes Triticites Diaphonetheca stage Protriticites Fusulinella Pre-keriotheca Profusulinella Tectum stage Fig.  2.13. The stratigraphic occurrence of the keriotheca, antetheca and stalactotheca stages (modified after Vachard et al., 2004). • Obsoletes Kireeva, 1950 (Type species: Fusulina obsoleta Schellwien, 1908). The test is elongate fusiform with tightly coiled early whorls becoming rapidly enlarged in the last whorls. Septa are fluted at the poles. Chomata are well developed and wide. Late Carboniferous (Kasimovian) (Fig. 2.13). • Paraeofusulina Putrya, 1956 (Type species: Eofusulina (Paraeofusulina) triangulifor- mis Putrya, 1956). The test is elongate fusiform with strongly fluted septa forming strongly arched loops and poorly differentiated pseudochomata in the early whorls. Late Carboniferous (Moscovian). • Parafusulinella Stewart, 1970 (Type species: Parafusulinella propria Stewart, 1970). The test is small, fusiform with septa undulating in the central region and slightly fluted towards the apices. Late Carboniferous (Moscovian). • Profusulinella Rauzer- Chernousova and Belyaev, 1936 (Type species: Profusulinella pararhomboides Rauzer- Chernousova and Belyaev, in Rauzer-C hernousova et  al., 1936). The test is fusiform with septa fluted in the polar regions. Chomata are prom- inent, asymmetrical with secondary thickening extending laterally towards the poles. Late Carboniferous (Middle Bashkirian to Early Moscovian) (Figs 2.12, 2.13; Plate 2.8, figs 1- 4). • Protriticites Putrya, 1948 (Type species:  Protriticites globulus Putrya, 1948). The test is fusiform, and septa are moderately fluted in the axial region. The wall has four- layers, with massive asymmetrical chomata. Late Carboniferous (lattermost Moscovian to Kasimovian) (Fig. 2.13). • Quasifusulina Chen, 1934 (Type species:  Fusulina longissima Moeller, 1878). The test is elongate with poles bluntly rounded. Septa are intensely fluted and may form cuniculi. The wall is very thin, with a tectum and diaphanotheca and a poorly defined tectorium. Chomata are weakly developed, axial filling is heavy. Late Carboniferous (Moscovian) to Early Permian (Sakmarian) (Plate 2.15, Fig. 8; Plate 2.16, figs 3, 5, 8, 10; Plate 2.20, 2- 3; Fig. 2.13). Carboniferous Permian Pennsylvanian Keriotheca Phase The Palaeozoic Larger Benthic Foraminifera 85 Chomata well marked in the juvenarium Tectum Keriotheca Change of Chomata growth rate Fluted septa Triticites Pseudoschwagerina Schwagerinidae Penultimateantethca “septum” Ultimate Cuniculi Axial antetheca Resorbed salient fillings Antethecal salient Cuniculi Parafusulina Polydiexodina Multiple tunnels at the base of the septa Parachomata Tectum Keriotheca Transverse septulae Axial septula Verbeekina Foramen Verbeekinidae Foramen Multiple tunnels Yabeina Parachomata Neoschwagerinidae Fig. 2.14. Schematic features of the advanced forms of fusulinides: the Schwagerinidae, the Verbeekinidae and the Neoschwagerinidae. • Skinnerella Coogan, 1960 emend. Skinner, 1971 (Type species: Parafusulina schuberti Dunbar and Skinner, 1937). The test is elongate with strong secondary deposits that may completely fill the chambers of the early whorls and tend to spread away from the axis to outer whorls. Low cuniculi are well developed. Early Permian (Kungurian). • Wedekindellina Dunbar and Henbest, 1933 (Type species: Fusulinella euthusepta Henbest, 1928). The test is fusiform with straight septa in the central area of the test, but may be slightly fluted toward the extremities. Late Carboniferous (Moscovian). Superfamily SCHWAGERINOIDEA Vachard, Pille and Gaillot, 2010 The schwagerinoids are large and fusiform to irregularly cylindrical. The test is planispi- rally enrolled, involute to irregularly coiled. The spirotheca is thick, composed of tectum and alveolar keriotheca (Fig. 2.10E). The septa are fluted in last whorls in primitive gen- era and completely across the test, forming multiple tunnels or cuniculi in more advanced genera (Fig. 2.14). Late Carboniferous to Permian (Kasimovian to Capitanian). Family Schwagerinidae Dunbar and Henbest, 1930 The Schwagerinidae have tests that are large, fusiform to sub- cylindrical, planispiral and involute. They have spirothecal structures showing a tectum and a glossy layer, keriotheca (including a finely or coarsely alveolar) phase (Fig.  2.10E). In the most 86 Evolution and Geological Significance of Larger Benthic Foraminifera advanced forms, such Polydiexodina Dunbar and Skinner 1931, the folded septa with a reduced single-l ayer wall have multiple tunnels or cuniculi throughout the test and heavy fillings along the axis (see Fig. 2.14; Plate 2.19; Plate 2.21; Plate 2.22). During the early phase of the keriotheca, the Schwagerinidae possess chomata that range from thin to massive (Figs 2.14– 2.16). The intensity of the folded septa varies from genus to genus. Late Carboniferous to Permian (Kasimovian to Capitanian). Significant genera include: • Carbonoschwagerina Ozawa, Watanabe and Kobayashi, 1992 (Type spe- cies:  Pseudoschwagerina morikawai Igo, 1957). The test is subspherical, septa are closely spaced, and weakly fluted. Chomata are distinct, and massive in inner whorls. This form has a smaller proloculus and much heavier chomata than in Pseudoschwagerina. Late Carboniferous (Gzhelian). • Chalaroschwagerina Skinner and Wilde, 1965 (Type species:  Chalaroschwagerina inflata Skinner and Wilde, 1965). The test is small fusiform, septa are strongly fluted in the early part, but merely wavy in the late part. Weak chomata occur on the pro- loculus. Early Permian (Artinskian to Kungurian). • Chusenella Hsu, 1942 (Type species: Chusenella ishanensis Hsu, 1942). The test is fusiform, and later septa are tightly fluted. The wall has a tectum and keriotheca. Chomata may be lacking throughout but axial filling is prominent. Early to Middle Permian (Sakmarian to Capitanian) (Fig. 2.13). • Cuniculinella Skinner and Wilde, 1965 (Type species: Cuniculinella tumida Skinner and Wilde, 1965). The test is fusiform, and septa are intensely fluted with folds reach- ing the top of the septa. Those of adjacent septa join to form chamberlets. Chomata are present only on the proloculus. Permian (Artinskian to Kungurian). • Daixina Rozovskaya, 1949 (Type species: Daixina ruzhencevi Rozovskaya, 1949). The test is fusiform, having septa with strong and irregular fluting. Chomata are rarely pre- sent in the early whorls. Late Carboniferous to Early Permian (Gzhelian to Asselian) Chomata tunnel Proloculus keriotheca Septa Fig. 2.15. Morphological features of an axial section of Triticites. Wolfcamp Beds, Wolfcamp Hills, N.E. Marathon Texas, UCL collection. The Palaeozoic Larger Benthic Foraminifera 87 • Dunbarinella Thompson, 1942 (Type species:  Dunbarinella ervinensis Thompson, 1942). The test is fusiform with a straight axis of coiling. It differs from Schwagerina and Pseudofusulina in the presence of heavy axial fillings and distinct chomata. Late Carboniferous to Early Permian (Gzhelian to Sakmarian) (Plate 2.8, Fig. 5; Plate 2.16, Fig. 11; Plate 2.23, Fig. 8; Plate 2.24, Fig. 4). • Eoparafusulina Coogan, 1960 (Type species:  Fusulina gracilis Meek, 1864). The test is subglobular, tightly coiled and expanding uniformly. Late parts of the septa are planar, while early parts are strongly fluted throughout the test, producing chamberlets, with low tunnel bordered by pseudochomata. Rudimentary cho- mata are present in in first whorls. Early Permian (Late Asselian to Artinskian and?Kungurian). • Eopolydiexodina Wilde, 1975 (Type species: Polydiexodina afghanensis Thompson, 1946). Septa are intensely fluted throughout the entire length and in all whorls of the test, with cuniculi well developed. It differs from Polydiexodina in lacking a median tunnel and in having only sporadic supplementary tunnels. Permian (Middle Guadalupian). • Monodiexodina Sosnina, 1956 (Type species: Schwagerina wanneri var. sutschanica Dutkevich, 1934). The test is subcylindrical, with septal fluting with folds restricted to the early part, forming low cuniculi. Early whorls have chomata. Early to Middle Permian (Cisuralian to Late Guadalupian). • Montiparus Rosovskaya, 1948 (Type species: Alveolina montipara Ehrenberg, 1854, emend. Möller, 1878). The test is fusiform. Septa are irregularly plicate. Chomata are sub- quadrate. Carboniferous (Kasimovian) (Fig. 2.13; Plate 2.8, Fig. 7). • Parafusulina Dunbar and Skinner, 1931 (Type species:  Parafusulina wordensis Dunbar and Skinner, 1931). The test is very large with intense folding but with no chomata (Fig. 2.9). Middle to Late Permian (Guadalupian to Lopingian) (Fig. 2.14; Plate 2.16, figs 1- 2; Fig. 2.19, figs 5, 7). • Paraschwagerina Dunbar and Skinner, 1936 (Type species:  Schwagerina gigantea White, 1932. The test is fusiform to subspherical. The septa are distinctly fluted throughout test, with high, strong, and parallel sided folds. The wall is thin, com- posed of thin tectum and coarsely alveolar keriotheca. The chomata are small in the first whorls, but absent in the later whorls. Early Permian (Asselian to Sakmarian) (Plate 2.17, figs 4,6; Plate 2.18, Fig. 10). • Polydiexodina Dunbar and Skinner, 1931 (Type species: Polydiexodina capitanensis Dunbar and Skinner, 1931). Septa are numerous, regularly and strongly fluted, with narrow septal folds, the fluting of adjacent septa producing cuniculi where in contact. Permian (Late Guadalupian) (Fig. 2.14; Plate 2.17, Figs 3, 7). • Pseudofusulina Dunbar and Skinner, 1931 (Type species: Pseudofusulina huecoensis Dunbar and Skinner, 1931). The test is fusiform with a large proloculus and fluted septa, most strongly towards the poles where it is densely folded to produce chamber- lets. Late Carboniferous to Permian (lattermost Gzhelian to Kungurian) (Fig. 2.13; Plate 2.18, Fig. 13). • Pseudoschwagerina Dunbar and Skinner, 1936 (Type species:  Schwagerina uddeni Beede and Kniker, 1924). The test has inflated outer whorls and strongly folded septa throughout (Fig. 2.9). Chomata are well marked at juvenile stage only. They 88 Evolution and Geological Significance of Larger Benthic Foraminifera are rudimentary or lacking in the final stage. Early to Middle Permian (Asselian to Kungurian) (Fig. 2.14; Plate 2.19, Fig. 5B; Plate 2.18, figs 3, 4, 8, 11; Pate 2.22, figs 3, 5). • Schwagerina von Mölle, 1877 (Type species: Borelis princeps Ehrenberg, 1842). Septa are regularly and intensely fluted, and folds of septa divide the early part of the cham- bers into small chamberlets. Chomata are weak or absent, with narrow tunnels. Early Permian (Sakmarian to Artinskian) (Plate 2.15, Fig. 5; Plate 2.16, figs 4,6; Plate 2.17, Fig. 5A; Plate 2.20, figs 4, 5; Plate 2.23, figs 3-4 , 6; Plate 2.24, Fig. 6; Plate 2.25, figs 1, 3- 4). • Skinnerina Ross, 1964 (Type species:  Skinnerina typicalis Ross, 1964). Septa are numerous and intensely fluted from pole to pole, resulting in high and well devel- oped cuniculi. Permian (Early Guadalupian). • Taiyuanella Zhuang, 1989 (Type species: Taiyuanella subsphaerica Zhuang, 1989). The test is fusiform, with spirotheca composed of a tectum and stalactotheca. Septa are strongly fluted throughout the length of the test. Chomata are weakly developed. Middle Permian (Wordian) (Fig. 2.13). • Triticites Girty, 1904 (Type species: Miliolites secalicus Say, in James, 1823). Septa are weakly folded over the equator, with massive chomata (Fig. 9). Late Carboniferous Chomata Tectum keriotheca Septa Proloculus Fig. 2.16. Morphological features of an equatorial section of Triticites. The Palaeozoic Larger Benthic Foraminifera 89 to Early Permian (Kasimovian to Sakmarian) (Figs 2.14–2 .16; Plate 2.8, Fig.  10; Plate 2.15, figs 1- 3, 7; Plate 2.20, Fig. 6; Plate 2.25, figs 5- 6; Plate 2.11, figs 1- 2). • Zellia Kahler and Kahler, 1937 (Type species: Pseudoschwagerina (Zellia) heritschi Kahler and Kahler, 1937). The test is globular with a thick wall composed of a tec- tum and coarsely alveolar keriotheca, with low, widely spread chomata. Permian (Sakmarian) (Plate 2.25, Fig. 2). Superfamily NEOSCHWAGERINOIDEA Solovieva, 1978 Found in the Late Permian showing the last stage of the keriotheca phase, these Neoschwagerinoids are different from the fusulinoids in having a wall made of a tec- tum and a glossy layer with alveoles running through it (the keriotheca) (Fig. 2.10E), and they differ from the schwagerinoids in having secondary layers and multiple cho- mata or parachomata, with multiple tunnels. Permian (Asselian to Wordian). Family Verbeekinidae Staff and Wedekind, 1910 The test is globose with discontinuous parachomata. Permian (Asselian to Early Lopingian). • Verbeekina Staff, 1909 (Type species: Fusulina verbeeki Geinitz, in Geinitz and von der Marck, 1876). Spherical with a tiny proloculus, parachomata are absent or rudimen- tary (Fig. 2.9). Permian (Roadian to Capitanian) (Figs 2.9, 2.14; Plate 2.15, Fig. 9). • Brevaxina Schenck and Thompson, 1940 (Type species: Doliolina compressa Deprat, 1915). The test is subspherical with a small proloculus, followed by an endothyroid early stage. The wall is thick, with tectum, high and broad parachomata and numer- ous foramina. Permian (Kungurian to early Wordian). Family Misellinidae Miklukho- Maklay, 1958 The test is fusiform with well-d eveloped parachomata. Permian (Sakmarian to Early Lopingian). • Misellina Schenck and Thompson, 1940 (Type species:  Doliolina ovalis Deprat, 1915). The test is ellipsoidal with numerous chambers, and numerous plane, unfluted septa. The keriotheca is thick and the parachomata are “saddle-s haped”. Permian (Kungurian to early Wordian). Family Pseudodoliolinidae Leven, 1963 Ellipsoidal test with planar septa and narrow, high parachomata. The parachomata supplement the primary pair of ridges in the equatorial zone of the shell laterally and polewards, regularly intercalated between supplementary tunnels until the polar end of the chamber Permian (Artinskian to Capitanian). • Pseudodoliolina Yabe and Hanzawa, 1932 (Type species:  Pseudodoliolina ozawai Yabe and Hanzawa, 1932). The test is bluntly circular with thin wall and reduced keriotheca. Septa are planar and unfluted. The parachomata are high, continuous and may reach the top of the chambers adjacent to the septa, appearing as septula in axial section. Permian (Artinskian to Capitanian). 90 Evolution and Geological Significance of Larger Benthic Foraminifera Family Neoschwagerinidae Dunbara and Condra, 1927 In the Neoschwagerinidae (see Fig. 2.11) the finely alveolar keriotheca join the para- chomata to form transverse septula which form complete partitions in the chamber. They occur in the Permian (Asselian to Early Lopingian). • Cancellina Hayden, 1909 (Type species: Fusulina (Neoschwagerina) primigena Hayden, 1909). The test is fusiform and broadly inflated in the centre. The septa are arcuate and broad transverse septula are formed by the extension of the lower part of the kerio- theca. Broad parachomata are present. Early to Late Permian (Roadian to Capitanian). • Neoschwagerina Yabe, 1903 (Type species: Schwagerina craticulifera Schwager, 1883). The test is subspherical. Chambers are divided into chamberlets by transverse septula of first order which connect with the poorly developed parachomata. The wall con- sists of a tectum and a very thick keriotheca that may be differentiated into an upper finely alveolar layer and a lower layer with fewer and coarser alveoli and with upper and lower tectoria. Early to Middle Permian (Wordian to Capitanian) (Plate 2.20, Fig. 1; Plate 2.23, figs 1- 2; Plate 2.26, figs 5- 7). • Sumatrina Volz, 1904 (Type species: Sumatrina annae Volz, 1904). The test is elon- gate fusiform with a large proloculus and long, thin septa. Up to four short sec- ondary septula develop between primary septula. The wall is very thin with a thin keriotheca. Massive parachomata and numerous foramina occur throughout the length of the test. Middle Permian (Capitanian). • Yabeina Deprat, 1914 (Type species:  Neoschwagerina (Yabeina) inouyei Deprat, 1914). The test is large and fusiform with well-d eveloped parachomata connected to the early ends of the septula. Up to three secondary spiral septula in outer whorls and nine axial septula per chamber. Middle Permian (Guadalupian) (Fig.  2.14; Plate 2.24, figs 1- 2). ORDER LAGENIDA DELAGE AND HÉROUARD, 1896 As an order the Lagenida are characterised by having walls made of orientated calcite crystals. It appears that this characteristically evolved independently, in parallel several times in the Palaeozoic, and that the superfamilies Robuloidoidea and Nodosinelloidea have different phylogenetic roots (see Fig. 2.1). Members of this order (Late Silurian to Holocene) were referred to the Nodosariida Calkins 1926 by Vachard et  al. (2010) and Mikhalevich (2013). They have alveo- les. Primitive taxa are without secondary lamination, more advanced forms have a secondary lamination. The wall structures in selected lagenides have been restudied recently (Groves et al., 2004). “Monolamellar” refers to the primary single- layered septal wall. In many lagenides, secondary lamellar develop, where extensions of the primary wall of a given chamber overlap some of the previous chambers. Reiss (1963) proposed an extensive reclassification of lamellar forms, and was among the first to note that radial walls are secreted on an organic substrate. Grønland and Hansen (1976) by examining Holocene lagenides developed a new terminology for lamellarity: • “Atelo-m onolamellar” refers to forms that lack secondary lamellarity and they rep- resent the earliest lagenides, e.g. the simple Nodosaria (Plate 2.5, Fig. 2.11); The Palaeozoic Larger Benthic Foraminifera 91 • “plesio- monolamellar” refers to forms in which secondary lamellarity envelops some but not all previous chambers; • “ortho-m onolamellar” refers to forms in which the primary wall of each chamber secondarily envelops all previous chambers; and • “poly- monolamellar” refers to forms in which yet another lamella secondarily covers the entire test. The plesio- monolamellar forms are thought to have originated in Early Jurassic, fol- lowed by ortho- monolamellar and poly- monolamellar types in the Cretaceous and Cenozoic (Groves et al., 2004). Superfamily ROBULOIDOIDEA Reiss, 1963 The tests of members of this superfamily are those of a typical Lagenida, but without secondarily lamellar or with slight lamination in younger taxa. The aperture is primi- tive and cylindrical. Late Silurian to Early Cretaceous. Family Syzraniidae Vachard, 1981 The test has two chambers and is elongate, composed of a proloculus and an undivided or subseptate tubular chamber partitioned to varying degrees by internal thickenings of the wall, which is calcareous and may have two layers. Late Silurian to Late Permian (Pridolian to Lochkovian; Tournaisian to Changhsingian). • Amphoratheka Mamet and Pinard, 1992 (Type species: Amphoratheka iniqua Mamet and Pinard, 1992). The septa are slightly more pronounced than in Tezaquina. Late Carboniferous to Early Permian (Gzhelian to Sakmarian). • Rectostipulina Jenny-D eshusses, 1985 (Type species: Rectostipulina quadrata Jenny- Deshusses, 1985). The test consists of a long narrow tapering tube with a polygonal transverse section. Septa are short and nearly perpendicular to the test wall. The wall is atelo- monolamellar. The aperture is a simple opening in the center of the apertural face. Late Permian (Wuchiapingian). • Syzrania Reytlinger, 1950 (Type species: Syzrania bella Reytlinger, 1950). The test has a spherical proloculus followed by a non-s eptate tubular chamber. The wall has a “fibrous” layer and an inner microgranular layer. The aper- ture is a simple terminal opening. This genus may have evolved from another simple robuloidoid or most likely from the earlandiides Earlandia in Middle Pennsylvanian. Carboniferous to Late Triassic (Moscovian to Rhaetian) (Fig. 3.3). • Tezaquina Vachard, 1980 (Type species: Tezaquina clivuli Vachard, 1980). Groves in Groves and Boardman (1999) emended Tezaquina to include nonseptate syzra- niid forms whose tubular second chamber is subdivided into pseudo- chambers by weak internal thickenings of the wall, which is thick hyaline fibrous and finely per- forate. The test consists of a spherical proloculus followed by a second chamber that is partitioned into elongate, subcylindrical pseudo-c hambers by weak internal thickenings of the wall. The aperture is a simple terminal opening, and the wall is atelo- monolamellar. Carboniferous to Late Triassic (late Moscovian to Rhaetian) (Fig. 3.3). 92 Evolution and Geological Significance of Larger Benthic Foraminifera • Tuborecta Pronina, 1980 (Type species: Tuborecta vagranica Pronina, in Petrova and Pronina, 1980). The test is small, forming a long narrow tube. The wall is calcare- ous, of two layers, a thin inner finely granular, dark layer and a thick radial, light outer layer. The aperture is terminal, at the end of the tube. Late Silurian to Early Devonian (Pridolian to Lochkovian). Family Protonodosariidae Mamet and Pinard 1992 This family includes the uniserial Lagenida with atelo-m onolamellar or plesio- monolamellar wall structures consisting of radial-fi brous forms. Late Carboniferous to Late Permian (Kasimovian to Changhsingian). • Protonodosaria Gerke, 1959 (Type species:  Nodosaria proceraformis Gerke, 1952). The test is elongate, circular in thin section. Sutures are horizontal and straight. The aperture is round. Late Carboniferous to Late Permian to Late Triassic (Kasimovian to Rhaetian). Family Ichtyolariidae Loeblich and Tappan, 1986 The test is elongate uniserial with a single layered wall, and may show some second- ary lamination, with a simple terminal aperture. Early Permian to Early Cretaceous (Artinskian to Albian). • Pseudotristix Miklukho- Maklay, 1960 (Type species: Tristix (Pseudotrislix) tcher- dvnzevi Miklukho- Maklay, 1960). The test consists of a spherical proloculus fol- lowed by a uniserial arrangement of triangular chambers. The sutures are nearly horizontal, the aperture is a simple opening, and the wall is atelo- monolamellar. Late Permian (Lopingian). Family Robuloididae Reiss, 1963 The test is uniserial, enrolled with an atelo-m onolamellar wall and a terminal aperture. Middle Permian to Late Jurassic (Guadalupian to Kimmeridgian). • Calvezina Sellier de Civrieux and Dessauvagie, 1965 (Type species: Calvezina otto- man Sellier de Civrieux and Dessauvagie, 1965). The test consists of a proloculus followed by five or more uniserial chambers that are weakly trochospirally enrolled in the early stage, later becoming rectilinear. The chambers are slightly compressed laterally and their size increases rapidly from early hemispherical ones to later lobate or irregular ones. The aperture is a simple terminal opening accompanied by a slight thickening of the wall. Late Permian (Lopingian). • Eocristellaria Miklukho- Maklay, 1954 (Type species:  Eocristellaria permica Miklukho-M aklay, 1954). The test is planispirally enrolled with chambers increasing rapidly in breadth, but slowly in height, so the apertural face extend back towards the proloculus. The calcareous wall is composed of an outer hyaline radiate layer and an inner dark microgrannular layer. Late Permian (Lopingian). • Robuloides Reichel, 1946 (Type species:  Robuloides lens Reichel, 1946). The test is planispiral with oblique septa, arching back towards the periphery. Permian to Middle Triassic (Artinskian to Ladinian). The Palaeozoic Larger Benthic Foraminifera 93 Superfamily NODOSINELLOIDEA Rhumbler, 1895 The Nodosinelloidea evolved from the Earlandiida which in turn evolved from the Parathuramminoidea (see Fig. 2.2). Their tests were free, with more than one distinct chamber, partially septate or fully septate, with a simple microgranular wall or evolving into two layers with an extra fibrous inner layer. The Nodosinelloidea range from the Late Silurian to the Permian. Family Earlandinitidae Loeblich and Tappan, 1988 The Earlandinitidae (Fig. 2.2) evolved from an earlandioid, Caligellidae, by becom- ing elongate, uniserial, semi-s eptate, with a simple microgranular wall, and range from Late Devonian to Late Carboniferous (Bashkirian). • Darjella Malakhova, 1964 (Type species: Darjella monilis Malakhova, 1964). The test is uniserial, with globular chambers and strongly constricted sutures. Early Carboniferous (Early Visean). • Earlandinita Cummings, 1955 (Type species:  Nodosinella perelegans Plummer, 1930). The test is subseptate with a terminal round aperture. Carboniferous (See Fig. 2.2). • Lugtonia Cummings, 1955 (Type species: Nodosinella concinna Brady, 1876. The test is rectilinear with depressed chambers overhanging the sutures. Carboniferous. Family Nodosinellidae Rhumbler, 1985 This family (Fig. 2.2) evolved from the Earlandioidea, and is characterised by having a uniserial test which develops in the Late Carboniferous to having a double layered wall, with an inner fibrous layer and an outer microgranular layer. They range from the Late Silurian to Permian. • Biparietata Zolotova, 1980 (Type species: Biparietata ampula Zolotova, 1980). The test is rectilinear with elongate chambers and sutures not apparent externally. Middle Permian (Guadalupian). • Nodosinella Brady, 1876 (Type species: Nodosinella digitata Brady, 1876). The test is free and fully septate. Permian (see Figs 2.2, 2.4). Superfamily COLANIELLOIDEA Fursenko, 1959 In this superfamily, the tests are uniserial and fully septate, and chambers are strongly overlapping with partitions. The walls occur with an outer vitreous layer and an inner granular layer. They range over the Late Devonian to Late Permian. Family Colaniellidae Fursenko, 1959 The family evolved from the Nodosinellidae by developing strong internal partitions or septa (Fig. 2.2). They range from the Late Devonian to Late Permian. • Colaniella Likharev, 1939 (Type species: Pyramis parva Colani, 1924). Elongate and uniserial, having a calcareous wall with a fibrous structure with strongly overlapping sutures. The aperture is terminal radiate. Late Permian (see Fig. 2.2). 94 Evolution and Geological Significance of Larger Benthic Foraminifera • Cylindrocolaniella Loeblich and Tappan, 1985 (Type species:  Wanganella ussuriensis Sosnina, in Kiparisova et  al., 1956). The test is slightly arcuate, narrow and cylindrical in form with numerous short septa. Late Permian (Lopingian). • Pseudowanganella Sosnina, 1983 (Type species:  Pseudowanganella tenuitheca Sosnina, 1983). The test is narrow, flattened and elliptical in cross section. Chambers are curved centrally and overlapping near the periphery. Septa are more thickened near the aperture. Late Permian (Lopingian). Superfamily GEINITZINOIDEA Bozorgnia, 1973 The tests of this superfamily are uniserial and appear similar to the Nodosinelloidea, but the microgranular layer is the inner dark layer while the radially fibrous layer is an outer layer. Advanced forms have secondary lamellarity. Late Devonian to Middle Triassic. Family Geinitzinidae Bozorgnia, 1973 The tests are free and uniserial. They evolve from species that are rounded in cross- section to species that are compressed with a terminal aperture. The range is Late Devonian to Late Permian. • Geinitzina Spandel, 1901 (Type species:  Geinitzella (Lunucammina) permiana Spandel, 1898). The test is uniserial elongate, laterally compressed, exhibiting in transverse view two planes of bilateral symmetry and an oval aperture in the centre of the apertural face. Permian (Cisuralian to Lopingian) (Figs 3.3 and 3.4). • Howchinella Palmieri, 1985 (Type species: Frondicularia woodwardi How-c hin, 1895). The test is laterally compressed, with arched or chevron sutures. The wall is calcare- ous granular with a dark thin inner organic layer and an outer hyaline layer of opti- cally radial calcite. The aperture is radiate with a tooth. Early Permian (Sakmarian). • Nodosinelloides Mamet and Pinard, 1992 (Type species: Nodosinelloides potievskayae Mamet and Pinard, 1996 (for Nodosaria gracilis Potievskaya, 1962 preoccupied)). The test consists of a spherical proloculus followed by as many as 10 or more unise- rial circular chambers with pronounced septal inflection and a simple aperture. The wall lacks secondary lamellarity (atelo-m onolamellar) and the aperture is a simple terminal opening. Carboniferous to Triassic (Kasimovian to Rhaetian) (Fig. 3.3). Family Pachyphloiidae Loeblich and Tappan, 1984 This family has a free uniserial compressed test with a simple calcareous microgranular wall, but with secondary thickenings on both sides of the test. The range is from Early to Late Permian. • Pachyphloia Lange, 1925 (Type species:  Pachyphloia ovata Lange, 1925). The test consists of a globular proloculus followed by a uniserial succession of laterally com- pressed, strongly overlapping chambers. The wall is plesio- or ortho-m onolamellar with thin dark inner layer and thicker hyaline outer layer. The aperture is ter- minal, oval, radiate (toothed). Early Permian to Late Permian (Sakmarian to Changhsingian) (Plate 2.5, figs 2- 4; Fig. 3.3). The Palaeozoic Larger Benthic Foraminifera 95 ORDER MILIOLIDA DELAGE AND HÉROUARD, 1896 The miliolides have tests that are porcellaneous and imperforate, which are made of high Mg- calcite with fine randomly oriented crystals. They range from the Carboniferous to the Holocene. Superfamily CORNUSPIROIDEA Schultze, 1854 The test is free or attached, and composed of a globular proloculus followed by a tubular enrolled chamber. The coiling is planispiral or trochospiral, evolute or invo- lute and may become irregular. There is a simple aperture at end if the tube. Early Carboniferous to Holocene. Family Cornuspiridae Schultze, 1854 The tests are free or attached, composed of a proloculus followed by an undivided planispiral to streptospiral, involute or evolute second tubular chamber. Early Carboniferous (Visean) to Holocene. • Cornuspira Schultze, 1854 (Type species: Orbis foliaceus Philippi, 1844). The test is discoidal with a globular proloculus and undivided planispirally enrolled and tubu- lar second chamber. Carboniferous (Moscovian) to Holocene. • Rectocornuspira Warthin, 1930 (Type species: Rectocornuspira lituiformis Warthin, 1930). Test elongate, with a globular proloculus and planispirally enrolled to later uncoiling and rectilinear undivided tubular second chamber. Late Carboniferous to Early Triassic (Moscovian to Induan). Family Hemigordiidae Reitlinger, 1993 Characterized by a test with at least an early trochospiral stage, but that may later become streptospiral. Carboniferous. • Hemigordius Schubert, 1908 (Type species: Cornuspira schlumbergeri Howchin, 1895). The test has early whorls that are streptospiral, later becoming planispiral and evo- lute. Early Carboniferous to Late Carboniferous (Late Visean to Gzhelian) (Fig. 3.5). • Neohemigordius Wang, 1973 (Type species: Neohemigordius maopingensis Wang and Sun, 1973). The test is lenticular, with a lamellar thickening over the umbilical area. Early Permian (Asselian to Sakmarian) (Fig. 3.5). • Pseudovidalina Sosnina, 1978 (Type species: Pseudovidalina Sosnina, 1978). The test is discoid with a globular proloculus followed by planispirally, evolute, enrolled tubu- lar undivided second chamber with secondary thickenings in the umbilical region. Middle Permian (Kungurian). Family Hemigordiopsidae Nikitina, 1969 Forms in this family have a test with at least an early streptospiral stage, that later may be planispiral. Early Carboniferous (Visean) to Holocene. • Hemigordiopsis Reichel, 1945 (Type species: Hemigordiopsis renzi Reichel, 1945). The test is globular with a proloculus followed by an undivided tube, which is strongly involute. Early to Middle Permian (Guadalupian) (Fig. 3.3). 96 Evolution and Geological Significance of Larger Benthic Foraminifera • Shanita Brönnimann, Whittaker and Zaninetti, 1978 (Type species: Shanita amosi Brönnimann, Whittaker and Zaninetti, 1978). The proloculus followed by an undi- vided enrolled second chamber, with a streptospiral in early coiling, becoming pla- nispiral and involute. Later whorls with alternating vertical pillars. Late Permian (Wuchiapingian to Changhsingian) (Fig. 3.6). 2.3 Biostratigraphy and Phylogenetic Evolution 2.3.1 The fusulinides and related orders The organic- walled allogromiides gave rise to the most primitive Palaeozoic larger benthic foraminifera, the Parathuramminida, in the Silurian (Figs 2.1, 2.2). The simple form parathuramminides gave rise rapidly to the pseudoseptate Pseudoammodiscidae belonging to the earlandiides in the Devonian, and eventually to the pseudoseptate tournayellides in the Late Devonian (Fig. 2.6). The latter gave rise to the completely septate endothyrides. The endothyrides include coiled forms and advanced uniserial genera with uniserial adults (e.g. Haplophragmella). They also exhibit species with a strong development of secondary deposits of calcite (chomata) on the chamber floor (e.g. Endothyra, Quasiendothyra). The chomata enfold the wall to give septal chamberlets in the Bradyinidae, which communi- cate with the exterior via septal pits (e.g. Bradyina, Figs 2.9, 2.10). The endothy- rides reached their acme and largest sizes in the Early Carboniferous (Visean), with trends towards evolute planispiral coiling and well developed septa in early whorls in the Loeblichiidae (e.g. Loeblichia). The latter finally gave rise to the higher Fusilinida (see Fig 2.17). The Loeblichiidae developed walls that were differentiated into two or more layers in the most advanced forms, leading to the first appearance of the Ozawainelloidea lineage of the fusulinides in the Early Carboniferous (Visean) (Còzar and Vachard, 2001). The fusulinides started as simple tiny organisms in the Early Carboniferous, but had evolved into large and complicated forms by the end of the Carboniferous and Permian. Their very large sizes in the Early Permian, some up to 8cm length, are related to the microspheric generations, as dimorphism becomes evident in larger benthic fora- minifera. The rapid proliferation of genera, the increasingly strongly folded septa, the insertion of septula, the development of chomata and axial deposits, the replacement of the central tunnel by numerous foramina, the development of axial infilling in the very large species, and the development of parachomata in the most advanced forms, led to the evolution of the different superfamilies of the fusulinides, the Staffelloidea, the Schubertelloidea, the Fusulinoidea and the more advanced Schwagerinoidea, and Neoschwagerinoidea (Fig.  2.18). Because of their complexity, size and stratigraphi- cal usefulness, the evolution and range of the Fusulinida superfamilies deserve more detailed discussion, and are the focus of the remainder of this section. The Fusulinida and related orders are the only significant forms of foraminifera without living representatives (Groves et  al., 2005). The evolutionary relationships between many members of the Fusulinida and other Palaeozoic groups are summa- rized in Figs 2.1 and 2.2, while their biostratigraphic ranges are shown later in Figs. 2.17 to 2.19. The Palaeozoic Larger Benthic Foraminifera 97 Period, Epoch and Stage 251.9 298.9 323.2 346.7 Fig.  2.17. Evolution the fusulinides. Abbreviations:  A.  =  Asselian, S.  =  Sakmarian, Ar.  =  Artinskian, K. = Kungurian, R. = Roadian, W. = Wordian, C. = Capitanian, W. = Wuchiapingian, Ch. = Changhsingian. Age (Ma) Carboniferous Permian Upper Mississippian Pennsylvanian Cisuralian Guadalupian Lopingian Visean Serpukhovian Bashkirian Moscovian Kasimovian Gzhelian A. S. Ar. K. R. W. C. W. Ch. Loeblichia Loeblichiidae Ozawainellidae Millerella Profusulinella Fusulinella Fusulinidae Fusulina Schubertella Schubertellidae Triticites Pseudoschwagerina Schwagerinidae Parafusulina Polydiexodina Sphaerulina Staffellidae Staffella Verbeekina Verbeekinidae Yabeina Neoschwagerinidae 98 Evolution and Geological Significance of Larger Benthic Foraminifera Period, Epoch and Stage Palaeofusulina Codonofusiella Yabeina Neoschwagerina Cancellina Misellina Schwagerina Ticites Fusulina Profusulinella Pseudostaffella Eostaffella Fig. 2.18. Evolution and biozonations of the fusulinides (drawings of foraminifera from Jin- Zhang (1990) and are not here to scale). One of the most important morphological features of the fusulinides and the related orders is their mode of coiling. Planispiral coiling is dominant (70%; Haynes, 1981) espe- cially in the Early Carboniferous. On the other hand, a fusiform chamber arrangement is common in the most advanced forms of fusulinides from Late Carboniferous to Late Permian. Many of the structural features of the large fusulinides tests can be interpreted from external observation. The traces of the septa on the external surface are called septal furrows. The apertural face of the last chamber is the antetheca and the external wall is the spirotheca (see Figs 2.15 and 2.16). The antetheca lacks true Age (Ma) Carboniferous Permian Upper Mississippian Pennsylvanian Cisuralian Guadalupian Lopingian A. S. Ar. K. R. W. C. W.Ch. Fusulinida Zones Loeblichiidae Ozawainellidae Staffellidae Fusulinidae Schubertellidae Schwagerinidae Neoschwagerinidae Verbeekinidae The Palaeozoic Larger Benthic Foraminifera 99 apertures, but the internal septa show single foramen or multiple foramina that are formed by resorption of the wall (Haynes, 1981; Leppig et al., 2005). In the Fusulinidae (except for few genera such as Paraeofusulina) and the Schwagerinidae (except for few genera such as Polydiexodina and Eopolydiexodina) a single large foramen is formed, while in the neoschwagerinids Verbeekinidae and Neoschwagerinidae multiple foramina are characteristic. However, like all other larger foraminifera, they cannot be identified solely on external features, and thin sections are essential for the identification of fusu- linids, neoschwagerinids and staffellids. Two sections cut through the test reveal all the complicated structures needed for classification (see Figs 1.7, 2.15, 2.16). One of these is the cut through the axis of coiling (the axial section), and the other is the cut normal to the axis of coiling (the equatorial or sagittal section). One of the diagnostic features to be seen in either section is the morphological development shown by the wall structures of the Late Carboniferous and Permian fusulinides. These wall structures range from simple calcareous granulated, in the prim- itive fusulinides, to compound microgranular, and then very complicated in the large fusulinids and schwagerinids of the Permian. As described above, the fusulinides can exhibit one of five types of wall structure (see Figs 2.10), which can be recognised as major evolutionary events (see Fig. 2.13): • a Pre-k eriotheca Phase, with a thin spirotheca, composed of a tectum only, the tec- tum stage, such as in the fusulinid Profusulinella and an early thicker, less dense layer, the diaphanotheca. This diaphanotheca stage includes forms such as Fusulinella, Ozawainella and Protriticites (see Fig. 2.10, A-D ; Fig. 2.13). This phase corresponds to the Middle Carboniferous. • a second phase (the Keriotheca Phase), in which the spirotheca is composed of a tec- tum and a coarsely alveolar keriotheca (Fig. 2.10E), is divided according to Vachard et  al (2004) into three stages, the keriotheca s.s., the anthotheca and the stalacto- theca (see Fig. 2.13). These stages evolve respectively from the Late Carboniferous to the late Middle Permian. During the first two stages in the Late Carboniferous and Early Permian, the family Schwagerinidae dominated. On the chamber floors of these advanced fusulinides secondary deposits, or chomata, were laid down and are related to the form of the chamber communications (see Fig. 2.13, 2.15). During the stalactotheca stage, in the Middle Permian, the Verbeekinidae and Neoschwagerinidae were the dominant fusulinides. They had multiple foramina and multiple small cho- mata, thus developing parachomata (see Figs 2.9, 2.14). In the Neoschwagerinidae, the parachomata became linked with the primary spiral septula to form complete partitions. • a Post- keriotheca Phase, exemplified by Nanlingella and Palaeofusulina, which became abundant after most of the Schwagerina- Verbeekina- Neoschwagerina types had disappeared (see Fig. 2.19). The oldest and most primitive keriothecal wall structure is seen in Triticites (Hageman and Kaesler, 1998). The wall composition is microgranular and the alveo- les are filled by a structureless microspar (Vachard et al., 2004). In other forms, such as Sakmarella (Pseudofusulina- like form), Praeskinerella (Parafusulina- like form), and Taiyuanella (Chusenella- like), the alveoles are occupied by radial calcite prisms, which 100 Evolution and Geological Significance of Larger Benthic Foraminifera Period, Epoch and Stage 251.9 259.1 298.9 323.2 358.9 Age (Ma) Carboniferous Permian Mississippian Pennsylvanian Cisuralian Guadalupian Lopingian Tournaisian Visean Serpukhovian Bashkirian Moscovian Kasimovian Gzhelian A. S. Ar. K. R. W. C. W. Ch. Archaesphaeridae Parathuramminidae Chrysothuramminidae Ivanovellidae Parathuramminoidea Uralinellidae Usloniidae Tuberitinidae Earlandiidae Pseudoammodiscidae Earlandioidea Pseudolituotubidae Archaediscidae Archaediscoidea Lasiodiscidae Lasiodiscoidea Caligellidae Caligelloidea Paratikhinellidae Earlandinitidae Nodosinellidae Nodosinelloidea Geinitzinidae Geinitzinoidea Pachyphloiidae Colaniellidae Colanielloidea Ptychocladiidae Ptychocladioidea Semitextulariidae Palaeotextulariidae Palaeotextularioidea Biseriamminidae Tournayellidae Tournayelloidea Palaeospiroplectamminidae Endothyridae Endothyroidea Bradyinidae Tetrataxidae Pseudotaxidae Tetrataxoidea Valvulinellidae Abadehellidae Fig. 2.19. Ranges of the Parathuramminoidea and Robuloidoidea in the Carboniferous and Permian. The Palaeozoic Larger Benthic Foraminifera 101 are arranged like narrow- centred petals (see Fig. 2.13), called anthotheca (Vachard et al., 2004). This structure is thought to represent the calcification of some bacteria or cyanobacteria (Riding, 1977; Feldmann and McKenzie, 1998). Finally, another texture appearing like large- centred petals is known as the stalactotheca (Zhuang, 1989). In this structure, the alveoles are filled by a blocky granular calcite, but radiat- ing prisms of calcite also exist at the periphery (Fig. 2.13). Some more advanced forms have well developed perforations (Haynes, 1981). Algal symbionts, common in mod- ern larger foraminifera, had evolved by the Carboniferous (Tappan, 1971). Larger perforations were developed in taxa with an anthotheca, so cyanobacteria could cal- cify within the pores of the wall. The diaphanotheca (internal clear layer in the wall) was probably partially filled in the living organism by cyanobacterial endosymbionts and microgranules that were possibly aragonitic (Vachard et  al., 2004). This hypo- thesis of a mixture of organic and mineral components has also been suggested by Bender and Hemleben (1988). Vachard et  al. (2004) also suggested a phylogenetic lineage from the keriotheca to anthotheca, and stalactotheca stages (see Fig. 2.13). In the Neoschwagerinidae, the alveolar keriotheca seem to appear independently from those of the fusulinids. The inner structures, the arrangement of internal foramina, and the straight septa which divide the chambers completely into chamberlets seen in the neoschwagerinids and verbeekinids are morphologically similar to those seen in the later alveolinids (Mikhalevich, 2004; 2009), and provide an example of parallel evolution over time. In advanced forms of the Fusulinoidea the septa become strongly folded. The fold- ing first appeared near the poles, then migrated up to the centre of the test near the sutures and subsequently fill the test. In the Schwagerinoidea the septa became very strongly folded and produced tunnel-l ike cuniculi (see Fig. 2.10) running in the direc- tion of the coiling (e.g. in the Carboniferous Paraeofusulina, and also in the Permian Parafusulina) (see Plate 2.16, figs 1- 2). Due to the problem of recrystallisation, tracing possible evolutionary relation- ships between the different genera within the fusulinides has been difficult, because detailed ultra- structure is over printed by diagenetic processes. Their evolution, however, has been discussed by many authors (Haynes, 1981; Rauzer- Chernosova, 1963; Loeblich and Tappan, 1980). Jin- Zjang (1990) considered the small lenticular endothyrides as the ancestral form to the Chinese fusulinids. They evolved into the Ozawainelloidea in the Late Mississippian, which are the earliest primitive ances- tors of the superfamily Fusulinoidea. The Ozawainellidae family are characterised by small discoidal shells with unfluted septa. The most common genera of this family are minute, spherical forms belonging to Ozawainella and Millerella (Fig. 2.11), and are mostly common in the mid-C arboniferous (Serpukhovian and Bashkirian). The wall structure of these genera is of the Ozawainella-t ype. The number of the genera of the ozawainellids decreased in the Late Carboniferous, but the family persisted to the Late Permian. The Fusulinidae family (as well as the less significant Staffellidae family) arose from the Ozawainellidae in the Middle Carboniferous (Moscovian), where there was a rapid increase in the number of genera. The septa became increasingly folded and the fam- ily persisted into the Middle Permian, but with a marked decrease in the number of genera. Most of the genera of the Middle Carboniferous had a wall structure of the 102 Evolution and Geological Significance of Larger Benthic Foraminifera Profusulinella-t ype, including genera with a wall composed of a tectum, late and early tectorium, or the Fusulina- type, where the spirotheca is composed of a tectum, diapha- notheca and thick late and early tectoria. This family persists into the Middle Permian but with a decreasing diversity of genera (Fig. 2.12). The Schubertellidae evolved from the Ozawainellidae in the middle Bashkirian (Sinitsyna and Sinitsyn 1987; Nikolaev 2005). The earliest primitive schubertel- lid, Grovesella, may have evolved from the ozawainellid Eostaffella by developing loosely coiled whorls, a two−layered wall (as opposed to the undifferentiated wall in Eostaffella), planispiral coiling and an absence of chomata or pseudochomata that are always present in Eostaffella (Davydov, 2011). The affinity between Grovesella and ozawainellids or staffellids or schubertellids is still disputable. Grovesella has been considered by Leven (2009) as belonging to the Zarodella Sosnina lineage of the ozawainellids, however Davydov (2011) demonstrated that this taxon belongs to the schubertellids. Grovesella evolved rapidly into the larger Schubertina in the late Bashkirian (Sinitsyna and Sinitsyn 1987; Nikolaev 2005). Schubertina, with its wall differentiated into two layers, survived up to the Wordian (Davydov, 2011). The fusi- form Schubertella first appeared in the Moscovian (Rauser−Chernousova et al. 1951). Schubertella, with a wall differentiated into three layers, lived from the Moscovian through to the Lopingian with several acme zones in the Moscovian– Kasimovian, late Asselian– early Sakmarian and late Artinskian times (Davydov, 2011). In the early Gzhelian, Schubertella evolved into the relatively large schubertellid Biwaella, which survived through Artinskian– Kungurian time. In the latest Gzhelian, the latter devel- oped fluted septa in Dutkevitchites, which in turn developed into a highly- specialized form in the Permian, Sphaeroschwagerina (Davydov 1984). Another advanced schuber- tellid, Mesoschubertella, is documented throughout the Permian, but Davydov (2012) speculated that its origination could have been in the Sakmarian–A sselian or even in the late Gzhelian. The Schwagerinidae arose from Fusulina in the Late Carboniferous (Fig.  2.17). The Schwagerinidae of the Late Carboniferous are dominated by Triticites, a mod- erately folded form with strong chomata and a spirotheca with an alveola keriotheca (Figs 2.15 and 2.16). There was a progressive decrease in the number of wall layers, accompanied with a progressive increase in the opacity of the diaphanotheca, from Fusulinella to Prototriticites, Montiparus and Triticites in the Late Carboniferous (van Ginkel and Villa, 1999) (see Fig. 2.13). Most of the Late Carboniferous genera became extinct at the top of the Gzhelian, except members of the Schubertellidae, which did not become extinct until the end of the Permian. Schubertella itself is a Permian form (see Fig. 2.17). In the Upper Carboniferous strata, the marker species of the middle Kasimovian is Montiparus montiparus (Fig. 2.13), and is widely distributed in the East European Basin and Tethyan province. In the Gzhelian, Montiparus evolves into Carbonoschwagerina in the peri-G ondwana realm (Ozawa et al., 1990). The precise position of the Carboniferous- Permian boundary has long been the subject of debate (Leven and Gorgij, 2006; Ozawa et al., 1990). However, many studies agree that the Carboniferous-P ermian boundary in the East European Basin is defined on the first appearance of Pseudoschwagerina (see Fig. 2.19). The latter evolved from Triticites at the base of the Permian. In the Tethyan The Palaeozoic Larger Benthic Foraminifera 103 realm and the peri- Gondwana part of Tethys the base of the Permian coincides with the appearance of Sphaeroschwagerina fusiformis (Ozawa et al., 1992). Schwagerinids evolved slowly during the Late Carboniferous. However, during the Asselian and Sakmarian part of the Early Permian a dramatic evolution of the Schwagerinidae produced 50 genera (Leven, 2003). These forms became intensely folded (Fig.  2.14) as well as becoming increasingly abundant and diverse. While Triticites decreased in abundance and importance (until it disappeared completely at the top of the Early Permian) about half of the schwagerinids became extinct before the Kungurian. From Early Permian to Middle Permian, Schwagerinids exhibited evolutionary lineages with chamber inflation, such as in genera Pseudoschwagerina, Sphaeroschwagerina and Zellia. Sphaeroschwagerina is considered to have evolved in the early Asselian from a species of Dutkevitchites, a member of the Schubertellidae (Ozawa et al., 1992). Cuniculi also appeared in two separate lineages. The earliest lineage contains the North American Eoparafusulina in the Early Permian and the cosmopolitan Monodiexodina in the Early to Middle Permian and, appearing slightly later, the sec- ond lineage contains the cosmopolitan Parafusulina (Guadalupian- Lopingian), and the North American Skinnerina and Polydiexodina (late Kungurian to late Guadalupian). The wall structure of the Schwagerinidae became more complicated and of the Schwagerina type. Most of these forms became extinct in the Middle Permian. The Schwagerinidae evolved into even more complicated forms in the Late Permian, such as the Polydiexodina and its related genera (Fig. 2.14). These forms are characterised by a single- layer wall but a very large test with intensely folded septa. A rapid evolution of the Neoschwagerinoidea followed the first appearance of the Misellinidae in the Kungurian; Misellina evolved directly from the ozawainel- lid Pamirina in the Kungurian by developing “saddle-s haped” parachomata (Leven, 2009). The Sumatrina group of the Neoschwagerinidae sprung directly from Misellina at the base of the Guadalupian. This horizon also marks the first occurrence of the Verbeekinidae, which formed one of the most distinctive lineages in the Tethyan fau- nal realm. The earliest verbeekinid, Brevaxina, which first appeared during the earliest part of the Guadalupian, witnessed a rapid evolution to lineages that culminated with Verbeekina and Yabeina. It descended either indirectly from Staffella via the interme- diate form Sphaerulina (Ross, 1967), or from the ozawainellids Pamirina (Kanmera et al., 1976) and rapidly evolved into a succession of specialised genera (see Fig. 2.17). These fusulinides tests have parachomata and evolved into forms with complete sys- tems of foramina in each chamber (see Fig. 2.14). The parachomata are discontinuous in Verbeekina, but continuous in the neoschwagerinid Yabeina. The end Guadalupian saw the extinction of all large and morphologically complex forms assigned to the Schwagerinidae and Neoschwagerinidae (Leven and Korchagin, 2001), and only 15 genera in the families Schubertellidae and Staffellidae persisted into the Lopingian (Groves and Altiner, 2005). The diversity of schubertellids reached its maximum at the end of the Guadalupian, and it underwent a minor burst of evolu- tionary diversity during the Lopingian (see Fig. 2.20). The new genus, Codonofusiella has irregularly coiled shells with strongly fluted septa, becoming uncoiled at maturity (see Fig.  2.13), while others developed thick fusiform shells having thickened walls, such as Palaeofusulina. However, this late diversification was followed by the extinction 104 Evolution and Geological Significance of Larger Benthic Foraminifera Fig. 2.20. The biostratigraphic range and diversity of the main Palaeozoic superfamilies (as shown by the horizontal scale of the spindles) found in the Carboniferous- Permian. of the remaining schwagerinids, verbeekinids, and neoschwagerinids at the end of the Permian. 2.3.2 Other Palaeozoic Larger Foraminifera The Archaediscida evolved from the parathuramminides in the Carboniferous. They include the bilocular Archaediscoidea (Fig. 2.21) and Lasiodiscoidea (Fig. 2.22A). The archaediscids have a streptospirally enrolled second chamber, while the lasiodiscids are conical with some genera (e.g. Howchinia) developing a high helicoidal spire. The archae- discids remain undivided and differ from the fusulinides in that the outer layer of their wall has a radial hyaline structure. They reach their maximum abundance in the Visean. In addition to the fusulinides described above, the earlandiides gave rise to many other diversified descendants, such as the plurilocular Semitextulariidae of the Palaeotextulariida in the Devonian. The Semitextulariidae include biserial, flattened forms, often becoming monoserial with broad, fully septate chambers. Members of the Palaeospiroplectamminidae represent a biserial development from the Tournayellidae, The Palaeozoic Larger Benthic Foraminifera 105 and may be ancestral to the Palaeotextulariidae. Palaeospiroplectammina have an initial coil and a two- layered wall indicating a possible origin from a planispiral endothyrid (Haynes, 1981). It gave rise to the first Palaeotextulariidae, Palaeotextularia in the Early Carboniferous (Tournaisian), which in turn evolved into Palaeobigenerina and related genera in the Visean. The Palaeotextulariidae (see Figs 2.2 and 2.4) have biserial to uni- serial genera that closely resembles the Textulariidae, but have a dark granular calcar- eous outer wall and an inner clear to yellowish “fibrous” layer with stacks of granules perpendicular to the surface. Cummings (1956) demonstrated that these forms evolved from simple agglutinated forms in the Devonian where, in some cases, a partially aggluti- nated outer layer is still present (e. g. Cribrogenerina). The aperture is basal in the biserial forms, becoming terminal and cribrate in the uniserial part (e.g. Climacammina). Like the archaediscids, the paleotextulariids reach their maximum abundance in the Visean. The Tetrataxida (Fig. 2.5) are highly conical, multilocular forms with trochospiral arrangements and an open central umbilicus at the base of the cone. The wall structure is two-l ayered and very similar to the paleotextulariids from which they are inferred to have evolved. They are divided into the single- layered pseudotaxids and valvulinellids and the two layered tetrataxids. The valvulinellids differ from the pseudotaxids and the tetrataxids in having subdivided chambers with vertical and horizontal partitions. streptospirally enrolled second chamber - coiled in successively changing planes Radially fibrous outer layer Proloculus Dark inner layer Fig. 2.21. Main morphological features of Archaediscus test. Scale bars = 0.5mm. 106 Evolution and Geological Significance of Larger Benthic Foraminifera Umbilical region filled fibrous calcite A Broad umbilical cavity B Vertical partitions C Fig. 2.22. Main morphological features of A) Howchinia bradyana (Howchin); B) Vissariotaxis cummings Hallet; C) Valvulinella youngi (Brady). Scale bars = 0.3mm. The two- layered walls of the tetrataxids suggests a phylogenetic relationship with the endothyrides but this has not been proven yet. The partially to fully spetate nodosinelloids of the Palaeozoic Lagenida evolved from the parathuramminides in the Silurian via the subseptate caligellids of the earlandi- ides. The nodosinelloids include the earlandinitids with the single- layered wall and the nodosinellids with the double- layered wall. The Syzraniidae evolved within the Robuloidoidea in the Upper Silurian of North Urals in Russia with the first appearance of Tuborecta. The latter has a two- layered wall struc- ture resembling the other Syzraniidae, but with a simpler morphology. Tuborecta died out soon after the beginning of the Devonian and it is only in the Late Carboniferous, the Moscovian- Kasimovian interval, that the Syzraniidae reappeared. The evolution- ary radiation of this family, with its oldest known Carboniferous genus Syzrania, might have evolved from the another simple robuloidoids or most likely from the earlandi- ides Earlandia in Middle Pennsylvanian time (see Chapter 3), through the addition of a hyaline- radial layer external to the ancestral microgranular wall (Groves et al., 2004). Palaeozoic lagenides are poorly studied in comparison with most other foraminif- eral groups. They did not show significant evolutionary radiation until the Moscovian (Middle Pennsylvanian), but early lagenides were minor members of foraminiferal faunas for most of the remaining Palaeozoic Era, while fusulinides were spectacularly diverse (Groves et al., 2003). Lagenides became conspicuous only after fusulinides suffered steep declines in both diversity and abundance at the end of the Guadalupian epoch (Leven The Palaeozoic Larger Benthic Foraminifera 107 and Korchagin, 2001) and their almost complete disappearance at the end of the Permian (Groves, 2005). It is noteworthy that the Palaeozoic lagenides survived the end Permian extinction, and gave rise to the Mesozoic- Cenozoic lineages (Groves et  al., 2003; Hallam and Wignall, 1997). Groves (2005), in studying the assemblage of lagenides in lattermost Permian rocks of the central Taurides in Turkey, stated that out of the 22 species in 16 genera, only two identifiable species in the primitive long- ranging “Nodosaria” (such as N. radicula (Linné), see Plate 2.5, Fig. 11) and indeterminate Syzrania (see Chapter 3) survived the end Permian mass extinction. The last occurrences of most taxa fall within the last half- meter of the Permian strata, a pattern consistent with abrupt extinction when tested for the Signor- Lipps effect. Generic diversity within the lagenides exceeded that of all other calcareous foraminiferal groups throughout most of the Mesozoic Era prior to the Late Cretaceous rapid diversification of rotaliines. The suborder includes approximately 120 extant genera and ranks behind only the Miliolida and Rotaliida as the most diverse group of living calcareous foraminifers (Tappan and Loeblich, 1988). They will be discussed further in the following chapter. The miliolides of the Triassic made their first appearance in the Carboniferous and Permian. During this time, they were small foraminifera, and not until the Triassic, after the extinction of most of the larger fusulinides, did they evolve into different larger forms, filling the empty niches left by the extinction of the Permian larger ben- thic foraminifera. The evolutionary relationships within all of these groups are also explored further in Chapter 3. 2.4 Palaeoecology The dominant earlier Palaeozoic foraminifera were mainly characterised by having an undivided tubular chambers with diverse types of coiling. The uncoiled planispiral, biserial to uniserial, or uniserial genera represented the infaunal assemblage of the Carboniferous. They were mainly infaunas living an endobenthic mode of life within the sediment or at the seawater/s ediment interface (Vachard et al., 2010). The unilocu- lar parathuramminides, and the plurilocular semitextulariids of the palaeotextulariides were mainly distributed along the Late Silurian and Devonian reef, restricted to back- reefs (lagoons) and fore-r eef (middle ramps), or off- reef environments (Krebs, 1972; Vachard, 1974, 1994; Préat and Kasimi, 1995), while the moravamminoides typically lived on the Devonian reefs or inner ramps (Vachard et al., 2010). The palaeotextulari- ides (e.g. Palaeotextularia) are believed to have thrive mostly in low energy environ- ments, where turbidity, current and wave action were minimal, with the most favorable conditions being in water deeper than 20m (Stevens, 1971). Also, palaeotextulariides could live in shallower environments (5–1 0m depth), where they probably found shel- tered niches among algal thalli (Gallagher, 1998). In contrast, the distribution of large palaeotextulariides (e.g., Climacammina, Cribrogenerina) shows a preference for wave- agitated environments (Porta et al., 2005). In the Carboniferous, the first trochospirally coiled, conical Tetrataxida (e.g., Tetrataxis, Pseudotaxis and Abadehella) occur in a wider range of environments than most fusulinides. They have a greater tolerance to decreased light. They are abundant 108 Evolution and Geological Significance of Larger Benthic Foraminifera in the inner platform facies and are common in shallow reefal facies (Toomey and Winland, 1973). Attached forms are rare, and forms such as Tetrataxis might have lived attached during a part of its life cycle (Kochansky- Devidé, 1970; Cossey and Mundy, 1990; Gallagher, 1998; Vachard and Krainer, 2001; Pille, 2008; Vachard et al., 2010). The late Tournaisian-V isean Tetrataxis could also tolerate decreased light conditions at water depth of about 200m (Lees, 1997; Gallagher, 1998). Similarly, archaediscoids could thrive in turbulent environments (Haynes, 1965; Brenckle et al., 1987; Gallagher, 1998), and in the relatively lower energy and deeper (below wave base) environments where they frequently occur with lasiodiscoids and pseudoammodiscids (Porta et al., 2005). It was suggested by Gallagher (1998) that the lenticular shape of the archaedis- coid test contributed to the stability in turbulent environment and the calcareous pris- matic wall enabled algae symbiosis. In the late Tournaisian and Visean, the Endothyrida and Tournayellida became diversified and spread on the entire inner ramp (Vachard et al., 2010). They appear to have been mostly endobenthic (Pille, 2008). The earliest forms Eotournayella and Septatournayella were found in deep water limestones (Vachard, 1973, 1974; Gallagher, 1998; Cózar and Rodríguez, 2003; Mohtat-A ghai et al., 2009). Compressed small endothyrids and tournayellids are most abundant in near- shore and shelf lime- stones (Ross, 1973), flourishing in warm shallow, moderately high energy environ- ments (Skipp, 1969). In comparing them with modern benthic foraminifera, such as Elphidium, Haynes (1981) suggested that the endothyrids lived on the sea bed “clinging by their pseudopods to various plant and animal substrates, crinoids, bryozoans, etc.” During the Late Carboniferous (Bashkirian to Gzhelian), the larger benthic foramin- ifera extended their habitats from the more confined, shallower areas of the shelf or inner ramp. The endothyrids, staffellids and bradynids may have made their habitats in or near high-e nergy environments, patch- reef facies (Dingle et al., 1993; Porta et al., 2005). The small lenticular/ subspherical test facilitate their transport (Rich, 1969). The endothyrids were common in low- and high-e nergy settings. The occurrences of secondary deposits in the endothyrids could have stabilized their tests in turbulent environments (Haynes, 1981). The exclusion of the endothyrids from lagoonal environments with restricted circulation and variable salinity suggests that these forms preferred open- marine envi- ronments. (Porta et al., 2005). The lenticular staffellids, e.g., Pseudoendothyra, are found to be abundant in higher energy, patch-r eef facies, whereas the subspherical forms, e.g. (Staffella) were common in the quieter, back reef facies (Dingle et  al., 1993), in the shallowest setting and in paleoenvironments characterized by abnormally high tempera- tures and salinity. From the Bashkirian to the Middle Permian, the epiphytic staffellids increasingly adopted the ecological setting and taphonomic behaviour of its modern equivalents, the Recent miliolid genus Peneroplis, as it progressively occupied the semi- restricted platform environments (Vachard et al., 2010). The Bradyinidae were probably epiphytes and Bradyina is interpreted as a shallow- water taxon adapted to life in current-s wept environments (Haynes, 1981; Gallagher, 1998; Gallagher and Somerville, 2003). This contrasts with its occurrence in the Pennsylvanian Minturn Formation (Colorado, U.S.A.), where Bradyina seemed to have lived in deeper, open shelf environments at a depth of at least 15m, and relatively deeper than most fusulinids (Stevens, 1971). Malakhova (1961) suggested that the presence of The Palaeozoic Larger Benthic Foraminifera 109 Bradyina in deeper settings indicates the possibility of an adaptation to a planktonic life. However, this is contradicted by its highest abundance in shallow environments. Ponta et al. (2005) suggested that the lateral ‘‘chamberlets’’ of Bradyina were part of a chan- nel system separated from the life chambers and acted as a hydrostatic function which would allow these globose forms to be adapted to a planktonic way of life. The Carboniferous witnessed also the first appearance of the large groups of the fusulinides which gradually occupied a more extensive environmental spectrum than their ancestors. As the fusulinides are an extinct group, their palaeoecology has to be inferred by comparing their shape and their faunal and floral associations with morpho- logically similar groups from the Holocene. They seem to be very well adapted to several distinct ecological and environmental conditions. Their fossil shells occur in light grey, shallow- water limestones or calcareous shales. They are associated with phylloid algae, corals, chaetetids, crinoids, bryozoans, and brachiopods (Thompson, 1964; Ross, 1965, 1967, 1969; Stevens, 1971; Wilson, 1975; Haynes, 1981; Gallagher, 1998; Wahlman, 2002; Wahlman and Konovalova, 2002; Almazán et  al., 2007, Davydov, 2011). This type of benthic carbonate production prevails in shallow, oligotrophic, warm and sunlit environments, with minimal siliclastic input, and with production rates highest in very shallow waters above a maximum depth of 25-3 0m (Weidlich, 2007). However, at some localities they are also abundant in sandstones (Loeblich and Tappan, 1988). The wide variety of coiling modes, from lenticular to subcylindrical or globose, indi- cates many life strategies in shallow water carbonate sub-e nvironments were adopted by the fusulinides (Ross, 1969). Ross found in his study of the palaeoecology of the fusuli- nids that shape was an important factor in their distribution. Elongate, sub-c ylindrical forms (e.g. Parafusulina) occurred in very shallow sub- tidal environments, lagoons and bays, whereas, small forms with inflated chambers occurred in a wide variety of sedi- ments suggesting deeper water to wave base depths (see Fig. 2.23). However, only rare fusulinids occurred in sediments interpreted as forming in deep water environments. The large proloculus in the fusulinides is comparable to those of extant forms such as Marginopora (Ross, 1969). This suggests that the fusulinidean proloculus was formed within the parent shell and released. According to Ross this restricted fusulinidean dis- persal to nearby shallow shelf areas as the released proloculus could only have been carried a short distance by local currents. During the Late Carboniferous, fusulinides forms, such as Profusulinella, acquired heavy chomata. These heavy secondary deposits may completely infill large parts of early chambers so that the protoplasm is concen- trated in the outer whorls. The benefit of these deposits might be the redistribution of protoplasm or, perhaps, the additional weight of the shell may have kept the individ- ual from being easily dislodged by waves or currents (Ross, 1969). Along the upper Bashkirian platform, the abundance of Profusulinella, Pseudostaffella and Ozawainella increases from wave swept areas to the lower- energy facies, whereas, Eostaffella decreases in abundance in decreasing environmental energy (Porta et al, 2005). The marine biota of the fusulinides limestones frequently contains fragments of algae, corals and bryozoa (Plate 2.27). Although in some cases the calcareous algae were abun- dant, they were mostly facies sensitive and too long- ranging to permit the establishment of a precise stratigraphic zonation (Mamet and Zhu, 2005), but their occurrences led to the interpretation that larger foraminifera adopted a symbiotic mode of life, with the fusulinides behaving similarly to Holocene larger foraminifera (Vachard et  al., 2010). 110 Evolution and Geological Significance of Larger Benthic Foraminifera The fusulinides tests are certainly constructed so that they could have hosted photosyn- thetic symbionts. Many modern larger benthic foraminifera are apparently limited in their distribution by the water temperature, sunlight intensity, and the physical and chem- ical requirements of their symbiont (Murray, 2007). In fact, many of the fusulinides have a thin, two- layered wall in the last whorl that lacks the tectorial layers of the early ones. Others, such as many staffellids and boultonids, have translucent walls, and most of the schwagerinids and neoschwagerinids have keriothecal walls with alveoli that could have housed symbionts analogous to extant dinoflagellates (Ross, 1969). Housing symbionts suggests that the occurrence of these faunas is limited to shallow tropical to subtropical waters on shelves, reefs and platforms in the photic zone. Their distribution would be limited by cold oceans and deep oceanic barriers (Haynes, 1981). Vachard et al. (2004) suggested that the cyanobacteria, reported in Holocene Marginopora, are the probable symbionts in the keriothecal walls of the fusulinides, which offered them more light than previously received within the chambers of the early forms with only a tectum. The distribution of fusulinids, staffellids and schubertellids shows varied trends through the Moscovian platform. Fusiella and Schubertella were the only schuber- tellids observed in restricted peri- tidal lithofacies (Baranova and Kabanov, 2003), whereas Fusulinella and Fusulina were dominant in shallow, normal- marine environ- ments. Hemifusulina occurred in subtidal zones affected by storms (Porta et al., 2005). Profusulinella show a preference for shallower settings regardless of environmental energy and salinity (Porta et al., 2005). Schubertella was most common in shallow, higher energy settings (Baranova and Kabanov, 2003), but also tolerated restricted marine conditions. Schubertellids are also common in cooler/d eeper water environ- ments (Teodorovich 1949; Rauser−Chernousova 1951; Baranova and Kabanov 2003; Davydov, 2011). The presence of Schubertella, in the deeper and shallowest facies indi- cate its tolerance of all energy levels and restricted marine conditions. The dominant planispiral fusulinides of the Late Carboniferous to Early Permian exhibits the typi- cal fusiform shape shown by some Holocene larger foraminifera, such as the miliolides (e.g. Alveolinella) that today appear to be confined to normal shallow marine (down to depths of 80m), well-o xygenated, nutrient-r ich, tropical and subtropical waters. It is thus inferred that the fusulinides required normal marine salinity and thrived in shallow warm well- oxygenated, nutrient- rich waters. Their variations in shape, from elongate fusiform to sub- spherical globular may be directly related to specific adaptation to vary- ing conditions in the shallow water environment. The Carboniferous foraminifera lived on the substrata and were primarily sensitive to nutrients availability, physical and chem- ical ecological changes, such as temperature, water currents, and wave intensity. Large forms of Triticites are associated with shallow water algal meadows and banks of cri- noidal fragments, which suggest an environment similar to the modern Gulf of Mexico. Some elongate forms of Triticites are closely associated with sediment of impure silty limestone and fine to medium sandstone that indicates shallow bays, lagoons and wave- built bars and terraces. The occurrence of thick- shelled subglobose Triticites suggests the thickened shell wall was an adaptation to slightly more energetic environments, where resistance to abrasion, crushing, and breaking would be a positive selection fac- tor. Large subglobose species, that have high chambers and only gently folded septa, are common in clay deposits formed in considerably less vigorously agitated environments. Small fusiform Triticites are mostly common in poorly sorted limestones that most The Palaeozoic Larger Benthic Foraminifera 111 Fore reef Sheltered laggons high energy * * * * * Low energy reef Low energy outer shelf Shallow inner shelf Algae * Echinoderm debris Not to scale Ostracods Fig. 2.23. Schematic figure showing the palaeoecological distribution of the fusulinides. probably were deposited in deeper, shelf waters (Ross, 1969), located between 3–8 m deep (Ross, 1971). Another example of shape variation driven by adaptation to different environments are the Early Permian Eoparafusulina, an elongate fusiform form, found living on sand bars in very shallow tidal environments with relatively strong current and wave conditions. The Early Permian Pseudoschwagerina, with inflated chambers with a large protoplasm volume, may even have been pelagic as it is found in a wide variety of sediment types (Ross, 1969). During Late Carboniferous to Permian (Kasimovian to Capitanian), the Schwagerinidae reached the outer limit of the inner or middle platform as they have been commonly reworked in calciturbidites (Vachard et al., 2010). During the Middle Permian, the Neoschwagerinoidea spread rapidly, replacing the Schwagerinoidea in their habitats. The epiphytic verbeekinids and (e.g. Verbeekina) and the neoschwagerinids (e.g. Yabeina) have similar shapes and arrangements of their inter- nal foramina, as in the alveolinids (see Chapter 6) which facilitate the direct movement of both equatorial and radial flows of the cytoplasm (Davydov, 2011). They are also very sensitive to temperature fluctuations and survived only in warm environments where sur- face water exceeds a yearly temperature of about 220 C (Davydov and Arefifard, 2013). Very few taxa survived the hypersaline environments of the Late Permian (Vachard et al., 2010). Only some bradynids (e.g. Glyphostomella), pachyphloiids (e.g. Pachyphloia), schubertellids (e.g. Russiella) and pseudodoliolinids (e.g. Pseudodoliolina) persisted in the very shallow, evaporitic environments of the Lopingian. 2.5 Palaeogeographic Distribution of the Fusulinides and Related Forms The fusulinides and their related forms are found in the Late Palaeozoic basins and adja- cent marine shelves of Eurasia and the Western Hemisphere (Ross, 1967). Geographic 112 Evolution and Geological Significance of Larger Benthic Foraminifera Tectonic closure Jutland volcanic Emeishan volcanic event of E. European Basin event Tectonic onset of the Uralian orogeny 1 Basin closed 2 3 Siberian Major volcanic Traps event: Viluy Tectonic event, orogeny in the 4 Variscan Belt 5 Mississipian Pennsylvianian Cisuralian Guadalupian Lopingian Carboniferous Permian Fig.  2.24. The number of fusulinides genera as a function of time, after Leven (2003). 1)  all data, 2)  Tethys, 3)  East European basin, 4) N America, 5) peri- Gondwana part of Tethys (S. Afghanistan, S. Pamirs, Karakorum, S. Tibet). 3 5 8 . 9 T o u r n a i s i a n 3 4 6 . 7 V i s e a n 3 3 0 . 9 S e r p u k h o v i a n 3 2 3 . 2 B a s h k i r i a n 3 1 5 . 2 M o s c o v i a n 3 0 7 . 0 K a s o m o v i a n 3 0 3 . 7 G z h e l i a n 2 9 8 . 9 A s s e l i a n 2 9 5 . 0 S a k m a r i a n 2 9 0 . 1 A r t i n s k i a n 2 8 3 . 5 K u n g u r i a n 2 7 2 . 9 R o a d i a n 2 6 8 . 8 W o r d i a n 2 6 5 . 1 C a p i t a n i a n 2 5 9 . 1 W u c h i a p i n g i a n 2 5 4 . 1 C h a n g h s i n g i a n 2 5 1 . 9 The Palaeozoic Larger Benthic Foraminifera 113 distribution of these genera forms the basis for recognizing faunal associations, which can be grouped into four phases (see Figs 2.24 and 2.25): • the Tethyan realm (Middle Carboniferous (Visean) to Late Permian), reaching peaks in the Late Carboniferous and Middle Permian; • the East European Basin realm (Middle Carboniferous (Visean) to Late Permian, reaching its peaks in the Late Carboniferous and Early Permian; • the N American realm (Middle Carboniferous (Visean) to Middle Permian, (Capitanian)), reaching its peaks in the Late Carboniferous and Early Permian; • the peri- Gondwana realm (Early Permian (Cisuralian) to Late Permian), reaching its peak in the Middle to Late Permian. The Middle Silurian witnessed an important event in the history of the Palaeozoic foraminifera, namely the appearance of the microgranular wall as a building component of the test. These foraminifera evolved slowly during the Silurian and Devonian and it was not until after the Devonian- Carboniferous boundary that they evolved into many distinctive lineages. Their evolution in the Early Carboniferous was delayed by approx- imately 12.2 Ma as the survivors of the Devonian- Carboniferous boundary extinction event adapted to their new environments. This boundary event is referred to as the Late Devonian Hangenberg extinction (358.9 Ma ago) that occurred at the end of the Famennian and is globally associated with black shale facies (Streel, 1986; McLaren, Period, Epoch Tethyan-Boreal trends of fusulinides Sub-ArccNorth American trends and of fusuliniudes Stage Pseudoschwagerina Pseudoschwagerina 298.9 Quasifusulina Montiparus Triticites Triticites Obsoletes Fusulina Beedeina Beedeina Fusulina Protriticites Fusulinella 315.2 Fusulinella Profusulinella Profusulinella Mllerella Mllerella Fig. 2.25. Trends of evolution of Tethyan and American fusulinides. Age (Ma) Carboniferous Permian Pennsylvanian Cisuralian Bashkirian Moscovian Gzhelian A. S. Ar. K. 114 Evolution and Geological Significance of Larger Benthic Foraminifera 1990). It is marked by the extinction of diverse marine groups, anoxia and rapid sea level fluctuations, and is named from the Hangenberg Shale. Some major groups, such as the ammonoids, stromaporoids and trilobites, suffered an entire extinction (Bambach 2006) and the event saw the collapse of the reef ecosystems (Copper, 2002; Bambach, 2006). The timing of the Hangenberg extinction coincides with the last phase of the Devonian Southern Hemisphere glaciation (Sandberg et  al., 2002; Chen and Tucker, 2003; Haq, 2005). Several causes for this extinction have been suggested, including the widespread development of anoxia and a mini- glaciation (Caplan and Bustin, 1999; Kalvoda, 1989, 2002; Kaiser et al., 2006), possibly triggered or amplified by the develop- ment of the volcanic Viluy Traps in Siberia (Courtillot and Renne, 2003), or an impact event (McLaren and Goodfellow, 1990). The Hangenberg event, however, did not seem to affect the sustainability of the Parathuramminida, Tournayellida or the Endothyrida. As in every extinction event, the small resilient foraminifera (“disaster forms”) survived the end Devonian, but it took the whole of the Tournaisian for recovery, and it was not until the Visean (346.7 Ma) that the shallow reefal benthic foraminifera began to recover and evolve into different lineages. The Parathuramminida declined in number and virtually disappeared within the Early Carboniferous, with only the Tuberitinidae surviving into the Permian. The Tournayellida survived to the Bashkirian, while the endothyrid Loeblichia gave rise to the Fusulinida at the Tournaisian- Visean boundary. The Visean lasted for about 15.8 Ma. During this time, major changes in ocean cir- culation, biogeographic differentiation and high bio-p rovincialism contributed to the diversification of new groups of fauna, such as in the ammonoids, fresh water pelecy- pods, gastropods (Davydov et al., 2004), and of course the fusulinides. During that time animal life, both vertebrate and invertebrate, consolidated its position on land, as plant life had during the Devonian. Euramerica and western Gondwana drifted northwards and moved closer together (Fig. 2.26). This movement eventually gave rise to collision, leading to the Variscan- Hercynian orogeny. The fusulinides thrived from the Visean to the Permian, and gradually filled the reefal niches at the expense of other smaller foraminifera. At the onset of the Visean, they were represented by few genera, but soon they became global in their geographic distribution. They have been found on all continents except Australia, India and Antarctica. This can be explained by the fact that Australia and India were connected to Antarctica throughout the Late Carboniferous and Permian, and were at southern latitudes at which the equatorial fusulinides could not thrive. Based on the reference work of Rauser- Chernousova et  al. (1996), Leven gener- ated a plot showing the generic diversity of fusulinides over the period from their first appearance in the Visean to their complete extinction in the Late Permian. Leven’s data is recast in Fig. 2.27 to show the number of new fusulinides genera throughout the Carboniferous and Permian. It is clear that in Tethys, East Europe and North America a large increase in the number of genera occurred in the Moscovian. Two more peaks of generic increase in the Tethys occur in the Asselian and Capitanian. On the other hand, fusulinides were absent from East Europe and North America in the Late Permian. Fig. 2.28 shows that a high number of extinctions of genera occurred at the Moscovian- Kasimovian boundary, but these were instantly replaced by new genera. The Palaeozoic Larger Benthic Foraminifera 115 Fig. 2.26. Palaeogeographic and tectonic reconstruction of the Late Devonian (by R. Blakey, http:// jan.ucc. nau.edu/ ~rcb7/p aleogeographic.html). On the other hand, the highest number of extinctions, combined with a low number of new genera, is seen at the Capitanian and Wuchiapingian boundary. The fusulinides were almost cosmopolitan, but the American mid- continental prov- ince showed less diversity than those of Tethys and East European Basin (Fig. 2.24). Fusulinides are only found rarely in North Africa, and as the North American mar- gins were largely connected to Palaeotethys (Figs 2.26, 2.29), the fusulinides of the Early Visean showed less provinciality. However, the diversity of fusulinides genera in the North American Basin decreases as the connection with the Tethyan and East European Basin became limited or intermittent from the Late Visean to Permian time. This explains the similarities between the American trend in fusulinides numbers and that of the Tethyan- East European trend plotted by Leven (Fig. 2.24 and 2.27). Tethys was the fusulinides main breeding ground, with more than 900 species being described from this realm (Haynes, 1981). The geographically specific curves of Leven (Fig. 2.24) show that the distribution of the fusulinides greatly depends on the evolu- tionary history of Tethys, and predominantly the equatorial part of that ocean. The 116 Evolution and Geological Significance of Larger Benthic Foraminifera Fig. 2.27. Number of fusulinides genera through the Carboniferous and Permian. fact that the northern hemispheric continental mass was closer to the equator in the Carboniferous (Fig. 2.29) than the southern hemispheric mass explains the wider dis- tribution of the fusulinides in the former hemisphere. The Tournaisian- Visean boundary coincided with the first occurrence of Eoparastaffella simplex Vdovenko, 1954 (Davydov et al., 2004; Devuyst and Kaldova, 2007). The Visean foraminifera were dominated by the endothyrids, which are par- ticularly common in algal limestones (Haynes, 1981). The endothyrids increased steadily during the Visean, however, their increase stopped abruptly at the end of the Serpukhovian. According to Walliser (1995) the faunal overturn was not an abrupt change in all affected fossil groups, but rather just a rapid transition that he correlated with the final orogeny in the Variscan Belt (Figs 2.29 and 2.30). This orogeny was not a catastrophic event and did not cause major changes in the faunal distribution, but rather a slight reduction in the diversity of the faunas. It was associated with cooling associated with changes in ocean circulation patterns and the closing of the equato- rial seaway (Daydov et  al., 2004). This palaeotectonic event (equatorial seaway clo- sure) resulted in the partial, thermal isolation of the Paleo- Tethys, which became a great semi- tropical bay. The climate, which had been fairly equable throughout most of the Mississippian, became more strongly zonal. These dramatic climate changes in the late Serpukhovian occurred at just about the same time as the sudden spread of The Palaeozoic Larger Benthic Foraminifera 117 Fig. 2.28. The number of fusulinides generic extinctions compared to the speciation of genera throughout the Carboniferous and Permian. the fusulinides and the onset of their provincialism. During the latter part of the Early Carboniferous and much of the Middle Carboniferous, seas gradually expanded and flooded low lying portions of continental areas to increase greatly their aerial extent (Ross, 1967). During the Early Carboniferous, genera and species of Ozawainellidae and Staffellidae were common on shallow carbonate shelves and in basins in nearly all parts of the world. At the Serpukhovian-B ashkirian boundary, the diversity of the fusulinides and their related forms increased, with three different provinces becoming recognisable. The East European Basin, characterised by the abundance of Eostaffella and Bradyina; the Tethyan realm, where the palaeotextulariides were abundant; and the North American realm, where Bradyina did not appear before the Bashkirian. The basal beds of the Bashkirian are characterized by the appearance of the foraminiferal species Pseudostaffella antiqua. During the Pennsylvanian, diversity increased steadily in the three provinces of Eastern Europe, Tethys and, to a lesser degree, the Americas (Figs. 2.24, 2.25 and 2.27). New fusulinides, with complicated internal structures, evolved rapidly and new forms such as Fusulina and Fusulinella appeared globally. The increase in size of the fusuli- nids during the Carboniferous coincides with a large increase of atmospheric oxygen levels (Fig. 3.10). This is also confirmed by the presence of giant insects, Meganeura in the Carboniferous (Chapelle and Peck, 1999). According to Moore and Thompson (1949) and Groves et al. (1999), the base of the Moscovian approximates with the first appearance of the genus Profusulinella in the sub-A rctic region of the North American 118 Evolution and Geological Significance of Larger Benthic Foraminifera province. However, this genus had occurred earlier, roughly 4–5 Ma, on the Russian platform of the East European basin. The genus originated in the latter area in the late Early Bashkirian and then underwent significant diversification, so that by the early Moscovian a range of shell morphologies existed (Groves et al., 2007). Although pre- vious work suggested that North American Profusulinella spp. may have been derived from a local ancestor such as Eoschubertella, Groves et  al. (2007) interpret the first sub- Arctic North American species of the genus as immigrants from Eurasia, with their migration through the Franklinian corridor having been facilitated by generally east- to-w est currents during a glacio- eustatic flooding event (Fig. 2.31). This analysis has recently been greatly expanded by Davydov (2014), who points out that the first occurrences of Tethyan fusulinides in North America are associated with paleoclimatic warming events, and that the time of the delay of the first occurrences (see Fig. 2.25) depended on the scale and intensity of the warming episodes during those periods. Fusulinides were by now common in shallow water carbonate banks, which trans- gressed far on to the edges of the land masses along the Tethyan seaway, found today Fig. 2.29. Palaeogeographic and tectonic reconstruction of the Visean (by R. Blakey, http:// jan.ucc.nau. edu/ ~rcb7/ paleogeographic.html). The Palaeozoic Larger Benthic Foraminifera 119 in northern Spain, northern China, Manchuria, and Korea, as well as in North and South America. Towards the end of the Moscovian, a sharp decline in the diversity of the fusulinids in all the three provinces is clearly seen in Figs. 2.27 and 2.28. Most of the fusiform species (e.g. Fusulina, Fusulinella) disappear at the top of the Moscovian and only the small staffellids survived. This extinction may have been related to signifi- cant climate change, which is also thought to have driven the Carboniferous Rainforest Collapse, dated around 305Ma, due to a trend toward increased aridity and changes in global glacial cover (e.g. Groves and Lee, 2008). In the Late Carboniferous, the Tethyan realm contained relatively few fusulinides genera in comparison with earlier (Moscovian) or later (Asselian and Sakmarian) assemblages. The general distribution of the fusulinides in the East European Basin followed the same trend as those in Tethys from the Visean towards the Sakmarian (Fig. 2.24). This can be explained by the fact that the pre-S akmarian East European basin was closely connected to Tethys (Fig. 2.29). After the partial extinction of the fusulinides, at the end of the Moscovian, new forms appeared such as Triticites. These forms dominate the fusulinides assemblages of the Late Pennsylvanian of North America (Loeblich and Tappan, 1988). Wide diversification within Triticites first became pronounced only near the end of the Carboniferous as new, more specialized branches appeared. However, the presence of a few species, at scattered localities in the Western Hemisphere, that are similar to those abundant in eastern Europe (such as Daixina), suggests that the dispersal processes for these groups were limited at this time. Also, the rare occurrence at that time of a species of Triticites in Eastern Europe that is similar to one abundant in the North American realm, suggests that dispersal was also limited in the other direction. Unusually, the Carboniferous- Permian boundary (298.8 Ma) is not marked by any major foraminiferal extinction event. However, it does closely correlate with a large scale volcanism (Fig. 2.32), the Jutland basalt event, which affected Europe and North Africa at this time (Smythe et al., 1995). This event was associated with the develop- ment of the Oslo graben and, amongst others, the Whin Sill in Britain. In the Early Permian (Fig. 2.33) the fusulinides again became diverse and cosmo- politan, but a few genera were not widely distributed. Some are endemic to North American, e.g. Chalaroschwagerina and Cuniculinella, and others are only found in the Tethyan realm, e.g. Sphaeroschwagerina (Plate 2.18, figs 6,7,9) and Zellia (Plate 2.25, Fig. 2). In the Asselian of the Tethyan province, Pseudoschwagerina and Sphaeroschwagerina occur together, while only Pseudoschwagerina occur in both the Tethyan and North American realms. At the top of the Asselian the tectonic closure of the East European basins iso- lated the fusulinides and their numbers dwindled until, at the end of the Kungurian, the tectonic closure became complete, and the East European province fusulinids disappeared completely (Fig. 2.24). During the latter part of the Early Permian, the Schubertelloidea evolved new genera of which Russiella and Minojapanella were endemic to Tethys, while Boultonia was restricted to the Tethyan faunal realm in the early part of the Late Permian, but later it was briefly cosmopolitan before becoming extinct towards the end of the Permian. During this same time, the peri- Gondwanan parts of Tethys (now found in south- ern Afghanistan, southern Pamirs, eastern Hindu Kush, Karakorum, southern Tibet, 120 Evolution and Geological Significance of Larger Benthic Foraminifera Fig. 2.30. Palaeogeographic and tectonic reconstruction of the Late Pennsylvanian world (by R. Blakey, http:// jan.ucc.nau.edu/ ~rcb7/p aleogeographic.html). Himalayas and the Salt Range) moved away from the near glacial conditions in Gondwana towards the equator (Fig. 2.33), thus creating a warm breeding ground for the fusulinides. Their diversity increased in the Sakmarian, before it decreased again at the end of the stage. Many species then disappeared gradually towards the end of the Early Permian, such as the cosmopolitan Pseudoschwagerina (Plate 2.18, figs 3,4,8,11), which appearance defines the base of Permian boundary and disappearance the top of the Kungurian (see Fig. 2.17). Newly recorded occurrences of fusulinides assemblages from the Sakmarian of Central Oman (Angiolini et  al., 2006)  have extended the distribution of the Pseudofusulina (Plate 2.18, Fig. 13) species to the South Tethyan seas. This assemblage compares with those described by Leven from the peri-G ondwanan parts of Tethys (Leven, 1993, 1997). Their presence in Oman expands the area of distribution of the warm water fusulinides during the Sakmarian, indicating the onset of the beginning of the “greenhouse” climate which resulted in the Gondwanan deglaciation. The Palaeozoic Larger Benthic Foraminifera 121 Fig. 2.31. Migration route of Profusulinella spp. from Eurasia to the sub- Arctic North American province through the Franklinian corridor (modified from R. Blakey http://j an.ucc.nau.edu/ ~rcb7/ paleogeographic. html). The decrease in the diversity at the end of the Sakmarian probably coincides with the Uralian orogeny and the end of the Permo-C arboniferous glaciation (Erwin, 1996). It affected the diversity in the East European basins, and the development of the Eastern Tethyan foraminifera. These new foraminiferal assemblages, which fil- led in the empty niches after the Early Permian (Kungurian), were dominated by the verbeekinids. These latter reached their peak in the Wordian (Leven, 2003). The ver- beekinid association forms the main part of the Permian Tethyan fusulinides faunal realm and is found from modern day Tunis to Greece, Yugoslavia, Sicily, Afghanistan and Timor (and New Zealand) and in eastern Asia and the Japanese islands. However, many species such as Verbeekina reach as far south as south California in the North American realm. The Capitanian saw a slight increase in the diversity of the fusulinides during a short- term transgression in that stage (Leven, 2003). Immediately after the trans- gression, the North American basin became isolated and underwent a rapid salini- zation that caused the complete extinction of the fusulinides in North America. The Late Guadalupian saw an extinction that was one of the largest in the Palaeozoic. All large and morphologically complex forms assigned to the Schwagerinidae and Neoschwagerinidae were eliminated. It has been suggested that this extinction could have been triggered to a flood basal event which occurred in the Late Permian in south- western China. The Emeishan basalts extend (Fig. 2.34) over an area in excess of half a million square kilometres (Courtillot and Renne, 2003). The Late Guadalupian cri- sis affected the fusulinides more than the other major extinction event in the Late Moscovian. According to Vachard et al. (2010), two causes can explain the total dis- appearance of the keriothecal forms, Schwagerinoidea and Neoschwagerinoidea, the disappearance of the endosymbionts of the giant foraminifers and the oceanic cooling event, the high- productivity “Kamura” event. This event is estimated to have lasted over 3–4 Ma (Isozaki, 2007). 122 Evolution and Geological Significance of Larger Benthic Foraminifera Fig. 2.32. The Jutland Flood Basalt event at the end of the Carboniferous. The remaining fusulinides biota were under stress during the Lopingian and diversity was in decline. The end Lopingian is, however, marked by the end Permian extinction: the most severe of the entire Phanerozoic. All of the inshore taxa (98%), including the large fusulinides, became extinct. Some small endothyrids survived, only to die out subsequently in the Early Triassic. Globally, 90 to 96% of all marine inver- tebrate species went extinct (Sepkoski, 1986) as did all but one of 90 genera of reptiles (McLaren and Goodfellow, 1990), most corals, brachiopods and large terrestrial spe- cies (see Benton (2005) for an extensive review). In a comprehensive study of the end Permian mass extinction horizon recorded in the Meishan section, South China, Kaiho et al. (2001) recorded that the extinc- tion event was characterized by the abrupt and catastrophic disappearance of major benthos, and according to these authors, the extinction horizon coincides with an abrupt decrease in the 34S/ 32S ratios of seawater sulphate, 87Sr/ 86Sr ratio, and an increase in Fe  – Ni grains. There was also a pronounced negative excursion in the carbon recorded in P– Tr boundary carbonate rocks and organic matter (e.g. Berner, 2002). The cause of the end Permian mass extinction remains debatable, and numerous theories have been formulated to explain the events of the extinction. Historically, geologists invoked in the past theories such as climate change, global warming, marine The Palaeozoic Larger Benthic Foraminifera 123 Fig. 2.33 Palaeogeographic and tectonic reconstruction of the early Permian world (by R. Blakey, http:// jan.ucc.nau.edu/~ rcb7/ paleogeographic.html). anoxia or tectonic processes as the cause of such global mass extinctions. But these are difficult to reconcile with the relatively abrupt nature of such extinction events. Furthermore, cause and effect are often difficult to differentiate, and indeed, the need to make a distinction between kill and trigger mechanism was stressed by Knoll et al. (2007), who referred to the first as the physiologically disruptive process that causes death, and to the second by the critical disturbance that brings one or more kill mech- anisms into play. So, if gradualist processes cannot provide an explanation for the rapidity of the observed event, then catastrophic causes of the P- Tr mass extinction must be invoked. These have included either an asteroid impact or flood basalt volcan- ism. However, even though many of the ecological features of the P-T r event compare with the results of the asteroid impact that famously is thought to have concluded the Mesozoic, the end Palaeozoic strata have not yielded an unambiguous signature of a bolide impact (Wardlaw et al., 2004). This leads many researchers to consider that a major volcanic event was the more probable cause of the end Permian extinction 124 Evolution and Geological Significance of Larger Benthic Foraminifera (Benton and Twitchett, 2003). The P- Tr mass extinction coincides with the eruption of the Siberian traps flood basalts (Fig. 2.26), the largest known Phanerozoic conti- nental igneous province (Reichow et al., 2002; Courtillot et al., 2003). The eruption of this vast volume of basalt, in a short time (< 1 Ma), would have released aerosols and greenhouse gases, which could have triggered a rapid climate change that would have caused a mass extinction of both marine and continental biota (Erwin et  al., 1994; Benton and Twitchett, 2003; Chen and Benton, 2012). A sudden release of huge volumes of carbon dioxide might have poisoned all marine and terrestrial life (Ward and Brownlee, 2000). The event may have also triggered widespread marine anoxia, and the destabilization of seafloor methane clathrates, which could account for the carbon isotope excursion noted by Berner (2002). Siberian Trap volcanism was a major trigger of these extinctions, especially if it was combined with the profound sea- level low stand, unprecedented global high temperatures and marine deep- water stagnation and anoxia (Macleod, 2013). Despite the globally devastating effect of the end Permian event on foraminiferal life, some creatures survived. For the forms related to the fusulinides, the survival of the endothyrides can perhaps be explained by their small size and by their not needing symbionts or much oxygen for existence. This allowed them to survive the Fig. 2.34 The Emeishan Large Igneous Province and the Siberians Traps. The Palaeozoic Larger Benthic Foraminifera 125 adverse conditions of the end of the Permian, and to occupy the empty niches and the new ecosystems found in the Early Triassic (see Chapter 3). According to Knoll et al. (2007), hypercapnia (being able to cope with the physiological effects of elevated PCO2) is a feature that seemed to be common to the survivors of the P- Tr event. This may provide an explanation of the significant loss of most calcareous foraminifera, but the survival of the small agglutinated foraminifera. It also seems that other reefal groups were affected in the same way, so that the Late Permian corals disappeared but their unskeletonized relatives, the sea anemones (which would give rise to scleractinian cor- als in the Triassic) survived (Knoll et al., 2007). Similarly, the skeletonized dasyclad green algae disappeared, but multiple unskeletonized sister groups survived (Aguirre and Riding, 2005). Whatever the catastrophic cause and the physiological needs of the survivors, the large fusulinides never recovered after the end Permian event. The occurrence of two consecutive extinction events in a relatively short time (the end Guadalupian crisis and the end Permian crisis were only ~10 Ma apart) exhausted this group and completely destroyed their ability to continue to fill their shallow marine niche. This left a major opening for new fauna and ecosystems to develop. As will be seen in Chapter 3, the Triassic was a very different world (Chen and Benton, 2012). B 1 2 3 4 5 A 6 7 8 9 A B B A 10 11 12 A B 13 14 15 16 17 18 Plate 2.1 Scale bars = 0.4mm Fig. 1-3 , 5. Earlandia elegans Rauser-C hernousova and Reitlinger, Vertical sections, Simonstone Limestone, Whitfield Gill, Yorkshire, England, 1) UCL coll., DH 109; 2) UCL coll., DH46; 3) UCL coll., D150 A; 5) UCL coll., DH688. Fig. 4. Earlandia vulgaris (Rauser- Chernousova and Reitlinger), Gayle Limestone, Duerley Beck, Gayle, Yorkshire, England, UCL coll., DH81. Fig.  6. A) Draffania biloba Cummings, B) Earlandia sp., Transverse section, Gayle Limestone, Duerley Beck, Gayle, Yorkshire, England, UCL coll., DH49. Figs 7-9 Brunsiella sp., 7) UCL coll., DH 109; 8) UCL coll., DH46; 9) UCL coll., DH82. Fig. 10. A) Draffania biloba Cummings B) Palaeonubecularia cf. uniserialis Reitlinger, Hawes Limestone, Gayle Beck, Yorkshire, England, UCL coll., DH 33a. Fig.  11. A) Howchinia sp., B) Saccaminopsis fusulinaformis (M’Coy), Hawes Limestone, Gayle Beck, Yorkshire, England, UCL coll., DH 33a. Fig. 12. Palaeonubecularia cf. uniserialis Reitlinger, Gayle Limestone, Duerley Beck, Gayle, Yorkshire, England, UCL coll., DH48. Figs 13, 17, 18. Draffania biloba Cummings, Gayle Limestone, Duerley Beck, Gayle, Yorkshire, England, 13A) axial section, 13B) transverse section, UCL coll., DH 81; 17) UCL coll., DH 272; 18)  transverse section, Gayle Limestone, Duerley Beck, Gayle, Yorkshire, England, UCL coll., DH 811. Fig.  14. Alga incerta, Gayle Limestone, Duerley Beck, Gayle, Yorkshire, England, UCL coll., DH 81. Fig. 15. Diplosphaerina sphaerica (Derville), Underset Limestone, Howgate Head, Sleddale, UCL coll., DH327. Fig. 16. Diplosphaerina inaequalis (Derville), Simonstone Limestone, Whitfield Gill, Yorkshire, England, UCL coll., DH117. Plate 2.2 Scale bars: Figs 1- 2, 4, 6-1 1, 17- 19 = 0.25mm; Figs 3, 5, 13 = 0.4mm; Figs 12, 15 = 1mm. Figs 1, 5. Tetrataxis conica Ehrenberg. 1) figured by Al- Habeeb from Mumbles, South Wales, M14b/B ; 52) Gayle limestone, Duerley Beck, Gayle, Yorkshire England. Fig. 2. Tetrataxis pusillus Conyl and Lys, Norton Quarry, South Wales, N6A/ A, UCL coll. Fig. 3. Valvulinella lata Grozdilova and Lebedeva, Gayle limestone, Duerley Beck, Gayle, UCL coll., DH 46. Fig. 4. Valvulinella tchotchiai Grozdilova and Lebedeva, Gayle limestone, Duerley Beck, Gayle, Yorkshire England, UCL coll. Fig.  6. Eotuberitina cornuta Hallet, Moss Kennels, Northumberland, England, UCL coll., DH349. Fig. 7. Tubeporina magnifica Hallet, Simonstone Limestone, Whitfield Gill, Yorkshire, England, UCL coll., DH117. Fig.  8. Eotuberitina sp., figured by Hallet (1966) from Trowbarrow Quarry, Silverdale, UCL coll., DH531. Fig. 9. Eotuberitina reitlingerae Miklucho- Maclay, Simonstone Limestone, Whitfield Gill, Yorkshire, England, UCL coll., JER 867. Fig. 10. Diplosphaerina inaequalis (Derville), figured by Hallet (1966) from Arngill Beck, Askrigg, Yorkshire, England, UCL coll., DH586. Fig. 11. Tuberitina sp., Simonstone Limestone, Whitfield Gill, Yorkshire, England, UCL coll., JER 867. Figs 12-1 5. Saccaminopsis fusulinaformis (M’Coy), 12- 14) figured by Hallet (1966) from 12) Buckden Beck, UCL coll., DH4; 13-1 4) Hawes Limestone, Duerley Beck, UCL coll., DH35; 15) Gayle Beck, Hawes Limestone, Yorkshire, England, UCL coll., DH47. Fig. 16. Draffania biloba Cummings, figured by Hallet (1966) from the Simonstone Limestone, Whitfield Gill, Yorkshire, England, UCL coll., DH613. Fig.  17. Paratuberitina sp., Hawes Limestone, Duerley Beck, UCL coll., DH35. Fig. 18. Eotuberitina fornicata Hallett, Gayle limestone, Duerley Beck, Gayle, Yorkshire, England, UCL coll., DH10. Fig. 19. Parathurammina aper- turata Pronina. Middle Devonian, hypotype, from Zadorozhnyy and Yuferev (1984), Tomsk District, USSR. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 A A B B A C C 1 2 A B 3 4 5 A B C 6 7 8 9 10 11 12 Plate 2.3 Scale bars: Figs 1 - 12 = 0.25mm. Fig. 1. Thin section photomicrographs of A) Earlandia vulga- ris (Rauser- Chernousova), B) Chrysothurammina sp., C) Draffania biloba Cummings, Gayle Beck, Hawes Limestone, Yorkshire, England, UCL coll., DH50. Fig. 2. A) Millerella designata Zeller, B) Elenella Pronina, C) Plectogyra irregularis Zeller, Hardraw Scar Limestone, Muker, Yorkshire Limestone, England, UCL coll., DH 708. Fig. 3. Sogdianina angulata Saltovskaya, Early Carboniferous (Visean), Tadzhikistan, USSR, cen- tred section, from Petrova (1981). Fig. 4. A) Eostaffella spp., B) Auroria, Gayle Beck, Hawes Limestone, Yorkshire, England, UCL coll., DH50. Fig. 5. Eovolutina sp., Antropov, Hardraw Scar Limestone, Muker, Yorkshire Limestone, England, UCL coll., DH 708. Fig.  6. Parphia Miklukho-M aklay, Draffania biloba Cummings, Archaediscus sp., Hardraw Scar Limestone, Muker, Yorkshire Limestone, England, UCL coll., DH 708. Fig. 7. Ostracod sp., Simonstone Limestone, Whitfield Gill, Yorkshire, England, UCL coll., DH115. Fig. 8 Tuberitina Galloway and Harlton, a specimen showing bulbous chambers in straight series, Gayle Beck, Hawes Limestone, Yorkshire, England, UCL coll., DH152. Fig. 9. Brunsia Mikhaylov, Gayle Beck, Hawes Limestone, Yorkshire, England, UCL coll., DH33. Fig. 10. Biseriella parva, late Visean, S. Urals, USSR from Chernysheva (1948). Fig. 11. Palaeospiroplectammina Lipina, Tournaisian section showing early coil, Russian Platform, USSR, from Lipina (1965). Fig. 12. Abadehella tarazi Okimura and Ishi, axial section of holotype from Okimura et al., (1975), Abadeh Formation, central Iran. 1 2 3 4 5 6 7 8 9 10 161 12 13 14 15 16 C B C A 17 18 Plate 2.4 Scale bars: Figs 1-6 , 9, 12- 15, 17, 18 = 0.25mm; Figs 7, 8, 10, 11, 16 = 1mm. Fig. 1. Cribrostomum inflatum Cummings, Gayle Limestone, Duerley Beck, Gayle, Yorkshire, England, UCL coll., DH48. Figs 2- 4, 9, 14. Palaeotextularia longiseptata Lipina, 2, 4, 9) Gayle Limestone, Duerley Beck, Gayle, UCL coll., DH44; 3) Hawes Limestone, UCL coll., DH22; 14) figured by Al-H abeeb (1977) from Port Eynon, South Wales. Figs 5, 13, 15. Cribrostomum sp., 5) Gayle Limestone, Duerley Beck, Gayle, UCL coll., DH11; 13) fig- ured by Al-H abeeb (1977) from Norton Quarry, South Wales; 15) Carboniferous Limestone, West Sahara, B188. Fig.  6. Koskinotextularia sp., Simonstone Limestone, Whitfield Gill, UCL coll., DH120. Fig.  7. Climacammina sp., Carboniferous Limestone, West Sahara, B1 288. Fig. 8. Deckerella quadrata Cummings, Simonstone Limestone, Whitfield Gill, Yorkshire, England, UCL coll., DH110. Fig. 10. Palaeotextularia sp., Simonstone Limestone, Whitfield Gill, UCL coll., DH33. Figs 11, 16. Deckerellina sp., 11) Simonstone Limestone, Whitfield Gill, UCL coll., DH3. 16) Deckerellina sp., Carboniferous Limestone, West Sahara, B1 426, UCL coll.. Figs 12. Palaeotextularia sp., Schlykovn, Limestone, Duerley Beck, Gayle, UCL coll., DH 50. Fig. 17. A) Palaeotextularia longiseptata Lipina, B) Tetrataxis conica Ehrenberg, C) Archaediscus inflatus Schlykovn, Limestone, Duerley Beck, Gayle, UCL coll., DH 50. Fig. 18. Palaeotextularia angulata Cummings, Cowey Sike, Grindon Hills, Northumberland, England, UCL coll., JER 1086. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 A B 18 18 21 19 20 Plate 2.5 Scale bars:  Figs 1-2 0  =  0.25mm. Fig.  1. Howchinia bradyana (Howchin). Gayle Beck, Hawes Limestone, Yorkshire, England, UCL coll. Figs 2. Pachyphloia asymmetrica (Miklukho- Maklay), Permian, Thai Peninsula, Thailand, AM 35b, NHM OGS coll. 231. Fig. 3. Pachyphloia depressa (Miklukho- Maklay), Permian, Thai Peninsula, Thailand, AM 35b, NHM OGS coll. 231. Fig. 4. Pachyphloia magna (Miklukho- Maklay), Permian, Thai Peninsula, Thailand, AM 35b, NHM OGS coll. 231. Fig. 5. Paratikhinella cylindrica (Brady), Main Limestone, Fossdale Gill, UCL coll., DH150. Fig. 6. Palaeotextularia sp., TP18, Permian, Khlong Pha Saeng, Thailand, Map. 4737 I. 808879, AHG Mitchell OGS coll. 231. Fig. 7. Climacammina sp., a twisted vertical section, Gayle Beck, Hawes Limestone, Yorkshire, England, UCL coll. Fig. 8. Lucammina jonesi (Brady), Hydraulic Limestone, Staffordshire, UK, UCL coll., F6. Figs 9- 10. Climacammina antiqua (Brady), Permian, Thitsipin Limestone Formation., Shan States Burma, NHM coll., 1970/2 4. Fig.  11. Nodosaria radicula (Linné) showing atelo-m onolamellar wall structure. Permian, Brady, NHM coll., fig- ured in Carboniferous and Permian Monograph, 1876. Figs 12- 17. Glomospirella paseudopulchra Lipina, 12- 13, 15) Gayle Limestone, Duerley Beck, Gayle, UCL coll., DH49; 14) Hardraw Limestone, Clough River, JER868; 16- 17) Hawes Limestone, Duerley Beck, UCL coll., DH38. Fig.  18. A) Calcifolium punctatum Maslov, a codiacean alga; B) Archaediscus karreri Bradi, Carboniferous, UK, UCL coll. Fig. 19. A pack- stone of Eostaffella mosquensis Vissarionova, Hawes Limestone, Duerley Beck, Yorkshire, England, UCL coll., DH802. Fig. 20. A packstone of Endothyra spp., Carboniferous Limestone, Staffordshire, UCL coll.. Plate 2.6 Scale bars; Figs 1-1 6 = 0.4mm. Figs 1-2 . Endothyra sp., SEM photographs figured by Al-H abeeb (1977) from Kittle, South Wales K85, 1)  of etched polished surface of a specimen in limestone; 2)  Part of the outer wall. Figs 3, 7- 9. Archaediscus complanatus Conil, Gayle Limestone, Duerley Beck, Hawes, Yorkshire, England, UCL coll., DH 50; 9) figured by Al- Habeeb (1977) from Oxwich, South Wales. Fig. 4. Archaediscus gigas Rauser- Chernousova, Gayle Limestone, Duerley Beck, Hawes, Yorkshire, England, UCL coll., DH32. Fig. 5. Archaediscus electus Ganelina, Simonstone Limestone, Whitfield Gill, Yorkshire, England, UCL coll., DH115. Fig. 6. Archaediscus sp., an SEM enlargement of the outer wall figured by Al- Habeeb (1977) from Wrexham, South Wales, W3. Figs 10, 12, 14, 15, 16. Archaediscus karreri Brady, 10) fig- ured by Al-H abeeb (1977) from Mumbles, South Wales, M5B/A ; Gayle Limestone, Duerley Beck, Hawes, Yorkshire, England, UCL coll., DH50. Fig. 11. Archaediscus inflatus Schlykova, Gayle Limestone, Duerley Beck, Hawes, Yorkshire, England, UCL coll., DH50. Fig. 13. Archaediscus sp., Carboniferous Limestone, West Sahara, UCL coll., B1-4 96. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 Plate 2.7 Scale bars: Figs 1-6  = 0.15; Figs 7-1 8 = 0.4mm Fig. 1. Hemiarchaediscus angulatus (Soanina), fig- ured by Al-H abeeb (1977) from Three Yard Limestone, Gray Gill, Yorkshire, England, UCL coll., DH227. Fig. 2. Hemiarchaediscus compressus Al- Habeeb, 1977. Paratype, Underset Limestone, Cowgill Beck, Widdale Fell, UCL coll., DH322. Fig. 3. Asteroachaediscus pressulus (Grozdilova and Lebedeva), figured by Al-H abeeb (1977) from Main Limestone, Gunnerside Beck, Swaledale, UCL coll., DH714. Fig. 4. Propermodiscus sp., Simonstone Limestone, Whitfield Gill, Yorkshire, England, UCL coll., DH 109. Fig. 5. Planoarchaediscus emphaticus Al- Habeeb, holotype, Limestone IVA, River Cover, UCL coll., DH567. Fig. 6. Archaediscus sp., Gayle Beck, Hawes Limestone, Yorkshire, England, UCL coll., DH 36a. Fig.7. Loeblichia sp., Carboniferous Limestone, Sahara, UCL coll., B1693. Figs 8, 12, 13. Tetrataxis Bradyi Hallett, 1966, 8) AM 82 Permian Thitsipin Limestone Formation, 469491, southern Shan States Burma. NHM coll., OGS 197/1 1; 12, 13) Gayle Limestone, Duerley Beck, Gayle, UCL coll., DH52. Figs 9, 10, 11, 14. Tetrataxis conica Ehrenberg, Hawes Limestone, Duerley Beck, Hawes, Yorkshire, England, 9) UCL coll., DH137A; 11) UCL coll., DH349; 14) UCL coll., DH50. Figs 16, 17. Tetrataxis sp., Yorkshire, England, 16) East Stone Gill, Coverdale, UCL coll., DH272; Hardraw Scar Limestone, Hardraw Force, UCL coll., DH77. Fig. 18. Tetrataxis subcylindricus Conil and Lys, Gayle Beck, Hawes Limestone, Yorkshire, England, UCL coll., DH 50. 1 3 2 5 4 6 7 8 9 10 A B 11 12 13 Plate 2.8 Scale bars: Figs 1-4 , 11,12 = 0.3mm; Figs 5- 10, 13 = 1mm. Figs 1- 4. Profusulinella sp., fusulinid Beds Upper Coal Measures, Southern Iowa, UCL coll. Figs 5. Dunbarinella ervinensis Thompson, Camp Creek foraminifera, on Saddle Creek, 9 m south of Rockwood, Texas, UCL coll. Fig.  6. Fusulina cylin- drica Fischer de Walheim, Limestone, Elmdale Formation Upper Gzhelian, Oklahoma, USA, UCL coll. Fig. 7. Montiparus montiparus (Rozovskaya), Medvedka River, USSR, UCL coll. Figs 8. Pseudofusulinella occidentalis (Thompson and Wheeler), Early Permian (Sakmarian), McCloud Limestone, California, USA, UCL coll. Fig. 9. Fusulina sp., Wolfcamp Beds, Wolfcamp Hills, 15m N. E. Marathon, Texas, UCL coll. Fig. 10. Triticites ventricosus (Meek and Hayden), equatorial and axial sections, Early Permian, Wolfcamp Fm., Wolfcamp Bed, 15m N.E. Marathon, Texas, UCL coll. Fig. 11. Millerella tortula Zeller, Carboniferous Limestone, West Sahara, UCL coll., B1 288. Fig. 12. A) Asteroachaediscus karreri Brady, B) Loeblichia sp., Carboniferous Limestone, West Sahara, UCL coll., B 1426. Fig. 13. Bradyina sp., Carboniferous Limestone, West Sahara, UCL coll., B1 1647a, 1 2 3 4 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 Plate 2.9 Scale bars Figs 1, 12 = 0.25mm; Figs 2- 11; 13-2 0 = 0.4mm. Fig. 1. Endothyranella sp., solid spec- imen, Carboniferous SEM, UCL coll. Fig. 2. Pseudoendothyra luminosa Ganelina, figured by Al- Habeeb (1977) from Port Eynon, South Wales, P35/ B. Fig. 3. Endothyranopsis sp., Carboniferous Limestone, West Sahara, UCL coll., B1 288g. Fig. 4. Valvulinella youngi (Brady). Hawes Limestone, Duerley Beck, Hawes, Yorkshire, England, UCL coll., DH37. Fig. 5. Eostaffella arcuata (Durkina, 1959) figured by Al-H abeeb (1977) from Pant Mawr, South Wales, TR6/ A. Fig.  6, 7.  Endothyranopsis crassa (Brady), figured by Al- Habeeb (1977) from Port Eynon, South Wales, 6) P28/ A; 7) K30/ B. Fig. 8, 10. Mstinia cf. bulloides Mikhailov, 8) Gayle Beck, Hawes Limestone, Yorkshire, England, UCL coll., DH 36a; 10) figured by Hallet (1966) from Gayle Limestone, Duerley Beck, Gayle, Yorkshire, England, UCL coll., DH4905. Fig. 9, 11, 12. Brunsiella buskensis (Brazhnikova), Hardraw Scar Limestone, Muker, Swaledale, Yorkshire, England, UCL coll., DH 708, 9) oblique axial section; 12) equatorial section; 11) axial section, Gayle Limestone, Duerley Beck, Gayle, Yorkshire, England, UCL coll., DH 46. Fig. 13. Pseudoendothyra composite (Dutkevitch), Hardraw Scar Limestone, Muker, Swaledale, Yorkshire, England, UCL coll., DH 708. Fig.  14. Carbonella sp., figured by Hallett (1966) from Gayle Limestone, Duerley Beck, Gayle, Yorkshire, England, UCL coll., DH4905. Fig. 15. Spinoendothyra phrissa (Zeller), figured by Al- Habeeb (1977) from Pwll Du, South Wales, DP3/D . Fig. 16. Chernyshinella exelikta (Conil and Lys), figured by Al- Habeeb (1977) from Kittle section, South Wales, K75/ E. Fig. 17. Forschia cf. subangulata Moller, figured by Al-H abeeb (1977) from Kittle section, South Wales, K85/ A. Fig. 18. Planoendothyra cf. aljutovica Reitlinger, figured by Al- Habeeb (1977) from Port Eynon, South Wales, 35/ 6. Fig. 19. Mikhailovella sp., Simonstone Limestone, Whitfield Gill, Yorkshire, England, UCL coll., DH 109. Fig. 20. Bradyina rotula (Eichwald), Simonstone Limestone, Whitfield Gill, Yorkshire, England, UCL coll., DH 110. B A 1 2 3 4 5 6 7 8 9 10 11 12 13 14 Plate 2.10 Scale bars: Figs 1-1 4 = 0.3mm Figs 1- 7. Howchinia bradyana (Howchin), 1) Hawes Limestone, Duerley Beck, Hawes, Yorkshire, England, UCL coll., DH47; 2-5 ) Simonstone Limestone, Whitfield Gill, Yorkshire, England, UCL coll., DH117. Figs 8-1 2. Vissariotaxis cummingsi Hallet, Yorkshire, England, UCL coll., 8)  Hawes Limestone; 9)  Simonstone Limestone; 10)  Gayle Limestone, Duerley Beck, Gayle, Yorkshire, England; 11- 12) Moss Kennels, Northumberland, UCL coll., DH 349. Figs 13- 14. Howchinia nuda Hallet, Moss Kennels, Northumberland, Yorkshire, England, UCL coll. 1 2 Plate 2.11 Scale bars: Figs 1- 2 = 2mm. Fig. 1. Assemblages of Triticites sp., Carboniferous Gap Tank strata (bed 10), 17m S.E. of Gap Tank, 23m N.N.E. of Marathon, Texas, UCL coll. Fig. 2. Assemblages of Triticites patulus Dunbar and Newell, Wolfcamp Beds, Wolfcamp Hills, 15 m. N. E. Marathon Texas, UCL coll. 1 2 3 4 5 6 7 8 9 10 11 12 B B A 13 14 15 Plate 2.12 Scale bars: 1- 15 = 1mm. Figs 1, 3. Plectogyra sp., Carboniferous Limestone, West Sahara, UCL coll., 1) B1 1630 (b); 3) B1 1647 (a). Fig. 2. Bradyina rotula (Eichwald), Carboniferous Limestone, West Sahara, UCL coll., B1 1647a. Fig. 4. Eostaffella radiata (Brady), Hardraw Scar Limestone, Hardraw Force, Yorkshire, England, UCL coll. Fig. 5. Plectogyra cf. geniculata (Ganelina), Gayle Beck, Hawes Limestone, Yorkshire, England, UCL coll., DH37. Fig. 6. Millerella tortula Zeller, Hunter’s Stone Bank, Coverdale, Yorkshire, England, UCL coll., DH562. Fig. 7. Endothyranopsis crassa (Brady), Main Limestone, Fossdale Gill, Yorkshire, England, UCL coll., DH151. Fig. 8. Endothyra bowmani Brown, Indiana, USA, UCL coll. Fig. 9. Eostaffella ornata (Brady), figured by Al- Habeeb (1977) from Oxwich, South Wales, K28/ A,. Fig. 10. Endothyranopsis aff. E. macrus (Zeller), figured by Al - Habeeb (1977) from Kittle, South Wales. Fig. 11. Endothyranopsis pechorica (Rauser-C hernousova), Hardraw Scar Limestone, Hardraw Force, Yorkshire, England, UCL coll., DH77. Fig. 12. Planoendothyra mameti Al Habeeb, holotype, figured by Al- Habeeb (1977), Kittle, South Wales K38/ A. Figs 13, 14. Eostaffella spp. 13) Hardraw Limestone, Hardraw Force, Yorkshire England, UCL coll., DH77; 14)  Three Yard Limestone, Walden Beck, Yorkshire Limestone, England, UCL coll., DH699. Fig.  15. A) Pseudoendothyra struvii (von Möller), B) Brunsiella buskensis (Brazhnikova), Three Yard Limestone, Walden Beck, Yorkshire Limestone, England, UCL coll., DH699. 1 2 3 4 5 6 7 8 9 B A A 10 B 11 Plate 2.13 Scale bars:  Figs 1- 11  =  0.25 Figs 1- 3. Endothyra bowmani Brown, Hydraulic Limestone, Waterhouses, Leek, Staffordshire, UCL coll. Fig.  4-9 . Endothyranopsis crassa (Brady), 4)  Carboniferous Limestone, West Sahara UCL coll., B1-2 88e; 5)  Moss Kennels, Northumberland, England, UCL coll., DH349; 6-9 ) Carboniferous Limestone, West Sahara, UCL coll., B1- 1630. Fig.  10. A) Bradyina sp., B) Plectogyra sp., Carboniferous Limestone, West Sahara, UCL coll., B1647. Fig.  11. A) Endothyranopsis crassa (Brady), B) Loeblichia sp., Carboniferous Limestone, West Sahara, UCL coll., B1- 1630. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 Plate 2.14 Scale bar: Figs 1- 15 = 0.4mm. Figs 1, 4, 5. Endothyranopsis crassa (Brady), Main Limestone, Fossdale Gill, Yorkshire, England, UCL coll. Figs 2, 13. Pseudoendothyra struvii (von Möller), Yorkshire, England, UCL coll., 2) Main Limestone, Fossdale Gill, UCL coll., DH79; 13) Hardraw Limestone, UCL coll., DH78. Fig. 3. Plectogyra cf. pandorae Zeller, Main Limestone, Fossdale Gill, Yorkshire England, UCL coll., DH 146 A2. Fig. 6. Plectogyra bradyi (Mikhailov), Carboniferous Limestone, West Sahara, UCL coll., B1 288 (g). Figs 7, 11. Plectogyra irregularis Zeller, Carboniferous Limestone, West Sahara UCL coll., 7) B1 1661; 11) B1 288(a). Fig. 8. Plectogyra excellens Zeller, Gayle Beck, Hawes Limestone, Yorkshire, England, UCL coll., DH33 3a. Fig. 9. Plectogyra phrissa Zeller, Gayle Beck, Hawes Limestone, Yorkshire, England, UCL coll., DH33a Fig. 10. Millerella designata Zeller, Hardraw Scar Limestone, Muker, Yorkshire, England, UCL coll., DH 708. Fig. 12. Plectogyra sp., Carboniferous Limestone, West Sahara UCL coll., B1 288(d). Figs 14, 15. Eostaffella mosquensis Vissarionova, Hardraw Scar Limestone, Hardraw Force, Yorkshire, England, UCL coll., DH80. 2 1 3 4 5 7 7 6 8 9 Plate 2.15 Scale bars: Figs 1-3 , 6, 8 = 2mm; Figs 4, 5, 7, 9 = 0.25mm. Fig. 1-3 , 7. Triticites ventricosus (Meek and Hayden), equatorial and axial sections, Early Permian, Wolfcamp Formation, Wolfcamp Bed, 15m N.E. Marathon, Texas, UCL coll. Fig. 4. Eostaffella sp., Macal Shale Group, Rio Trio, British Honduras, axial section, NHM P44725 (where wrongly identified as Ozawainella spp.). Fig. 5. Schwagerina sp., oblique axial section, Macal Shale Group, Rio Trio, British Honduras, NHM 99147. Fig.6. Fusiella sp., Am 87, Permian, Thitsipin Limestone Formation, 477 484, South Shan States Burma, NHM OGS coll. (where wrongly identified as Wedekindellina sp.), 1970/ 12. Fig. 8. Quasifusulina sp., Am 87, Permian, Thitsipin Limestone Formation, 477 484, South Shan States Burma. NHM OGS coll. (where it is identified as Fusulina prima Thompson), 1970/ 12. Fig. 9. Verbeekina verbeeki (Geinitz), Permian, Sumatra H.B., Brady, NHM coll. 1 2 3 4 4 5 5 6 7 6 9 8 9 10 11 10 1121 Plate 2.16 Scale bars: Figs 1-8 , 10- 11 = 2mm; Fig. 9 = 0.5mm. Figs 1- 2. Parafusulina sp., Early Permian, N.E. Marathon, Texas, UCL coll. Figs 3, 5, 8, 10. Quasifusulina sp., Permian, Thitsipin Limestone Formation, South Shan State Burma, NHM OGS coll., 1970/ 24. Figs 4, 6. Schwagerina sp., 4, 6) Permian, Thitsipin Limestone Formation, South Shan State Burma, NHM OGS coll. (where wrongly identified as Quasifulina), 4) 1970/2 4; 6) 1970/2 9. Fig. 7. Praeskinerella sp., Early Permian, USSR, UCL coll. Fig. 9. Neostaffella sp., Moscovian, Khlong Pha Saeng, Thai Peninsula, Thailand. Map. 4737 I. 808879, AHG Mitchell NHM OGS coll., 231. Fig. 11. Dunbarinella ervinensis Thompson, Camp Creek foraminifera, on Saddle Creek, 9m south of Rockwood Texas, UCL coll. 1 2 1 2 3 44 A B 5 6 7 Plate 2.17 Scale bars: Figs 1- 7 = 2mm. Fig. 1 Schubertella sp., Late Carboniferous, Yorkshire, England, UCL coll. Fig. 2. Beedeina sp., Late Carboniferous, Yorkshire, England, UCL coll. Fig. 3. Polydiexodina praecursor Lloyd, holotype, Geli Khana Section, North Kurdistan, Permian, NHM, P44395. Figs 4, 6.  Paraschwagerina sp., Am 88b, Permian, Thitsipin Limestone Formation, 477 484, South Shan States Burma, NHM OGS coll., 1970/ 24. Fig. 5. A) Schwagerina adamsi Ross, Early Permian, oblique axial sec- tions, B) Pseudoschwagerina cf. P. fusiformis (Krotow), Early Permian, an equatorial section of a hypotype from Geli Khana Section, Zinnar Formation, North Kurdistan, Iraq NHM, P44410. Fig.7. Polydiexodina praecursor Lloyd, Permian (Wordian), axial section of paratype, Kaista Section, Zinnar Formation, North Kurdistan, NHM P44386. 1 2 3 4 5 6 7 8 9 10 11 12 13 Plate 2.18 Scale bars: Figs 1- 13 = 2mm. Fig. 1. Robustoschwagerina geyeri (Kahler and Kahler), holotype, axial section, Carnic Alps, SFN coll. Fig. 2. Chalaroschwagerina stachei (Kahler and Kahler), identified as Paraschwagerina stachei Kahler and Kahler, Paratype, oblique axial section, Carnic Alps, SFN coll. Fig. 3. Pseudoschwagerina nitida Kahler and Kahler, holotype, equatorial section, Carnic Alps, SFN coll. Fig. 4. Pseudoschwagerina aequalis Kahler and Kahler, holotype, Carnic Alps, equatorial section, SFN coll. Fig. 5. Occidentoschwagerina alpina Kahler and Kahler, paratype, early part of the UPL, ZK, Carnic Alps, SFN coll. Fig. 6. Sphaeroschwagerina pulchra (Kahler and Kahler), identified wrongly as Pseudoschwagerina pul- chra Kahler and Kahler, holotype, Carnic Alps, SFN coll. Fig. 7. Sphaeroschwagerina carniolica (Kahler and Kahler), paratype, axial section, Carnic Alps, SFN coll. Fig. 8. Pseudoschwagerina elegans Kahler and Khaler, paratype, equatorial section, Carnic Alps, SFN coll. Fig. 9. Sphaeroschwagerina citriformis (Kahler and Kahler), paratype, axial section, Carnic Alps, SFN coll. Fig. 10. Paraschwagerina lata Kahler and Kahler, paratype, Col Mezzodi Formation, (Forni Avoltri), SFN coll. Fig. 11 .Pseudoschwagerina lata Kahler and Kahler, holotype, equatorial section, Carnic Alps, SFN coll. Fig. 12. Robustoschwagerina tumida (Likharev), wrongly identified as Pseudoschwagerina schellwieni Hanzawa, Kahler and Kahler, axial section, Carnic Alps, SFN coll. Fig. 13. Pseudofusulina sp. Kahler and Kahler, equatorial section, Carnic Alps, SFN coll. 1 2 Plate 2.19 Scale bars:  Figs 1-2   =  1mm. Fig.  1. Assemblages of schwagerinids sp., Wolfcamp Beds, Wolfcamp Hills, 15 m. N. E. Marathon, Texas, UCL coll. Fig. 2. Assemblages of Fusulina cylindrica Fischer de Walheim, Limestone Elmdale Formation, Upper Gzhelian, Oklahoma, USA, UCL coll. 1 2 3 4 151 1692 Plate 2.20 Scale bars: Figs 1-6  = 2mm. Fig. 1. Neoschwagerina sp., late Permian, Wolfcamp Formation, Wolfcamp Bed, 15m N.E. Marathon, Texas, UCL coll. Figs 2-3 . Quasifusulina sp., Permian, Thitsipin Limestone Formation, South Shan State, Burma, NHM OGS coll., 2) 1970/2 4; 3) 1970/2 9. Fig. 4. Schwagerina sp., equatorial section, NHM coll. (where it is identified wrongly as Neoschwagerina sp.). Fig. 5. Schwagerina sp., Late Carboniferous Limestone, USSR, oblique equatorial section, UCL coll. Fig.  6. Triticites sp., Guadalupian, USSR, oblique axial section, UCL coll. 1 2 3 64 5 6 7 8 9 Plate 2.21 Scale bars: Figs 1-9  = 2mm. Figs 1- 2, 4- 6, 9. Enlargement of tangential sections of solid speci- mens of fusulinids displaying septal fluting, Italy, UCL coll. Figs 3, 8. Fusulinid specimens showing deposits of levee- like chomata around tunnel, USSR, UCL coll. Fig. 7. Enlargement of a fusulinid test showing the development of the tunnel and futures giving fold, UCL coll. 1 2 3 4 5 6 7 8 Plate 2.22 Scale bars: Figs 1- 8 = 2mm. Figs 1- 3. Enlargement of tangential sections of solid specimens of fusulinids displaying septal fluting, Italy, UCL coll. Fig. 4. Fusulinid specimen showing deposits of levee- like chomata around tunnel, USSR, UCL coll. Figs 5, 6. Fusulinid specimens showing antetheca and septal plications, Italy, UCL coll. Fig. 7. Fusulinid specimen showing the equatorial section with septa and spiral theca, UCL coll. Fig. 8. Solid specimen of Schwagerina sp. UCL coll. Plate 2.23 Scale bars:  Figs 1- 8  =  2mm. Fig.  1. Neoschwagerina aff. craticulifera (Schwager), Permian, Tebaga, S Tunisia, NHM P43989. Fig.  2. Neoschwagerina sp., Limestone from Bukit Kepayang Quarry, Pahang, NHM P42182. Figs 3- 4. Schwagerina adamsi Ross, Early Permian, 3) holotype, NHM, P42647, 4)  paratype, NHM P42648, oblique axial sections showing the thick secondary deposits on the septa, Macusani, Peru, Permian. Fig. 5. Parafusulina sp., Permian, Kuzu Machi Tochigi Prefecture, Central Japan, UCL coll.. Fig. 6. Schwagerina sp., Permian, UCL coll. Fig. 7. Parafusulina kaerimizensis (Ozawa), Permian, Kuzu area, Central Japan, NHM coll. Fig. 8 Dunbarinella ervinensis Thompson, Camp Creek, on Saddle Creek, 9 m south of Rockwood, Texas, UCL coll. A 1 2 3 6 4 5 6 7 8 1 2 3 4 5 6 Plate 2.24 Scale bars: Figs 1, 2 = 1mm; Figs 3, 4- 6 = 2mm. Figs 1- 2. Yabeina globosa (Yabe), Georgia, Grimsdale coll. Figs 3, 5. Pseudoschwagerina sp., Guadalupian, USSR, oblique axial section, NHM coll. Fig. 4. Dunbarinella ervinensis Thompson, Camp Creek, on Saddle Creek, 9m south of Rockwood, Texas, UCL coll. Fig.  6. Schwagerina sp., Late Carboniferous Limestone, USSR, oblique equatorial section, UCL coll. 1 2 3 4 5 6 Plate 2.25 Scale bars: Figs 1- 6 = 2mm. Fig. 1, 3-4 . Schwagerina sp., Permian, Pontafel Austria, Stürtz, iden- tified wrongly as Fusulina sp., NHM coll., P5111. Fig. 2. Zellia heritschi mira Kahler and Kahler, holotype, EarlyPermian, Carnic Alps, SFN coll. Fig. 5- 6. Triticites patulus Dunbar and Newell, Permian, Copacabana Group (bed at top of hill), Nacusani, Peru, NHM, P412656- P42659. 1 2 3 4 5 6 7 Plate 2.26 Scale bars: Figs 1-7  = 2mm. Figs 1- 2. SEM photographs of a fusulinid test, 1) showing antetheca and septal placations, 2) enlargement of the septal fluting. Figs 3-4 . Fusulinid specimens showing antetheca and septal placations, Italy, UCL coll. Fig. 5- 7. 5) solid specimen of Neoschwagerina sp., 6- 7) tangential thin sections of the same specimens. 1 2 2 3 34 5 5 6 7 Plate 2.27 (Algae, Corals, Bryozoa) Scale bars: Figs 1-7  = 0.5mm. Fig.  1. Calcifolium okense Shvetzov and Birina, Northumberland, UCL coll., JER 110. Fig. 2. Oligoporella sp., Hardraw Limestone, Yorkshire, England, UCL coll. Fig. 3. Chaetetes depressus (Fleming), corals, Hardraw Limestone, Yorkshire, England, UCL coll. Figs 4- 5. Bryozoa sp., Gayle Limestone, Yorkshire, UCL coll. Fig. 6. Nanopora anglica Wood, Fossdale Gill, Yorkshire, UCL coll., DH146. Fig. 7. Calcifolium sp., Swaledale, Yorkshire, UCL coll., DH714. 161 Chapter 3 The Mesozoic Larger Benthic Foraminifera: The Triassic 3.1 Introduction As seen in the previous chapter, the end of the Palaeozoic saw one of the most significant events in the history of life on Earth, with two mass extinctions occurring within a period of 10 Ma of each other (Kaiho et al., 2001; Chen and Benton, 2012). As a result of these events, about 90% of calcareous foraminiferal genera became extinct. The most affected were the large Fusulinida, which were wiped out. The only survivors of the Palaeozoic were a single member of each of the Endothyroidea and Earlandioidea superfamilies, which at that time were morphologically minute. In contrast, the extinction event had less impact on, for example, the simple Textulariida, which lost only 30% of its genera (Loeblich and Tappan, 1988). Forms from the Allogromiida, Miliolida and Lagenida also survived the end Permian extinction, albeit with significantly reduced diversity and as morphologically small forms. The Involutinida with aragonitic tests made their first appearance in the Triassic (Olenekian), persisting to the Early Cretaceous (Cenomanian). In comparison with Permian larger benthic foraminifera, the Triassic larger for- aminifera have not been systematically studied on a global scale. A  revision of the taxonomy of the Early and Middle Triassic taxa was presented by Rettori (1995), Rigaud et al. (2015), and a stratigraphic summary of larger benthic foraminifera of the Tethyan realm was presented by Pybernes (in De Gracianski et al., 1998). The relation- ship between the microgranular Paleozoic and agglutinated textulariides was explored by Rigaud et al. (2015). In this chapter, the taxonomy of the main genera of the Triassic larger foramin- ifera is presented, and a number of revisions suggested. Although most of the Triassic foraminifera are relatively morphologically small they have complex internal structures that are distinguishable in thin section, and so for the purposes of this study they are considered as “larger foraminifera”. Most of the superfamilies and families found in the Mesozoic are long ranging, but this chapter is only concerned with the genera char- acteristic of the Triassic. 3.2 Morphology and Taxonomy of Triassic Larger Benthic Foraminifera The Triassic larger forms are found developed in six orders: • Textulariida • Lagenida • Earlandiida 162 Evolution and Geological Significance of Larger Benthic Foraminifera • Endothyrida • Miliolida • Involutinida The development and evolution of the superfamilies of these orders is schematically shown in Fig. 3.1. Below, are presented the morphological characteristics and taxo- nomic relationships of the major Triassic forms, while in the next section their biostrat- igraphic significance and their phylogenetic relations are discussed. ORDER TEXTULARIIDA Delage and Hérouard, 1896 The tests of these agglutinated foraminifera are made of foreign particles bound by organic cement. They range from Early Cambrian to Holocene. Superfamily AMMODISCOIDEA Reuss, 1862 The test is sub-s pherical or tubular, with an aperture at the end of a tube. Cambrian to Holocene Age Period Ma 66.0 LAGENIDA INVOLUTINIDA 145.0 201.3 EARLANDIID 251.9 MILIOLIDA 298.9 358.9 ENDOTHYRIDA 419.2 PARATHURAMMINIDA 443.8 TEXTULARIIDA 485.4 ALLOGROMIIDA Fig. 3.1. The evolution of the Triassic larger benthic foraminifer orders and superfamilies. Cambrian Ordovician Silurian Devonian Carboniferous Permian Triassic Jurassic Cretaceous The Mesozoic Larger Benthic Foraminifera: The Triassic 163 Family Ammodiscidae Reuss, 1862 Members of this family have a proloculus that is followed by an uncoiled non-s eptate tubular second chamber. Early Cambrian to Holocene. • Gandinella Ciarapica and Zaninetti, 1985 (Type species:  Gandinella apenninica Ciarapica and Zaninetti, 1985). The test is partly streptospiral. Early Triassic to Late Triassic (Olenekian to Rhaetian) (Plate 3.1, fig. 8). • Glomospira Rzehak, 1885 (Type species: Trochammina squamata Jones and Parker var. gordialis Jones and Parker, 1860). The first chamber is followed by a strepto- spirally irregularly coiled chamber. The aperture is terminal. Early Carboniferous (Visean) to Holocene (Plate 3.2, figs 1- 3, 5- 9). • Glomospirella Plummer, 1945 (Type species:  Glomospira umbilicata Cushman and Waters, 1927). Similar to Glomospira but later becoming planispirally coiled. Late Carboniferous (Bashkirian) to Miocene (Plate 3.1, fig.  1; Plate 3.4, fig. 7; Plate 3.5, fig. 15; Plate 3.6, figs 6-7: Plate 3.7, figs 6-7.). • Paulbronnimannia Rettori and Zaninetti, 1993 (Type species: Agathammina judica- riensis Premoli Silva, 1971). The test is fusiform and compressed, composed of a globular proloculus followed by a long non-s eptate tubular chamber. Middle Triassic (Anisian) (Plate 3.5, fig. 12). • Pilammina Pantic, 1965 (Type species: Pilammina densa. Pantić, 1965). The spheri- cal proloculus is followed by a narrow non- septate, elongate second chamber. Early Triassic to Middle Triassic (Induan to Anisian) (Plate 3.5, fig. 9). • Pilamminella Salaj, 1978 (Type species: Pilammina grandis Salaj, in Salaj et al., 1967). The initial coiling is the same as Pilammina, then the second chamber changes 90o to the plane of coiling, followed by two or three oscillating coils. Middle to Late Triassic (Anisian to Carnian). Superfamily ATAXOPHRAGMIOIDEA Schwager, 1877 Members of this superfamily have a multilocular, trochospiral test becoming biserial or uniserial in later stages. Middle Triassic to Holocene. Family Ataxophragmiidae Schwager, 1877 Members of this family have three or more chambers per whorl in the early stages. The aperture is high and terminal. Late Triassic to Palaeocene. • Palaeolituonella Bérczi- Makk, 1981 (Type species: Palaeolituonella majzoni Berczi- Makk, 1981). The elongate conical test, having initial whorls with four to five cham- bers, is followed by a biserial stage and uniserial stages with internal rudimentary radiations. Middle to Late Triassic (Anisian to Carnian). Superfamily LOFTUSIOIDEA Bradey, 1884 The test is planispiral, may uncoil in later stage. The wall is agglutinated with differenti- ated outer layer and inner alveolar layer. Late Triassic (Carnian) to Holocene. Family Mesoendothyridae Voloshinova in Bykova et al., 1958 The test is strepto- or planispirally coiled, has involute initial chambers, and later is uncoiled. Adult chambers are cylindrical or flattened, falciform to cyclical. They are 164 Evolution and Geological Significance of Larger Benthic Foraminifera simple, with radial partitions or with pillars. Walls may have alveoles or a hypodermic network. Late Triassic (?Carnian to early Norian). • Wernlina Rigaud et al., 2014 (Type species: Wernlina reidae Rigaud et al., 2014). The test is symmetrical and planispirally coiled. The wall is dark, thick and micro- granular, formed by an inner alveolar layer sealed by an outer imperforate layer. The septa are thick and non- alveolar. The aperture is single and basal. Rigaud et al (2014) included in Wernlina, the species Everticyclammina praevirgulina described and illustrated from the Sinemurian– early Pliensbachian of Spain by BouDagher- Fadel and Bosence (2007, pl. 3, fig.  6). However, the latter differs from typical Wernlina in having an initial streptospiral coiling, fewer and proportionally thicker chamber walls and septa (see Chapter  4). Late Triassic (?late Carnian to early Norian). Superfamily LITUOLOIDEA de Blainville, 1825 Members of this superfamily have a conical, multilocular, rectilinear and uniserial test. The early stage has plani- (strepto- ) or trochospiral coiling. The periphery of the cham- bers has radial partitions; but centrally they are with or without scattered, separated pillars. The septa are arched into hummocks (almost solid masses) between the aper- tures, with bases of the arches that can fuse to the hummocks of the previous septum, with the apertures then opening at the suture. The alignment of the apertures and thickening of the hummock walls produces the appearance of a series of “gutters”. No true pillars are formed. The walls are solid, non-a lveolar, non- canaliculate. The aper- ture is simple, with no internal tooth plates, areal or multiple, cribrate. Late Triassic (Carnian) to Holocene. Family Lituolidae de Blainville, 1827 The early stages of the tests are enrolled, but later they may become rectilinear. Walls are formed from agglutinated foreign particles. There are few chambers (less than 10) per whorl. Carboniferous to Holocene. Subfamily Ammomarginulininae Podobina, 1978 The early stage of the test is coiled, but it becomes uncoiled in later stages. Apertures are single. Carboniferous (Early Mississippian) to Holocene. • Ammobaculites Cushman, 1910 (Type species: Spirolina agglutinans d’Orbigny, 1846). The test is simple, not compressed and uncoils in the adult. Apertures are sin- gle, areal. Carboniferous (Mississippian) to Holocene (Plate 5.6, fig. 18). Subfamily Lituolinae de Blainville, 1827 Members differs from Ammomarginulininae in having a multiple apertures. Late Triassic to Holocene. • Lituola Lamarck, 1804 (Type species: Lituolites nautiloidea Lamarck, 1804). These forms have no internal partitions and a multiple cribrate aperture. Late Triassic to Holocene (Plate 5.5, fig. 14; Plate 5.6, fig. 9). The Mesozoic Larger Benthic Foraminifera: The Triassic 165 Superfamily COSCINOPHRAGMATOIDEA Thalmann, 1951 Members of this superfamily are attached and may be coiled in their early stages, but later are uncoiled or branched. Triassic to Holocene. Family Coscinophragmatidae Thalmann, 1951 The wall is canaliculated, and perforated with alveoles. Late Triassic to Holocene. • Alpinophragmium Flügel, 1967 (Type species: Alpinophragmium perforatum Flügel, 1967). The chambers are numerous. The test is uniserial with a terminal cribrate aperture. Late Triassic (Carnian to Rhaetian) (Plate 3.6, figs 1-5 ) Superfamily VERNEUILINOIDEA Cushman, 1911 Representatives of this superfamily have a trochospiral test throughout, or only in the early stage. They may be triserial, biserial or uniserial. Some forms have a streptospiral initial part. The aperture is single or multiple. Late Carboniferous to Holocene. Family Piallinidae Rettori, Zaninetti in Rettori et al., 1993 The test is multilocular and elongated, with a proloculus followed by a trochospiral development, or with a small streptospiral in the early stage followed by a high trocho- spire in the later stage. Late Triassic (Carnian). • Piallina Rettori, Zaninetti in Rettori et  al., 1993 (Type species:  Piallina tethydis Rettori and Zaninetti, 1993). The elongated test, with a globular proloculus, is fol- lowed by a trochospiral or initially streptospiral part. Late Triassic (Carnian) (Plate 3.5, fig. 8). ORDER LAGENIDA Delage and Hérouard, 1896 This order is characterised by having monolamellar walls, composed of low-M g cal- cite in which the optical c- axes of the crystal units are perpendicular to the outer sur- face of the test. Primitive taxa are without secondary lamination, but more advanced forms are found with secondary lamination and a thin microgranular inner layer (see special terminology defined in Chapter  2). They range from Late Silurian to Holocene. Superfamily DUOSTOMINOIDEA Brotzen, 1963 The test is enrolled, planispiral to trochospiral. The aperture is single or double, and interiomarginal. Triassic (Anisian) to Early Jurassic (Hettangian). Family Duostominidae Brotzen, 1963 Characterised by two interiomarginal apertures in the final chamber. Triassic (Anisian to Rhaetian). • Duostomina Kristan- Tollmann, 1960 (Type species: Duostomina biconvexa Kristan- Tollmann, 1960). The test is lenticular, and trochospirally coiled. Triassic (Anisian to Rhaetian) (Plate 3.2, fig. 4). 166 Evolution and Geological Significance of Larger Benthic Foraminifera Earlandiida Lagenida Period and Stage Fig. 3.2. The phylogenetic development of the lagenides through the Palaeozoic and Triassic (sections of some lagenides are modified from Sellier de Civrieux and Dessauvagie, 1965). Age (Ma) Carboniferous Permian Triassic Bashkirian Moscovian Kasimovian Gzhelian Cisuralian Guadalupian Lopingian Induan Olenekian Anisian Ladinian Carnian Norian Rhaetian Earlandia Syzriana Tezaquina Protonodosaria Nodosinelloides Geinitzina Pachyphloia Cryptoseptida Austrocolomia Grillina The Mesozoic Larger Benthic Foraminifera: The Triassic 167 Family Variostomatidae Kristan- Tollmann, 1963, emend. Rigaud, Martini and Vachard, 2015 • Cassianopapillaria di Bari and Rettori, 1998 (nom. subst. for Papillaria di Bari and Rettori, 1996, preoccupied (Type species: Papillaria laghii, di Bari and Rettori, 1996). The test is trochospirally coiled, with pillars filling the umbilical region. Triassic (Carnian). Superfamily GEINITZINOIDEA Bozorgnia, 1973 The tests of this superfamily are uniserial, and similar to the Palaeozoic form Nodosinelloidea, but the microgranular layer is the inner dark layer, while the radially fibrous layer is the outer layer. Late Devonian to Middle Triassic. Family Abriolinidae Zaninetti and Rettori in Zaninetti et al., 1992 Morphologically small foraminifera, being trochospiral, with slightly curved sutures and subglobular chambers. The aperture is not observed. Middle Triassic (Anisian to Ladinian). • Abriolina Luperto, 1963 (Type species: Abriolina mediterranea Luperto, 1963). The test has a large sub- spherical proloculus. Middle Triassic (Anisian to Ladinian) (Plate 3.5, fig. 10). Superfamily ROBULOIDOIDEA Reiss, 1963 The tests of members of this superfamily are those of a typical Lagenida, but without secondarily lamellar or with slight lamination in younger taxa. The aperture is primi- tive and cylindrical. Late Silurian to Early Cretaceous. Family Ichtyolariidae Loeblich and Tappan, 1986. The test is elongate and uniserial with a single layered wall, and may show some sec- ondary lamination. The aperture is simple and terminal. Early Permian to Early Cretaceous. • Austrocolomia Oberhauser, 1960 (Type species: Austrocolomia marschalli Oberhauser, 1960). The chambers are cylindrical, gradually enlarging and divided by straight sutures, perpendicular to the long axis of test. Triassic (Anisian to Rhaetian) (Figs 3.2; 3.3). • Cryptoseptida de Civrieux and Dessauvagie, 1965 (Type species: Cryptoseptida ana- toliensis de Civrieux and Dessauvagie, 1965). Loeblich and Tappan (1986) included Pachyphloides de Civrieux and Dessauvagie, 1965, with a question mark in the syn- onyms of Cryptoseptida, as the off- centred illustrated specimen of Pachyphloides might have strong lamination as in Cryptoseptida. Permian to Triassic (middle Sakmarian to Carnian) (Figs 3.2; 3.3). • Grillina Kristan- Tollmann, 1964 = Geinitzinita de Civrieux and Dessauvagie, 1965 (Type species: Geinitzinita oberhauseri Sellier de Civrieux and Dessauvagie 1965). The tricarinate, flattened or biconcave cross- sections are without internal protective lamellae. Triassic (Carnian to Rhaetian) (Figs 3.2; 3.3; Plate 3.5, fig. 6). 168 Evolution and Geological Significance of Larger Benthic Foraminifera Austrocolomia Grillina (Carnian to Rhaetian) (Anisian - Rhaetian) Cryptoseptida (Late Permian - Carnian) Pachyphloia (Permian) Geinitzina (Permian) Fig. 3.3. A schematic figure showing the convergence of lagenides features in the Late Permian and the Late Triassic forms. Family Robuloididae Reiss, 1963 The test is uniserial, enrolled with an atelo-m onolamellar wall construction. The aper- ture is terminal. Middle Permian to Late Jurassic. • Robuloides Reichel, 1946 (Type species: Robuloides lens Reichel, 1946). The test is nautiloid to lenticular with an acute periphery. It is involute and planispirally, or nearly planispirally, coiled. A primary monolamellar wall covers previous cham- bers (weakly to markedly plesio-m onolamellar), so that the umbilical region of the test appears as a nearly solid mass of calcite and the peripheral septa are oblique. The aperture is simple. Middle Permian to Middle Triassic (Ladinian) (Plate 3.5, fig. 11). ORDER EARLANDIIDA SABIROV IN VDOVENKO ET AL., 1993 EMEND. VACHARD ET AL., 2010 This order has simple, free or attached tests, consisting of a proloculus and a rec- tilinear second chamber. The aperture is terminal and simple. Late Silurian to Early Cretaceous. The Mesozoic Larger Benthic Foraminifera: The Triassic 169 Superfamily EARLANDIOIDEA Cummings, 1955 This superfamily is characterised by having a free, non-s eptate test with a globular first chamber followed by a straight tubular one. It ranges from the Late Silurian to Early Triassic. Family Earlandiidae Cummings, 1955 The Earlandiidae have a single free chamber and range from the Late Silurian to Early Triassic. • Earlandia Plummer, 1930 (Type species: Earlandia perparva Plummer, 1930). The test is free, elongate, and composed of a globular proloculus followed by an undivided straight tubular chamber. The wall is calcareous and microgranular. Late Silurian to Early Triassic (Induan) (Plate 3.3, fig. 2, Fig. 3.2). ORDER ENDOTHYRIDA FURSENKO 1958 Members of this order have a lenticular, planispirally coiled test. The wall is dark and microgranular, but sometimes can be bilayered or multilayered. Apertures are simple, basal or cribrate. Late Devonian to Triassic. Superfamily ENDOTHYROIDEA Brady, 1884 nom. translat. Fursenko, 1958 Members of this superfamily have a streptospiral to planispiral tests with many chambers, followed by a rectilinear stage, which is biserial or uniserial in some forms. The wall is microgranular, calcareous, but some forms show two to three distinct layers; others may develop an inner perforate or keriothecal layer. Late Devonian to Triassic. Family Endotebidae Vachard, Martini, Rettori and Zaninetti, 1994 The test is free, and planispiral in early stages, but later uniserial to biserial. Walls are calcareous, grey, thick, and calcareous agglutinated. Apertures are simple. Late Permian to Triassic. • Endoteba Vachard and Razgallah, 1988 emend. Vachard, Martini, Rettori and Zaninetti, 1994 (Type species:  Endoteba controversa Vachard and Razgallah, 1988 emend. Vachard, Martini, Rettori and Zaninetti, 1994). The axial view is compressed. Late Permian to Late Triassic (Late Kungurian to Rhaetian) (Fig. 3.4). • Endotebanella Vachard, Martini, Rettori and Zaninetti, 1994 (Type spe- cies:  Endothyranella lwcaeliensis Dager, 1978). Similar to Endoteba, but the final stage is uniserial. Middle Triassic (latest Olenekian to Ladinian) (Fig. 3.4). Family Endotriadidae Vachard, Martini, Rettori and Zaninetti, 1994 Tests are trochospiral, almost planispiral, and may be uniserial in later stages with a microgranular wall. Apertures are basal, simple. Middle to Late Triassic. 170 Evolution and Geological Significance of Larger Benthic Foraminifera Endotriada Anisian – Norian Endotriadella Anisian – Norian Endotebanella Latest Olenekian - Carnian Endoteba Late Permian to Late Triasic Endothyra Late Devonian? or Early Carboniferous to Permian Earlandia Late Silurian - Early Triassic Fig. 3.4. A schematic evolution of the Endotebidae (drawings taken from Vachard et al. 1994) from the Endothyridae (Earlandia). • Endotriada Vachard, Martini, Rettori and Zaninetti, 1994 (Type species: Endotriada tyrrhenica Vachard, Martini, Rettori and Zaninetti, 1994). The chambers are hemi- spherical. Triassic (Anisian to Norian) (Fig. 3.4). • Endotriadella Vachard, Martini, Rettori and Zaninetti, 1994 (Type spe- cies: Ammobaculites wirzi Koehn-Z aninetti, 1968). The test has uniserial later stages. Triassic (Anisian to Norian) (Fig. 3.4). ORDER MILIOLIDA Delage and Hérouard, 1896 The miliolides have tests that are porcelaneous and imperforate, made of high Mg- calcite with fine, randomly oriented crystals. They range from the Carboniferous to the Holocene. Superfamily CORNUSPIROIDEA Schultze, 1854 The test is free or attached, and composed of a globular proloculus followed by a tubular enrolled chamber. The coiling is planispiral or trochospiral, evolute or involute, and may become irregular. Apertures are simple, at end of the tube. Early Carboniferous to Holocene. Family Arenovidalinidae Zaninetti, Rettori in Zaninetti, Rettori, He and Martini, 1991 Tests are lenticular with a globular proloculus and a second undivided chamber. Early to Middle Triassic. • Arenovidalina He, 1959 (Type species: Arenovidalina chialingehian- gensis He, 1959). The test is lenticular with a globular proloculus and a second tubular undivided The Mesozoic Larger Benthic Foraminifera: The Triassic 171 chamber. The coiling is planispiral throughout and involute, with a lamellar umbo- nal region on each side of the test. Early to Middle Triassic (Induan to Anisian, ?early Ladinian) (Fig. 3.5). • Paratriasina He, 1980 (Type species: Paratriasina jiangyouensis He, 1980). The test is planispiral, involute with a possibly irregular central area. Internal pillars may be present. The wall is porcelaneous. Early whorls show short zigzag bends as in Meandrospira Loeblich and Tappan, followed by later planispiral developments. Triassic (lattermost Anisian to early Ladinian) (Fig. 3.5). Family Ophthalmidiidae Wiesner, 1920 Tests are free, composed of a proloculus followed by an undivided coiled second cham- ber and chambers that commonly are one- half coil in length. The wall is porcelaneous and the aperture is terminal and simple. Triassic (Anisian) to Holocene. • Eoophthalmidium Langer, 1968 (Type species: Praeophthalmidium (Eoophthalmidium) tricki Langer, 1968). The coiling is involute. Middle Triassic (Anisian) (Plate 3.1, figs 3,5). • Karaburunia Langer, 1968 (Type species:  Karaburunia rendeli Langer, 1968). The proloculus is followed by sigmoidal coiling of two chambers per whorl, the final pair of chambers are approximately 180o apart. Middle Triassic (late Anisian) (Plate 3.1 figs 2,4). • Ophthalmidium Kübler and Zwingli, 1870 (Type species:  Oculina liasica Kübler and Zwingli, 1870). The coiling is planispiral and chambers have distinct floors and lateral extensions. Late Triassic (Carnian) to Late Jurassic (Kimmeridgian) (Fig. 3.5). Family Meandrospiridae Sadova, 1961 emend. Zaninetti et al., 1987 Tests are composed of a proloculus and an undivided second chamber with zigzag coil- ing. Apertures are simple and terminal. Permian to Holocene. • Meandrospira Loeblich and Tappan, 1964 (Type species: Meandrospira washitensis Loeblich and Tappan, 1946). The undivided second chamber forms an involute pla- nispiral zigzag of bends, with only those of the final whorl visible from the exterior. Early Permian (Artinskian) to Holocene (Plate 3.5, figs 1,2). • Meandrospiranella Salaj, 1969 (Type species: Meandrospiranella samueli Salaj, in Salaj et  al., 1967). The undivided tubular chamber coils streptospirally in short zigzag bends of about 5 whorls, later becoming somewhat irregular and uncoiling. Triassic (Anisian) to Late Cretaceous (Cenomanian) (Plate 3.5, fig. 3). • Turriglomina Zaninetti in Limongi, Panzanelli- Fratoni, Ciarapica, Cirilli, Martini, Salvini- Bonnard and Zaninetti, 1987 (Type species:  Turritellella meso- triasica Koehn- Zaninetti, 1968). The test is elongated and composed of a glob- ular proloculus and a second undivided tubular chamber with a first stage that is Meandrospira- like, followed by a long helicoidally compressed stage. Triassic (Anisian to Rhaetian) (Plate 3.4, fig. 9). 172 Evolution and Geological Significance of Larger Benthic Foraminifera Period, Epoch, Triassic miliolides Stage 201.3 ~227 Ophthalmidium ~237 ~242 Paratriasina 247.2 251.2 ? 251.9 Arenovidalina Shanita 259.1 Neohemigordius 272.9 Hemigordiopsis 298.9 Hemigordius 323.2 Fig. 3.5. The evolution of the Triassic miliolides from a Permian ancestor (modified from Zaninetti et al., 1991; and Rettori, 1995). Age(Ma) Pennsylvanian Cisuralian Guadalupian Lopingian TriassicEarly Middle Late Bashkirian Moscovian Kasimovian Gzhelian A S A K R. W. C. W. Ch. Induan Olenekian Anisian Ladinian Carnian Norian Rhaetian The Mesozoic Larger Benthic Foraminifera: The Triassic 173 Family Cornuspiridae Schultze, 1854 The tests are free or attached, composed of a proloculus followed by an undivided planispiral to streptospiral, involute or evolute second tubular chamber. Early Carboniferous (Visean) to Holocene. • Planiinvoluta Leischner, 1961 (Type species: Planiinvoluta carinata Leischner, 1961). The test is flat, attached, with tubular chambers, nonseptate, planispirally enrolled with an evolute coiling on the flat attached side and an involute coiling on the oppo- site side. Late Triassic (Rhaetian) (Plate 3.8, fig. 7). Family Hemigordiopsidae Nikitina, 1969 These forms are found with a test that has at least an early streptospiral stage, that later may be planispiral. Early Carboniferous (Visean) to Holocene. • Agathammina Neumayr, 1887 (Type species: Serpula pusilla Geinitz, in Geinitz and Gutbier, 1848). The proloculus is followed by an undivided, non- septate tubular chamber, coiling in five planes. Middle Carboniferous (Namurian) to Triassic (Plate 3.4, fig. 3). • Multidiscus. Miklukho- Maklay, 1953 (Type species:  Nummulostegina padangensis Lange, 1925 Family Hoynellidae Rettori, 1994 A free test, with a globular proloculus followed by an early miliolid stage and an undi- vided tubular chamber arranged on several vertical planes. Triassic to ?Early Jurassic. • Hoynella Rettori, 1994 (Type species: Hoynella sinensis Rettori, 1994). The test is ovoid, the tubular second chamber is followed by a planispiral evolute stage. Triassic to ?Early Jurassic (Plate 3.5, fig. 7). Superfamily SORITOIDEA Ehrenberg, 1839 These forms have chambers that are planispiral, uncoiling, flabelliform or cyclical, and may be subdivided by partitions or pillars. Late Permian to Holocene. Family Milioliporidae Brönnimann and Zaninetti, 1971 The tests are free or attached with a proloculus and tubular chambers arranged in var- ious planes of coiling, which may be irregular, oscillating, or sigmoidal. Late Permian to Late Triassic (Rhaetian). • Galeanella Kristan, 1958 (Type species: Galea tollmanni Kristan, 1957). The test is ovate in outline, planoconvex, planispiral to involute, and with thick coarsely perfo- rate walls. Late Triassic (Norian to Rhaetian) (Plate 3.4, fig. 11). • Kamurana Altiner and Zaninetti, 1977 (Type species: Kamurana bronnimanni Altiner and Zaninetti, 1977). The test is spherical in outline, with a globular proloculus fol- lowed by an undivided enrolled tubular second chamber. The aperture is terminal, and simple. Late Permian to Early Triassic (Wuchiapingian to Olenekian). 174 Evolution and Geological Significance of Larger Benthic Foraminifera ORDER INVOLUTINIDA HOHENEGGER AND PILLER, 1977 This order includes all forms with an enrolled second chamber. They have walls that are aragonitic, but commonly they are recrystallised to give a homogeneous micro- granular structure. They have an umbilical region with pillar-l ike structures on one or both sides of the test. They range from Triassic to Cretaceous. Superfamily INVOLUTINOIDEA Bütschi, 1880 These forms consist of a first chamber, followed by a planispiral to trochospiral enrolled tubular second chamber. Originally thought to date from Early Permian (Loeblich and Tappan, 1988; BouDagher, 2008), but here defined from Triassic to Late Cretaceous (Cenomanian). Family Triadodiscidae Zaninetti, 1984 The umbilical area is covered with additional lamellae added with each new whorl. Early Triassic to Late Triassic (Olenekian to Carnian). • Lamelliconus Piller, 1978 (Type species: Trocholina (Trocholina) biconvex Oberhauser, 1957). The test is lenticular with a globular proloculus, followed by a trochospiral undivided second tubular chamber. Triassic (Ladinian to Carnian) (Plate 3.5, fig. 13). • Triadodiscus Piller, 1983 (Type species:  Trocholina (Paratrocholina) eomesozoicus Oberhauser, 1957). The test is lenticular with a globular proloculus, followed by a planispiral to slightly trochospiral and involute second undivided tubular chamber that is planispiral. The whorls are formed by single lamella. Triassic (Olenekian to Norian/ ?Rhaetian) (Plate 3.7, fig. 2B; Plate 3.8, fig. 6). Family Aulotortidae Zaninetti, 1984 These forms have a lenticular to conical test with calcareous imperforate walls, and consist of a globular proloculus followed by a tubular, enrolled and undivided second chamber, lying against the previous whorl to form planispiral, oscillating or trocho- spiral coiling. Each half whorl is followed by the lamellar deposition of one or two layers. The aperture is at the open end of the tubular chamber. Triassic (Anisian) to Middle Jurassic. • Angulodiscus Kristan, 1957 (Type species: Angulodiscus communis Kristan, 1957). The second chamber is involutedly coiled. In axial section, the chamber lumen is curved and slightly tapering towards the umbilical area. Triassic (Norian to Rhaetian) (Fig. 3.6). • Auloconus Piller, 1978 (Type species: Trocholina permodiscoides Oberhauser, 1964). The second chamber is trochospirally enrolled and tubular. Each half whorl forms a lumina that covers the umbilicus, resulting in the build-u p of a thick and solid umbil- ical filling. Triassic (Norian to Rhaetian) (Plate 3.7, figs 1,6). • Aulotortus Weynschenk, 1956 (Type species:  Trocholina (Paratrocholina) oscillens Oberhauser. 1957). The enrolled second chamber is planispiral to slightly strepto- spiral, oscillating around the proloculus. Aulotortus has less regular planispiral coil- ing than Angulodiscus. Triassic (Anisian) to Middle Jurassic (Fig. 3.6; Plate 3.3, figs 3,5; Plate 3.5, figs 4,5; Plate 3.7, figs 2A, 9B, 11, 12B; Plate 3.8, figs 2, 3). The Mesozoic Larger Benthic Foraminifera: The Triassic 175 Family Triasinidae Loeblich and Tappan, 1986 The tests have a proloculus, followed by a broad tubular chamber. The interior is filled with cylindrical pillars. Triassic (Ladinian to Rhaetian). • Triasina Majzon, 1954 (Type species: Triasina hantkeni Majzon, 1954). Secondary thick- ening may occur in the umbilical area. Triassic (Norian to Rhaetian) (Plate 3.4, fig. 12). Family Involutinidae Bütschli, 1880 The globular proloculus is followed by a trochospiral coiled tubular second cham- ber. They show secondary lamellar thickenings on one or both umbilical regions. The aperture is at the open end of the tube. Members of this family may have shell mate- rial deposited on the umbilical area at one or both sides of the test. Triassic to Late Cretaceous (Olenekian to Cenomanian). • Aulosina Rigaud, Martini and Rettori, 2013 (Type species:  Triasina oberhauseri Koehn−Zaninetti and Brönnimann, 1968). The test is lenticular, with no inter- nal pillars, but with an evolute tubular chamber. The umbilical mass is wide. Late Triassic (Norian to Rhaetian). • Involutina Terquem, 1862 (Type species:  Involutina jonesi Terquem and Piette, in Terquem, 1862). The test is planispiral and evolute. Both umbilical regions are filled with lamellar deposits. Triassic to Late Cretaceous (Norian to Cenomanian) Fig. 3.8; (Plate 3.7, figs 3- 5, 7, 8B, 9A, 10, 12A). • Parvalamella Rigaud, Martini and Rettori, 2012 (Type species:  Glomospirella friedli Kristan- Tollmann, 1962). The test is lenticular to globular, lacking any internal struc- tures. The small proloculus is directly followed by a ball-l ike, enrolled, undivided tubu- lar chamber. The simple aperture is terminal. Triassic (late Ladinian to Rhaetian). • Semiinvoluta Kristan, 1957 (Type species: Semiinvoluta clari Kristan, 1957). Only the flat umbilical side is filled with pillar-l ike deposits. Triassic (Norian to Rhaetian) (Fig. 3.6). • Trocholina Paalzow, 1922 (Type species:  Involutina conica Schlumberger, 1898). A conical test consisting of a globular proloculus followed by a trochospirally enrolled tubular second chamber with pillars filling the umbilical area. The aperture is at the end of a tubular chamber. Trocholina closely resembles the simple small benthic Conicospirillina Cushman, but differs from the latter in having a solid plug instead of an open umbilical cavity. Trocholina is also less involute than Conicospirillina. Triassic to Cretaceous (Norian to Cenomanian) (Fig. 3.6; Plate 3.1, figs 6- 7). 3.3 Biostratigraphy and Phylogenetic Evolution 3.3.1 General Biostratigraphy The Permian-T riassic transition represents a critical period in foraminiferal evolution- ary history. Without question, the end Permian extinction resulted in the greatest reduc- tion of biodiversity in Earth history, claiming roughly 85 percent of all genera then in existence (Erwin et al., 2002; Chen and Benton, 2012). The dominant Palaeozoic order, Fusulinida, was eliminated, while diversity within the Lagenida, Miliolida and 176 Evolution and Geological Significance of Larger Benthic Foraminifera Involutinida Period and Stage 201.3 208.5 ~227 ~237 ~242 247.2 251.2 251.9 Fig. 3.6. Evolutionary lineages of the main genera of the Involutinida. (drawings of foraminifera are not to scale). Textulariida declined by 30, 50, and 80 percent respectively. As discussed in Chap. 2, the true Involutinida appeared as new forms in the Triassic. The end Permian mass extinction, and the following survival and recovery intervals, resulted in a dramatic turnover in the taxonomic composition of calcareous benthic assemblages (Groves and Altiner, 2004). During the Early Triassic, shallow marine foraminifera were simple and rare, with the first recorded forms appearing in the east- ern Palaeo-T ethyan realm (Sweet et al. 1992; Márquez, 2005), resulting elsewhere in a taphonomically induced gap in the understanding of the transition of the Permian sur- vivors (Groves, 2000). They were initially dominated by small arenaceous forms (Tong and Shi, 2000). These are the “disaster forms” which are, as defined by Fischer and Arthur (1977), ‘opportunistic species in the sense of MacArthur (1955) and Levinton (1970)’, and characteristic of the survival phase following a mass extinction (Hallam and Wignall, 1997). The subsequent disappearance of such forms, along with the pro- liferation of the survivors and the reappearance of ‘Lazarus taxa’, marks the beginning Age (Ma) Triassic Induan Olenekian Anisian Ladinian Carnian Norian Rhaetian Triadodiscus Triasina Aulotortus Angulodiscus Involutina Semiinvoluta Trocholina The Mesozoic Larger Benthic Foraminifera: The Triassic 177 Period and Stage Fig. 3.7. Biostratigraphic ranges of the main Triassic genera. of the recovery phase in the Ladinian (Groves and Altiner, 2004). The recovered forms seem to be dominated by the miliolides and later by the involutides in the Eastern Tethys area. The larger benthic foraminifera in the Early and Middle Triassic evolved slowly, and it was not till the Ladinian stage that highly diversified communities, such as the Ammodiscoidea and the Endothyroidea thrived. Among the most evolved foramin- ifera along with these communities are the Involutinida, such as Lamelliconus which became diverse and common in the Norian to the Rhaetian. Towards the end of the Triassic a gradual decline of the larger benthic foraminifera groups occurred, which climaxed with an end Triassic mass extinction event (Tanner et al., 2004), where most remaining larger benthic foraminifera disappeared. The biostratigraphic range of the main Triassic forms are summarized in Fig. 3.7. 3.3.2 The Palaeozoic- Triassic Lagenides One of the most important groups which survived the Permian extinction was the lagenides. Early Triassic lagenides are direct descendants of Permian forms and Age (Ma) Triassic Induan Olenekian Anisian Ladinian Carnian Norian Rhaetian Glomospira Glomospirella Pilammina Gandinella Pilaminella Paulbronnimannia Palaeolituonella Piallina Cryptoseptida Austrocolomia Abriolina Grillina Endoteba Endotebanella Endotriada Endotriadella Triadodiscus Lamelliconus Paratriasina Angulodiscus Auloconus Aulotortus Triasina Involutina Coronipora Semiinvoluta Trocholina Meandrospira Arenovidalina Meandrospiranella Turriglomina Paratriasina Duostomina Papillaria 178 Evolution and Geological Significance of Larger Benthic Foraminifera originated with the family Syzraniidae in Moscovian time. The wall in the earliest spe- cies of Syzrania (a form of Robuloidoidea) is dominantly microgranular, with only an incipiently developed hyaline-r adial layer. This led authors such as Altiner and Savini (1997) to retain the Syzraniidae within the Fusulinida. Similarly, in certain reference books (e.g. Loeblich and Tappan, 1998)  and in previous analyses of foraminiferal diversity across the P-T boundary, a number of lagenides genera were also regarded as either belonging to the Fusulinida or Rotaliida. In our modification of the phyloge- netic of the Carboniferous forms, we would suggest that they may have indeed derived from the Earlandiida. However, more recent work (Groves et al., 2003, 2004) suggests that the order Lagenida is monophyletic. Authors, such as Palmieri (1983), Groves and Wahlman (1997), Pinard and Mamet (1998), Groves (2000) and Groves et al. (2004), suggested there were very close similarities in wall structures between Palaeozoic and Mesozoic- Cenozoic lagenides, with no change in wall structure at the Permian- Triassic boundary. If so, the Triassic Robuloidoidea would have had its origins in the Parathuramminida (see Fig 3.1). This is still an unresolved issue. Studies by Grove et al. (2004) and Grove (2005) divided the lagenides, at the high- est level, into nonseptate and septate taxa. Within the nonseptate group there are forms whose tubular second chamber is unpartitioned (e.g. Syzrania, see Chapter 2), and forms whose second chamber is partitioned, to varying degrees, by constrictions or thickenings of the wall (e.g. Tezaquina, (Fig.  3.2); see also Chapter  2). Within the septate group there are uniserial (e.g. Nodosinelloides, see Fig.  3.2) and coiled forms (Calvezina, see Chapter  2). The uniserial forms are further differentiated on the basis of chamber symmetry when seen in transverse section:  i.e., radial circular (Nodosinelloides); radial bilateral flattened (Geinitzina); triangular (Pseudotristix); and polygonal (Rectostipulina). The phylogenetic evolution of the lagenides, as shown in Fig. 3.2, follows a morphological trend varying from uniserial tubular, non-s eptate forms to uniserial septate forms with rounded transverse section which become flat- tened and flaring in advanced forms. The Moscovian- Kasimovian interval of the Carboniferous saw the initial evolution- ary radiation of the early complex lagenides with the oldest known genus Syzrania, which as indicated above might have evolved from the Earlandioidea (Earlandia) in the Middle Pennsylvanian (see Figs 3.1 and 3.2), through the addition of a hyaline-r adial layer external to the ancestral microgranular wall (Groves et al., 2004). This is shown in the earliest species of Syzrania where the wall was still dominantly microgranular with only an incipiently developed hyaline-r adial layer (Fig. 3.2). The Syzranidae evolved in the Late Pennsylvanian and Permian into sub-s eptate and fully septate uniserial forms. In the late Ghzelian, the ancestor to all laterally compressed early lagenides, Geinitzina, made its first appearance, but it was not until the late Sakmarian that the Pachyphloia – Howchinella group finally developed. The Late Permian (Lopingian) lagenides were diverse and abundant throughout the Tethyan region and northern higher paleolatitudes, filling the niches that became vacant after the larger fusulinides became extinct. Among these lagenides were the compressed forms Geinitzina, Pachyphloia, Nodosinelloides, Protonodosaria and Robuloides. However, Early Triassic (Induan and Olenekian) rocks contain very few lagenides. These Early Triassic lagenides are morphologically simple forms, so- called “disaster forms”. It was not until the late Anisian that true recovery really began, with the reappearance of The Mesozoic Larger Benthic Foraminifera: The Triassic 179 more evolved Permian-l ike features, such as the laterally compressed test and the subdivided chambers, with radiate or modified apertures and secondary lamellarity of the wall (e.g. Grillina, see Fig. 3.3). The reappearance of these Permian features in the later Triassic and younger taxa is attributed to evolutionary morphological convergence (Groves et al., 2004). Indeed, this convergence almost certainly accounts for the many examples of so called “Elvis taxa” (Erwin and Droser, 1993) among Palaeozoic and Mesozoic- Cenozoic forms. 3.3.3 The Triassic Earlandiida and Endothyrida As discussed in Chapter 2, the initial decline of the fusulinides coincided with the end Guadalupian in the Late Permian. This crisis event eliminated all large and morpho- logically complex forms of fusulinides, and by the end of the Permian all fusulinides were extinct. Indeed, only two genera (Earlandia, Endoteba), from the earlandiides and the endothyrides respectively, survived the end Permian mass extinction (Fig. 3.4), but they never experienced significant taxonomic recovery (Groves and Altiner, 2004). The Endotebidae have their origin in the Permian where they evolved from the true Endothyridae.The type species Endoteba controversa Vachard and Razgallah, 1988 is known from Late Permian rocks in Tunisia, Yugoslavia, Italy, Afghanistan and Japan. However, according to Vachard et al. (1993) the Triassic occurrence of this genus is known mainly from the Anisian to Ladinian. They are very rare in the Carnian and Rhaetian throughout the Tethyan realm, with the only documented Early Triassic (Olenekian) occurrence being that of E. ex gr. bithynica Vachard et al., 1993. Endoteba is regarded as a Lazarus taxon, as no known Induan occurrences have been documented to date. A slight increase in the number of genera characterized the Middle Triassic, there being five gen- era but with no more than ten species in all. In Tethys, Endoteba evolved into three genera or species- groups, Endotebanella (uppermost Olenekian through Carnian), Endotriada (Middle Triassic), and Endotriadella (Middle Triassic). The endothyrides became entirely extinct in the Late Triassic, while the earlandiides died out within the Early Triassic (see Fig. 3.4). Earlandia have been suggested as having survived into the Early Cretaceous, but this has not been confirmed (Arnaud- Vanneau, 1980; Vachard et al., 2010). 3.3.4 The Triassic Miliolides Foraminifera in the order Miliolida first made their appearance at the beginning of the Carboniferous (Ross and Ross, 1991). Analysis by Flakowski et al. (2005) of SSU (small sub- unit) rRNA genes showed that the miliolides cluster with the calcareous Spirillinidae and agglutinated Ammodiscidae in an independent, fast evolving line- age (Pawlowski et al., 2003). This suggests that miliolides evolved directly from some allogromiide-l ike ancestor, which is consistent with earlier morphological observa- tions (Arnold, 1978). This was subsequently questioned by Gaillot and Vachard (2007) and Vachard et al. (2010). The latter argued that the early attached miliolides derived probably from the attached Tournayellida and not from the Ammodiscidae, which are homeomorph of the Tournayellida. The question is still open, but in Fig 2.1 and 3.1 they are shown as having an allogromiide-l ike ancestor. 180 Evolution and Geological Significance of Larger Benthic Foraminifera Although the miliolides proliferated during Middle and Late Permian, the end Permian extinction nearly eliminated the order, as almost all species disappeared. However, a few minute disaster forms survived the extinction, and forms such as Cornuspira and Rectocornuspira (see Chapter 2) are found in the Induan, apparently because of their toler- ance of environmental stress (Grove and Altiner, 2004). Indeed, the middle Induan is zoned on the occurrence of R. kahlori by many authors, such as de Graciansky et al. (1998). Like other forms, the miliolides were rare in the Early Triassic. The appearance of Meandrospira in the Olenekian represented a probable continuation of the Permian form assigned to Streblospira (which is according to Loeblich and Tappan is a synonym to Meandrospira), and is considered by Groves and Altiner (2004) as a Lazarus taxon. The late Olenekian of Tethys is zoned by M. pusilla (de Graciansky et al., 1998), and it was not before the middle Anisian that Meandrospira expanded, and subsequently evolved into forms such as Meandrospiranella and Turriglomina in the late Anisian. The appearance of the latter coincided with the disappearance of the Meandrospira species group from Tethys at the end of the Anisian. A parallel evolution to that of Meandrospira (= Streblospira) also took place (accord- ing to Zaninetti et al., 1991, and Rettori, 1995) in the Neohemigordius – Arenovidalina lineage. Neohemigordius in the Early Permian loses its prominent umbilical thickening and acquires a streptospiral early coil as is in Hemigordius and its species group of the Permian. The Hemigordius group disappears in the Late Permian. However, similar forms to Hemigordius, but with planispiral tests and thickening in the umbilical region, as in Neohemigordius, and referred to as Arenovidalina, reappear in the Olenekian. Arenovidalina could be regarded as a Lazarus taxon that gave rise, just before disap- pearing in the early Ladinian, but giving rise to the Mesozoic group ophthalmidiids, by losing the lamellar umbonal region on each side of the test and developing a strep- tospiral early coil with much irregular later coiling. The tubular undivided planispi- ral second chamber of Arenovidalina becomes subdivided into distinct chambers in Ophthalmidium. Another short lineage also to develop from Arenovidalina in the late Anisian is that of Paratriasina, with a streptospiral early coil that later becomes plani- spiral. This phylogenetic evolution is shown schematically in Fig. 3.5. 3.3.5 The Triassic Involutinides The Early Triassic witnessed the first appearance of the aragonitic involutinides (older texts such as Loeblich and Tappan (1988) identify them as originating in the Permian, but this is based on the erroneous classification of the porcelaneous genera Neohemigordius (Fig. 3.5) and Pseudovidalina as involutinides rather than miliolides, see Chapter 2). The order Involutinida includes all forms with an enrolled second chamber and pillar- like structures in the umbilical region on one or both sides of the test (see Fig. 3.6). The oldest Triassic representative of the group is Triadodiscus, which first appeared abruptly in the Olenekian with no known occurrences in the Induan. The ancestor of Triadodiscus is still unknown. One of the common hypothesis is that involutinides had an Archaediscoidean ori- gin (Groves and Altiner, 2004). However, Archaediscoidean forms were confined to The Mesozoic Larger Benthic Foraminifera: The Triassic 181 the Palaeozoic and were never recorded in the Triassic. Gargouri and Vachard (1988) proposed the ammodiscoides Glomospirella as being a possible ancestor, but other authors favoured a hemigordiopsid Multidiscus ancestor (Altiner et al., 2005). However, Multidiscus is not definitely known from the earliest Triassic. Furthermore, rDNA anal- yses suggest a significant genetic and evolutionary separation between miliolids and other testate foraminifera (Pawloski, 2000), which would mitigate against this proposal. In the Anisian, the planispirally coiled Triadodiscus evolved into streptospirally coiled Aulotortus (see Fig. 3.6). The late Ladinian of Tethys is recognised (de Graciansky et al., 1998) by the appearance of A. praegaschei (Koehn-Zaninetti). However, it was not before the Norian that significant diversification occurred, with changes in the shape of the coiled tests and the shape and occurrences of pillars. Several evolutionary trends in the involutinides occurred from Middle to Late Triassic, which are biostrati- graphically useful. The Norian and Rhaetian are zoned using Triasina oberhauseri and T. hantkeni (Plate 3.4, fig. 12) respectively. The development of Involutina (see Figs 3.6;  3.8) has been debated by many authors. Kristan- Tollman (1963) proposed a direct development from the trochospiral Lamelliconus. However, this theory was refuted by Rigaud et al. (2013) who demon- strated that trochospirally coiled involutinides are a separate group. On the other hand, Involutina is planispiral- evolute, non- septate, perforate, and possess at least two lami- nae per whorl, that are, as in Aulotortus, interfingered in the median part of the umbil- ical region (Rigaud et al., 2015). It was also postulated that Involutina evolved from Triadodiscus (Piller, 1978; Gaździcki, 1983; di Bari and Laghi, 1994), however, the lam- ina depositions of Triadodiscus are more discontinuous than those of Involutina. In the latter the interfingering laminar arrangement, the type of coiling with possible irregu- larities and tubular chamber morphology are more closely phylogenetically linked to Aulotortus (Rigaud et al. 2015) (see Figs 3.6; 3.8). Most involutinides disappeared at the end of the Triassic and only three species continued into the Mesozoic. Canals A Laminar Perforations Megalosphericextensions proloculus B Merged canals Fig. 3.8. Sections of Involutina sp., A) Schematic axial section: B) Photomicrograph of an Axial section. 182 Evolution and Geological Significance of Larger Benthic Foraminifera 3.4 Palaeoecology of the Triassic Larger Foraminifera The Triassic saw the single vast super-c ontinent of Pangea straddling the equator, with the Tethys Ocean intruding into Pangea from the east, leading to the division between Laurasia in the North and Gondwana in the South. Land animals were free to migrate from Pole to Pole (Fig. 3.9). The interior of Pangea was hot and for the most part dry and very arid. Warm temperate climates extended to the poles. Sea levels were low and the seas had little or no dissolved oxygen and were possibly alkaline (Woods, 2005). Rapid global warming at the very end of the Permian expanded the deserts and cre- ated a “super- hothouse” world, that may have contributed to or exacerbated the great Permian- Triassic extinction. The carbon dioxide content of the atmosphere was very high and the level of atmospheric oxygen was unusually low (Berner et al., 2003), see Fig. 3.10 The end Permian extinction led to a profound decline in reef production and caused the near- total elimination of metazoan reef systems (Hallam and Wignall, 1997), giv- ing rise to a so-c alled “reef gap” (Fagerstrom, 1987; Flügel, 1994) during the Induan and early Olenekian. Recovery was very slow for foraminifera, and ecological recovery was delayed until the middle Olenekian. The start of the Triassic was characterised by extremely impoverished taxa, dominated by opportunists and generalists (Rodland and Bottjer, 2001). Many genera which seemed to have disappeared at the Palaeozoic- Mesozoic boundary remained hidden, possibly in geographically restricted environ- ments, before returning as “Lazarus” taxa in the Olenekian or Middle to Late Triassic. Loeblich and Tappan (1988) suggested that the morphologies of the surviving faunas are characteristic of infaunal detrital feeders, which would have been therefore less sus- ceptible to primary productivity crashes, in contrast to the larger benthic foraminifera which lived symbiotically in reefal environments. On the other hand, Chen and Benton Fig. 3.9. Palaeogeographic reconstruction of the Early Triassic (by R. Blakey, http://j an.ucc.nau.edu/ ~rcb7/ paleogeographic.html). The Mesozoic Larger Benthic Foraminifera: The Triassic 183 Age Period Atmosphere oxygen levels % Ma 0 10 20 30 66.0 145.0 Low level of oxygen inhibits the development of large forms. 201.3 251.9 High level of oxygen enables 298.9 large sizes to be achieved. 358.9 419.2 443.8 485.4 Fig. 3.10. Oxygen variation over time, highlighting the low level of oxygen during Triassic time (modified after Berner et al., 2003). (2012) postulated that recovery of complex life in the Triassic could have been delayed either by a complex multispecies interaction combined with physical perturbations. It was not before the Anisian that larger benthic foraminifera became more wide- spread and cosmopolitan. During this time, a tropical sea formed between the super- continents of Gondwana (in the south) and Laurasia (in the north). This sea, called the Tethys (Fig 3.9), became home to large diverse assemblages of foraminifera. However, the larger Triassic benthic foraminifera never recovered the relatively gigantic sizes seen in the Permian, but rather remained small and indistinctive. This was caused most probably by the drop in atmospheric and marine oxygen levels at the Permian- Triassic boundary (Berner et al., 2003), see Fig. 3.10. Increasing atmospheric oxygen levels have, for example, been implicated in the rise of late pre- Cambrian and Cambrian marine faunas (Cloud, 1976; Runnegar, 1982; McMenamin and McMenamin, 1990) and are consistent with the presence of Permian-C arboniferous giant insects. Similarly, anoxia has been regularly correlated with the dramatic Permian- Triassic extinction events in the ocean and the small size of the larger foraminifera during the Triassic (Erwin, 1993; Knoll et al., 1996; Wignall and Twitchett, 1996). In the oceans today, many benthic invertebrates compensate for hypoxic conditions via a broad spectrum of behavioural Cambrian Ordo. S. Dev. Carb. Pem. Trias. Jurassic Cretaceous 184 Evolution and Geological Significance of Larger Benthic Foraminifera and morphological adaptations (Rogers, 2000). Similarly, infaunal penetration by the foraminifera may be a response to deplete sediment oxygenation in the Triassic (Berner et al., 2003). During the Anisian, carbonate platform development reoccurred with the impor- tant contribution of reefal debris, which as noted above are almost totally absent in Early Triassic carbonate platforms. This was accompanied by the appearance of scle- ractinian corals and the increasing importance of dasycladoidean green algae (e.g. Griphoporella sp., Plate 3.1, fig. 9), the increasing frequency of crinoids and molluscs, and the scarcity of ooids (Kiessling et al., 2003). During the Ladinian there was a clear stabilization of marine ecosystems, accompanied by the development of wide carbo- nate platforms bearing highly diversified communities (Márquez and Trifonova, 2000; Márquez, 2005). Several foraminiferal genera and species that emerged and/o r developed during the Anisian and Ladinian gave rise to characteristic faunal associations that have enabled some authors to establish biozones (Salaj et al.,1983; Trifonova, 1984; Salaj et al., 1988; Márquez, 2005). Among the Anisian forms are Meandrospira dinarica Kochansky- Devide and Pantic (Pl. 3.5, fig.  2) in the Carpathian- Balkan zone and in the Alps (Rettori, 1995), and Pilammina densa Pantic (Pl. 3.5, fig.9) in the Pelsonian of the Alps (Zaninetti et al., 1991). Among the most frequently found forms in the Ladinian of western Tethys are numerous representatives of the Involutinida, such as Lamelliconus gr. biconvexus- ventroplanus (Oberhauser), L.  multispirus (Oberhauser), L.  procerus (Liebus), L.  cordevolicus (Oberhauser), Aulotortus praegaschei (Koehn- Zaninetti), A.  sinuosus (Eynschenk) (Pl. 3.7, figs 2A, 9B, 11)  and A.  pragsoides (Oberhauser) (Pl. 3.8, fig. 2; Trifonova, 1992, 1993, 1994; Salaj et al., 1983; Oravecz-S cheffer, 1987; Márquez, 2005). The species Pilamminella gemerica Salaj appears in abundance in the Ladinian characterizing the P. gemerica zone, and is characteristic of reef facies and dasycladoidean platform carbonate sediments (Salaj et al., 1988; Márquez, 2005). The climate in the Carnian was unusually arid and hot and the seaway itself appears to have experienced a “salinity crisis” (general evaporation) and a breakdown of the reef system. In the Carnian, foraminiferal assemblages were dominated by endothy- rides and miliolides . The Late Carnian- Norian interval saw major reef extinctions in the Tethys province, with the loss of older coral species, but diversity was main- tained with reciprocal replacement by new taxa (Stanley, 2001). These in turn became extinct, possibly a result of a sudden global cooling. This is mirrored by the disap- pearance of many foraminiferal genera in western Tethys, which were subsequently replaced by new forms of Involutinida in the Norian. The early Norian Tethyan reef systems were replaced by much larger reef developments during the Late Triassic (mid Norian to Rhaetian) coinciding with a major rise in global sea-l evel. At this time the new foraminiferal taxa of Involutinida became dominant and corals diversified, con- stituting a new reef building consortium. There was a period of world-w ide expansion of carbonate platforms and maximum reef diversity (Stanley, 2003). Associations of Late Triassic benthic foraminifera dominated by Ammodiscoidea, Involutinoidea, and Duostominoidea are mainly found in shallow marine facies, often restricted and eco- logically highly unstable (Márquez, 2005). They are generally typical of low energy, bay or lagoon-t ype, protected settings, with a salinity sometimes higher than normal, on shallow carbonate ramps. The different reef facies include forms common to the Tethys The Mesozoic Larger Benthic Foraminifera: The Triassic 185 such as Triadodiscus mesotriasica (Koehn- Zaninetti), Palaeolituonella meridionalis (Luperto), Endoteba wirzi (Koehn- Zaninetti), Duostomina alta Kristan-T ollmann and D. astrofimbriata Fuchs, to mention a few. In these reef environments, the foraminiferal fauna are more numerous and diverse than in the restricted carbonate ramps of the Triassic basins (Márquez, 2005). Many of the Late Triassic species have morphological characteristics, such as the irregularly distributed papillose lamellae and tube infolding of Involutina, which have been interpreted as being rudimentary features associated with photosymbiosis (Rigaud et al., 2015), typically developed in shallow, high-e nergy tropical carbonate platforms (Martini et al., 2009). 3.5 Palaeogeographic Distribution of the Triassic Larger Foraminifera The Early Triassic was characterised by small and low- diversity taxa, and is viewed as a time of delayed biotic recovery that persisted until the Middle Triassic (Hallam, 1991; Galfetti et al., 2008; Chen and Benton, 2012). Explanations for this vary from invok- ing the sheer scale of the Permian extinction to the prolonged stresses of the environ- ment post end- Permian extinction (Hallam, 1991; Payne et al., 2004; Chen and Benton, 2012), or the effects of further extinction crises in the Early Triassic (Orchard, 2007; Stanley, 2009; Song et al., 2011). The Early Triassic larger benthic recovery shows roughly the same trajectory eve- rywhere. The foraminifera were mainly small, and they followed the model of the “Lilliput effect” described by Urbanek (1993), where he refers to the pattern of size change through extinction events, specifically the temporary appearance of dwarfish organisms (Twitchett, 2006). There was a significant decrease in maximum and mean size among species and genera across the Permian/T riassic boundary. This was fol- lowed by gradual size increase through the late Induan and into Olenekian. The recovery of reef-b uilding larger benthic for- aminifera began 4 Ma later, in the Anisian, but they never recovered the pre- extinction sizes of the Permian (Payne et  al., 2011; Song et  al., 2011; Rego et  al., 2012). The anoxia and global warming that prevailed through the Permian–T riassic period may have contributed to this “Lilliput effect” on larger benthic foraminifera in the after- math of the end-P ermian crisis. The distribution of the Triassic genera is plotted in charts (see Figs 3.11 to 3.13) in order to understand their palaeogeographic distribution. The genera are divided into three provinces: Europe and Russia, China and East Asia, and North America. There is a general and progressive increase in diversity from the Induan to the Anisian, with the end Olenekian being marked by the increase in the number of orders that exhibit larger morphologies. A possible minor extinction event occurred at the end of the Anisian affecting the Ammodiscoidea in Europe. The ancestor of the Involutinida appeared in all the three provinces, and the order increased in diversity in China and East Asia during the Ladinian, but it was not before the Norian that they became wide- spread in Europe. The Endothyroidea were common during the Anisian and Ladinian in China, East Asia and Europe, but they decreased gradually toward the end of the Norian. However, they did not disappear completely from these provinces until the end of the Rhaetian. 186 Evolution and Geological Significance of Larger Benthic Foraminifera Triassic Total Genera 18 16 14 12 10 8 6 4 Europe & 2 Russia China & E Asia 0 N America Carnian Norian Rhaetian Ladinian Induan Olenekian Anisian Fig. 3.11. The number of larger benthic foraminifera genera found throughout the Triassic in the three main palaeogeographic provinces. Triassic Speciation 10 9 8 7 6 5 4 3 2 Europe & 1 Russia China & E Asia 0 N America Rhaetian Ladinian Carnian Norian Induan Olenekian Anisian Fig. 3.12. The number of new larger benthic foraminifera genera occurring throughout the Triassic. The Mesozoic Larger Benthic Foraminifera: The Triassic 187 Triassic Extinctions 12 10 8 6 4 2 Europe & Russia China & E Asia 0 N America Rhaetian Ladinian Carnian Norian Induan Olenekian Anisian Fig. 3.13. The number of larger benthic foraminifera genera extinctions occurring throughout the Triassic. From Fig. 3.11, it can be seen that throughout the Triassic, the diversity in North America was constantly low, but there was a distinct increase in the number of gen- era throughout Europe, and to a lesser extent in China and East Asia, in the Anisian and Norian. From Fig. 3.12, we see that there was a sharp decline in speciation in the Ladinian, and that no new forms appeared at all in the Rhaetian. However, during the Anisian and the Norian the larger benthic foraminifera speciation was at its most productive. North American foraminifera speciation mirrored generic diversity but remained poor throughout the Triassic. There was a gradual increase in the number of extinctions from the Ladinian to the Carnian in all three provinces, that was followed by a large extinction event at the end of the Rhaetian (Fig. 3.13). The trends described above in all three provinces, can be understood to some extent from an analysis of the palaeogeography of the Triassic. After the great extinction of the end Permian, foraminifera were struggling to survive during the Induan interval. The recovery process was very slow in the Induan, but new forms were appearing, 80 % of the Induan foraminifera were new Triassic forms with the first records of foramin- ifera appearing in the eastern Tethys area (China) and Eastern Europe. They are domi- nated by opportunist, euryfacial forms, giving rise to associations that are sometimes numerous but of low species diversity, since they are mostly comprised of arenaceous genera (Tong and Shi, 2000) from the Ammodiscoidea (see Fig. 3.7). During the Early Triassic, the diversity of the foraminifera increases gradually in all three provinces, 188 Evolution and Geological Significance of Larger Benthic Foraminifera North America having the least diversity. This is probably due to its relative isolation during the Early Triassic (see Fig. 3.9). In the eastern Tethys area, the recovery process seems to be dominated by forms of Miliolida (Márquez, 2005). Among the most common species are Cornuspira mahajeri (Brönnimann, Zaninetti and Bozorgnia) and Rectocornuspira kalhori (Brönnimann, Zaninetti and Bozorgnia), which occur widely in many areas of Europe and Asia (Rettori, 1995). Trifonova (1993) recorded these species from the Early Triassic of Bulgaria, along with species of the genus Kamurana bronnimanni (Altiner and Zaninetti) and K. chatalovi (Trifonova). During the Olenekian, Triassic foraminifera started rapidly to diversify, leading to the large number of Anisian species (Fig. 3.12). The late Olenekian is characterised by the presence of the species Meandrospira pusilla (Plate 3.5, fig. 1), widely described by many authors including Zaninetti (1976), Gaździcki et al. (1975), Salaj et al. (1983), Trifonova (1978, 1993), Rettori (1995) and Márquez (2005). This species is particularly abundant in the late Olenekian and characterizes the M. pusilla Zone of Salaj et al. (1988). The Anisian saw the highest number of new genera in the Triassic (see Fig. 3.12). However, at the end of the Anisian a small extinction occurred, mostly among European Ammodiscoidea (affecting 35% of the genera), and within the Ladinian, 60% of the Cornuspiroidea from all three provinces disappeared. This may have been associated with environmental stress caused by the large volcanic event (ca. 235 Ma) along the northwest margin of North America, and which formed the Wrangellia Terrane. After this event, the foraminifera in North America and Europe did not recover completely until the end of the Carnian. The Far East seems not to have been affected by this event. The Wrangellia volcanic sequences were oceanic events, and so may not have had the same globally devastating environmental and climatic effects inferred to be associated with sub-a erial volcanic eruptions (e.g. the Siberian Traps). At the end of the Ladinian, there was a general decline in speciation after the blooming of the Early Triassic recovery, except after the end Carnian event which saw the expansion in the Norian of the superfamily Involutinida in European Tethys. The ancestral form of this family appeared in the Anisian, but only attained importance in the Norian and Rhaetian. In addition, the first forms of Lamelliconus emerged in the Ladinian, a genus that was to develop substantially during the Norian. The miliolide genus Turriglomina, which developed during the Anisian, gave rise to many species of stratigraphic significance, such as T.  conica (He), T.  mesotriasica (Koehn-Z aninetti) and T.  scandonei (Zaninetti, Ciarapica, Martini, Salvini- Bonard and Rettori). The species Palaeolituonella meridionalis (Luperto), widely cited for the Middle Triassic from several Tethyan localities, is especially common in the Ladinian (Trifonova, 1992; Márquez, 2005). Early Carnian Triassic foraminifera had become cosmopolitan, as a small sea- way connected the Tethys Ocean with North America (see Fig.  3.14). Foraminifera thrived until the end of the Carnian. The high number of extinctions at the end of the Carnian coincide with the occurrence of set of multiple of impacts (Spray et al., 1998). Five terrestrial impact structures have been found to possess comparable ages The Mesozoic Larger Benthic Foraminifera: The Triassic 189 Fig. 3.14. Palaeogeographic reconstruction of the Late Triassic (by R. Blakey http://j an.ucc.nau.edu/~ rcb7/ paleogeographic.html). (214+/ - 2 Ma), close to Carnian- Norian boundary (see Fig.  3.15). These craters are Rochechouart (France), Manicouagan and Saint Martin (Canada), Obolon (Ukraine) and Red Wing (USA). These five large, chronologically-c lose impact events might have triggered adverse ecological changes that in turn might have led to the high extinction rates observed. However, foraminifera were soon again to recover (see Fig. 3.12), with new forms filling the empty niches. The Norian saw the expansion of the Involutinida in European Tethys. The Late Triassic also witnessed the appearance in the West Pangean reefs of reportedly Permian Lazarus taxa (Stanley, 2001). Presumably these forms had remained rare and geographically isolated during most of Triassic time. Panthalassan volcanic islands might have served as refuges during time of crisis when the Tethys was affected (Stanley, 2003), and probably middle Norian conditions were similar to those of the Permian. 190 Evolution and Geological Significance of Larger Benthic Foraminifera Fig. 3.15. A map showing the position of the major, Carnian- Norian impact craters. During the Rhaetian, foraminiferal diversity in all provinces dwindled gradually to near full extinctions in the three provinces (Tanner et al., 2004; Scherreiks et al., 2010). These extinctions mirror the abrupt extinction seen for all other groups in the marine realm at the end of the Triassic. This Triassic-J urassic extinction is generally recog- nized as being one of the five largest in the Phaenorozoic (Hallam, 1990). Only 16% of the Triassic foraminifera survived the end Triassic extinction. There has been much speculation as to the cause of this extinction event, and debate as to whether there was a catastrophic event at the end of the Triassic, or if it was preceded by a gradual decline. Certainly, a gradual decline of the foraminifera (see Fig. 3.12) was apparent. According to Tanner et al. (2004) the Late Triassic extinction appears protracted rather than catastrophic. Marine regression reduced the available shallow marine habitats for the Triassic foraminifera, and consequent competition may have been the forcing mechanism for extinction (Newell, 1967). Tucker and Benton (1982) specifically cited climate- induced changes occurring during the Late Triassic as a possible cause for mass extinction. Evidence of anoxia has also been documented in Triassic- Jurassic sections and has been suggested to be a significant factor in this event (Hallam and Wignall, 1997; 2000). The Mesozoic Larger Benthic Foraminifera: The Triassic 191 However, the identification of clear evidence of iridium anomalies at the Triassic- Jurassic boundary (Tanner and Kyte, 2005) again raises the possibility that the mass extinction resulted from an extraterrestrial impact, or was triggered by flood basalt volcanism. The Triassic- Jurassic boundary coincides with large scale eruptions that created a flood basalt province that covered at least 5x105 km2 of northeastern North America. Deckart et al. (1997) proposed a large igneous province related to Pangaean rifting, based on correlation of outcrops in French Guyana, Guinea and Surinam. Marzoli et  al. (1999, 2004)  extended the range of the province, which they named Central Atlantic Magmatic Province (CAMP). The CAMP therefore includes eastern North America, northern South America, western Africa and southwestern Iberia (Tanner et al., 2004, see also Fig. 3.16). The synchronicity between the CAMP volca- nism and of marine faunal extinction at the Triassic-J urassic boundary was confirmed by Schaltegger et al. (2007), who carried out biostratigraphic correlation between zonal ammonites and zircon ages. CAMP volcanism would have created anoxic and dysoxic environments caused by rising CO2 levels and related greenhouse effects during the rapid eruption of the flood basalt. The eruption of the CAMP volcanics amplified the rapid global sea fluctuations and is associated with global cooling and glaciation and is Fig. 3.16. A map showing the location of the end Triassic Central Atlantic Magmatic Province (CAMP). 192 Evolution and Geological Significance of Larger Benthic Foraminifera closely associated with the end-T riassic extinction (Bonis et al., 2009; Whiteside et al., 2010; Schoene et al. 2010; Ruhl et al., 2011; Macleod, 2013; Blackburn et al., 2013; Percival et al., 2017), and specifically Deenen et al., (2010) claim that the extinction of the larger benthic foraminiferal correlate well with the detail timing of the CAMP eruptions. Thus, it would seem that a globally significant volcanic event at the end of the Triassic saw another catastrophic decline in larger foraminiferal assemblages and their associated corals. This, as will be seen in Chapter 4, was followed by the now familiar trend of a “reef gap” in the Early Jurassic. 2 1 4 5 3 6 7 8 9 Plate 3.1 Scale bars: Figs 1-9  = 0.2mm. Fig. 1. Glomospirella irregularis Moeller, Sample YA 840.1/ 2,?Middle Triassic, Nwabangyi Dolomite Formation, Shan States, Burma, NHM P4874. Figs 2, 4. Karaburunia rendeli Langer, paratypes figured by Langer (1968) from the Middle Triassic of Turkey, SMF XXVII 5825-6 5826. Figs 3, 5. Eoophthalmidium tricki Langer, paratypes figured by Langer (1968) from the Middle Triassic of Turkey, SMF XXVII 5821, 23. Figs 6-7 . Trocholina multispira Oberhauser, Late Triassic, Idd-e l-S hargi- 1 core, 10540 ft., Gulailah, UCL coll. Fig.  8. Gandinella falsofriedli (Salaj, Borza and Samuel), figured by Zamparelli, Iannace, Rettori (1995) from the Triassic of Italy. Fig. 9. Griphoporella sp. A fragment of dasy- clad algae, Late Triassic (?Norian), Burma, UCL col. 1 2 3 4 5 6 7 8 9 Plate 3.2 Scale bars: Figs 1-9  = 0.2mm. Figs 1, 5-9 . Glomospira meandrospiroides Zaninetti and Whittaker, 1)  Late Triassic, Lebanon; 5-9 ) Anisian, Nwabangyi Dolomite Formation, Yadanatheingi Area, East. Burma, Sample Am 43 (2), 5-6 ) NHM P50818; 7-9 ) NHM P50820. Fig. 2. Glomospira cf. tenuifistula Ho, Triassic, Anisian, Sample AM 37(7) Nwabangyi Dolomite Formation, East. Burma, B.J. Amos collection, NHM P50816. Fig. 3. Glomospira sp., Late Triassic, Lebanon, UCL coll. Fig. 4. Duostomina sp. Brönnimann, Whittaker and Zaninetti, Late Triassic (?Norian), Burma, NHM P49916- 49923. 1 2 3 4 5 Plate 3.3 Scale bars:  Figs 1-5   =  0.2mm. Fig.  1. Dolomitic facies with Glomospira meandrospiroides Zaninetti and Whittaker, Triassic, Nwabangyi Dolomite Formation 105 943, YA 840-1 , N.  Shan State Burma, NHM 1973/ 2. Fig.  2. Earlandia tintinniformis (Misik), Whittaker and Zaninetti, Late Triassic (?Norian), Nwabangyi Dolomite Formation, Burma, Sample YA196, NHM P48777. Figs 3, 5. Aulotortus sinuosus Weynschenk, Sample YA 672, Late Triassic (?Norian), Northern Shan States, Burma, NHM 49876- 49877. Fig. 4. Arenovidalina chialingchiangensis Ho, Triassic (Anisian), East Burma, Amos collection, NHM P50814. 1 2 3 4 5 6 7 8 9 10 11 12 Plate 3.4 Scale bars:  Figs 1- 12  =  0.2mm. Figs 1-2 . Duotaxis birmanica Brönnimann, Whittaker and Zaninetti, Sample Y673, Late Triassic (Norian), Northern Shan States, Burma, NHM 48765. Fig.  3. Agathammina?austroalpina Kristan-T ollmann and Tollmann, Late Triassic (?Norian), Nwabangyi Dolomite Formation, Burma, Sample YA196, NHM P48779. Figs 4-6 Glomospira spp., Late Triassic (Norian), Nwabangyi Dolomite Formation, Burma, Mitchell collection NHM AM 37. Fig. 7. Glomospirella irregu- laris Moeller, Sample YA 840.1/ 2,?Middle Triassic, Nwabangyi Dolomite Formation, Shan States, Burma, NHM P48745. Fig. 8. Ammobaculites sp. Late Triassic (?Norian), Nwabangyi Dolomite Formation, Burma, Sample YA196, NHM P48780. Fig.  9. Turriglomina mesotriasica (Koehn-Z aninetti), figured by Sartorio and Venrurini (1988) from the Ladinian of Abriola, Basilicata, Italy. Fig. 10. Endothyranella sp., figured by Sartorio and Venrurini (1988) from the Ladinian of the Adriatic Sea Fig. 11. Galeanella sp., figured by Sartorio and Venrurini (1988) from the Rhaetian of Germany. Fig. 12. Triasina hantkeni Mason, figured by Sartorio and Venrurini (1988) from the Rhaetian of Italy. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 Plate 3.5 Scale bars Figs 1- 7 = 0.2mm; Fig. 8 = 0.1mm; Figs 9- 15 = 0.2mm. Fig. 1. Meandrospira pusilla (Ho), figured by Sartorio and Venrurini (1988) from Late Olenekian, Trentino, Italy. Fig. 2. Meandrospira dinarica Kochansky- Devidé and Pantić, figured by Sartorio and Venrurini (1988) from the Anisian of Yugoslavia. Fig.  3. Meandrospiranella samueli Salaj, Anisian, Czechoslovakia, holotype figured by Salaj (1983). Figs 4, 5. Aulotortus friedli (Kristan-T ollmann), CQ1, Rhaetian, Chalkis Quarry, Greece, UCL coll. Fig. 6. Grillina sp., CQ22, Rhaetian, Chalkis Quarry, Greece, UCL coll. Fig. 7. Hoynella gr. sinensis (Ho), CQ27, Rhaetian, Chalkis Quarry, Greece, UCL coll. Fig. 8. Piallina bronnimanni Martini et al., holotype fig- ured by Martini et al. (1995), Carnian, Carpatho- Balkanides Belt. Fig. 9. Pilammina densa Pantic, YA 840-1 Triassic, Nwabangyi Dolomite Formation, 105 943, North Shan State Burma, NHM collection, 1973/2 . Fig. 10. Abriolina mediterranea Luperto, 1963, figured by Rettori (1995) from the Middle Triassic of Central Apennines, Italy. Fig. 11. Robuloides reicheli (Reytlinger), figured by Reytlinger (1965) from the Triassic of Precaucasus. Fig. 12. Paulbronnimannia judicariensis (Premoli Silva), figured by Zaninetti et al. (1994) from the Recoaro Limestone of Italy. Fig.  13. Lamelliconus multispirus (Oberhauser), figured by Sartorio and Venrurini (1988) from the Carnian of Pietratagliata, Friuli, Italy. Fig. 14. Thin section photomicrograph of Aulotortus sp., Rhaetian, Chalkis Quarry, Greece, UCL coll. Fig. 15. Thin section photomicrograph of Glomospirella sp., Triassic, Nwabangyi Dolomite Formation, 105 943, YA 840- 1, N.  Shan State Burma, NHM coll. 1973/ 2. Plate 3.6 Scale bars Figs 1-5 = 0.5mm; Figs 6-7 = 0.1mm. Figs 1-5. Alpinophragmium perforatum Flügel, paratypes from the Late Triassic reef limestones of the Northern Alps, SMF XXV11 4981-5. Figs 6-7. Glomospirella irregularis Moeller, Sample YA 840.1/2, ?Middle Triassic, Nwabangyi Dolomite Formation, Shan States, Burma, NHM P4874. 1 2 4 3 66 5 7 A B 1 2 3 4 5 6 7 A B A B A 8 9 10 A B 11 12 Plate 3.7 Scale bars: Figs 1- 12 = 0.2mm. Figs 1, 6. Auloconus permodiscoides (Oberhauser), Sample YA 672, Late Triassic (?Norian), Northern Shan States, Burma, NHM 49879. Figs 2. A) Aulotortus sinuosus Weynschenk, B) Triadodiscus eomesozoicus (Oberhauser), Sample YA 672, Late Triassic (?Norian), Northern Shan States, Burma, NHM 49880. Figs 3, 7. Involutina communis (Kristan), Sample YA 672, Late Triassic (?Norian), Northern Shan States, Burma, NHM 49880. Figs 4, 5. Involutina tenuis (Kristan), Sample YA 672, Late Triassic (?Norian), Northern Shan States, Burma, NHM 49887. Fig.  8. A) Gastropod sp., B) Involutina communis (Kristan), Sample YA 672, Late Triassic (?Norian), Northern Shan States, Burma, NHM P49895. Fig.  9. A) Involutina tenuis (Kristan), B) Aulotortus sinuosus Weynschenk, Sample YA 672, Late Triassic (?Norian), Northern Shan States, Burma, NHM P49896. Fig. 10. A) Involutina tumida (Kristan- Tollmann), B) Involutina communis (Kristan), Sample YA 672, Late Triassic (?Norian), Northern Shan States, Burma, NHM 49915. Fig. 11. Aulotortus sinuosus Weynschenk, Sample YA 672, Late Triassic (?Norian), Northern Shan States, Burma, NHM 49878. Fig. 12. A) Involutina tenuis (Kristan), B) Aulotortus sinuosus Weynschenk, Sample YA 672, Late Triassic (?Norian), Northern Shan States, Burma, NHM 49879. Plate 3.8 Scale bars: 1- 8 = 0.2mm. Figs 1,4 . Involutina communis (Kristan), Sample YA 672, Late Triassic (?Norian), Northern Shan States, Burma, NHM 49881. Fig. 2. Aulotortus pragsoides (Oberhauser), Sample YA 672, Late Triassic (?Norian), Northern Shan States, Burma, NHM 49882. Fig.  3. Aulotortus sinuosa (Weynschenk), Sample YA 672, Late Triassic (?Norian), Northern Shan States, Burma, NHM 49883. Fig. 5. Duotaxis birmanica Brönnimann, Whittaker and Zaninetti, holotype, Sample Y673, Late Triassic (?Norian), Northern Shan States, Burma, NHM 48765. Fig. 6. Triadodiscus eomesozoicus (Oberhauser), Late Triassic (?Norian), Northern Shan States, Burma, NHM 49871. Fig. 7. Planiinvoluta?mesotriasica Baud, Zaninetti and Brönnimann, Late Triassic (?Norian), Northern Shan States, Burma, NHM 49872. Fig. 8. Trochammina sp., Late Triassic (?Norian), Northern Shan States, Burma, NHM 49870. 1 2 A B 3 4 6 5 7 8 203 Chapter 4 The Mesozoic Larger Benthic Foraminifera: The Jurassic 4.1 Introduction The mass extinction in the marine realm at the end of the Triassic affected all groups of the larger benthic foraminifera that had previously survived the end Palaeozoic. Most notable was the total extinction of the Endothyrida. Having survived the much greater extinction event at the end of the Permian, small Triassic endothyrids persisted but with a steadily declining diversity through the Triassic, until their final demise, in the European realm, at the end of the Triassic. This fate was also shared by most other benthic foraminifera, except for a few small forms from the Textulariida, Involutinida and Miliolida orders. Of the different orders of foraminifera with large representatives present in the Jurassic, however, only the agglutinated textulariides exhibited impor- tant evolutionary developments; becoming large, complicated and forming many lineages. The aragonitic/c alcareous microgranular walls of their Palaeozoic ancestors were by now replaced by calcareous walls bonded by organic cement. The early Jurassic witnessed the steady evolution of the agglutinated forms from being small and sim- ple to being internally complicated, which became abundant from the Pliensbachian onwards, thereby giving the carbonate facies of the Jurassic a characteristic appearance that is recognizable throughout Tethys (Fig. 4.1). In contrast to the Triassic benthic foraminifera, the Jurassic larger benthic foramin- ifera have been systematically studied on a regional scale by a number of authors. Early Jurassic Hettangian- Sinemurian foraminifera from the present-d ay Mediterranean region were studied by Septfontaine (1981), Fugagnoli (2000), BouDagher- Fadel (2000), BouDagher- Fadel et al. (2001), Noujaim Clark and BouDagher-F adel (2001, 2004, 2005), BouDagher-Fadel et al. (2001), BouDagher-F adel and Lord (2002), Scherreiks et al. (2006, 2010, 2016), BouDagher-F adel and Bosence (2007), and by BouDagher-F adel (2016). These authors have proposed standard Jurassic biozona- tions for the Mesozoic realm on the basis of foraminiferal generic ranges and assem- blages. Benthic forms in comparable facies have also been illustrated and described from the Early Jurassic of north Italy (Sartoni and Crescenti, 1962; Bosellini and Broglio Loriga, 1971; Castellarin, 1972), and south and central Italy (e.g. Chiocchini et al., 1994; Barattolo and Bigozzi, 1996). Jurassic larger benthic foraminifera have also been studied from Saudi Arabia (e.g. Redmond, 1964, 1965; Banner and Whittaker, 1991; Wyn and Hughes, 2004), Morocco (Hottinger, 1967; Septfontaine, 1984), the southern Tethyan realm (Sartorio and Venturini, 1988), from the ‘vast carbonate plat- form’ (e.g. García- Hernández et al., 1978; Vera, 1988; Rey, 1997)  that included the external zones of the Betic Cordillera of southern Iberia (e.g. González- Donoso et al., 1974; Braga et al., 1981), Gibraltar (BouDagher-F adel et al., 2001), and the Western Mediterranean (BouDagher- Fadel and Bosence, 2007). 204 Evolution and Geological Significance of Larger Benthic Foraminifera A B C D Fig.  4.1. Scale bars  =  0.5mm. A) Thin section of a clayey limestone containing microspheric forms of Orbitopsella, Sinemurian, Morocco; B) Andersenolina elongata (Leupold), figured by Vincent et al. (2014) from the Tithonian of North-E ast Turkey; C) Alveosepta jaccardi (Schrodt), figured by Vincent et al. (2014) from the Kimmeridgian of North- East Turkey; D) Meyendorffina sp., Bathonian, France. In this chapter, a review and revision of the taxonomy of the main genera of the Jurassic larger foraminifera is presented, and their evolutionary lineages and phylo- genetic relationships are discussed. Finally, revised and updated biostratigraphic ages, palaeogeographic and palaeoenvironmental interpretations are presented. 4.2 Morphology and Taxonomy of Jurassic Larger Benthic Foraminifera The dominant larger foraminifera of the Jurassic were the agglutinated Textulariida. The Involutinida persisted throughout, while the Miliolida were present but were mainly composed of morphologically small genera, and it was not before the Cenomanian that larger miliolides played an important role in the benthic assemblages of carbonate platforms. The Lagenida were small and simple in the Jurassic and will not be discussed in this chapter, which will instead focus on the three orders: • Involutinida • Textulariida • Miliolida The Mesozoic Larger Benthic Foraminifera: The Jurassic 205 Age Period Jurassic larger benthic foraminifera Ma INVOLUTINIDA 145.0 201.3 251.9 ? 298.9 MILIOLIDA 358.9 TEXTULARIIDA 419.2 443.8 ALLOGROMIIDA 485.4 Fig.  4.2. The evolution of the Jurassic orders (thick lines) and superfamilies (thin lines) of larger foraminifera. The development and evolution of the major superfamilies of these orders is schemat- ically shown in Fig. 4.2. Order Involutinida Hohenegger and Piller, 1977 An order including all forms with an enrolled second chamber. They have walls that are aragonitic, but commonly they are recrystallised to give a homogenous microgranular structure. They show an umbilical region with pillar-l ike structures on one or both sides of the test. They range from Triassic to Late Cretaceous (Cenomanian). Superfamily Involutinoidea Bütschli, 1880 Forms consisting of a first chamber followed by a planispiral to trochospiral enrolled tubular second chamber. Triassic to Late Cretaceous (Cenomanian). Cam. Ordovician Silurian Devonian Carboniferous Pemian Triassic Jurassic Cretaceous COSCINOPHRAGMATOIDEA AMMODISCOIDEA TEXTULARIOIDEA SPIROPLECTAMMINOIDEA PFENDERINOIDEA VERNEUILINOIDE A LOFTUSIOIDE LITUOLOIDEA NEZZAZATOIDE BIOKOVINOIDE ATAXOPHRAGMIOIDE MILIOLOIDEA INVOLUTINOIDEA 206 Evolution and Geological Significance of Larger Benthic Foraminifera Family Involutinidae Bütschli, 1880 The globular proloculus is followed by a trochospiral, coiled tubular second chamber. Secondary lamellar thickenings on one or both umbilical regions. The aperture is at the open end of the tube. Late Triassic (Norian) to Late Cretaceous (Cenomanian). • Andersenolina Neagu, 1994 (Type species:  Andersenolina perconigi Neagu, 1994). The test maybe lenticular, and conical in shape. The spherical proloculus is fol- lowed by a tubular, trochospiral second chamber. The umbilical side is covered by perforated lamellae added with each whorl and surrounded by small rounded “col- larette” margins. A primary aperture is absent. It differs from Trocholina (Fig. 4.3) and Auloconus (see Chapter 3) by the presence of a perforated umbilical plate rel- ative to the former, and from the latter by the substitution of the primary aperture by pores and the absence of external lamellae. Middle Jurassic to Early Cretaceous (Bathonian to Aptian) (Plate 4.1, figs 4- 5; Fig. 4.8). • Involutina Terquem, 1862 (Type species:  Involutina jonesi Terquem and Piette, in Terquem, 1862). Both umbilical regions are filled with lamellar deposits (Figs 4.4; 4.5). Late Triassic to Late Cretaceous (Norian to Cenomanian) (see Chapter 3; Plate 4.1, fig. 1; Fig. 4.8). • Neotrocholina Reichel, 1956 (Type species:  Involutina conica Schlumberger, 1898). Loeblich and Tappan (1988, p. 300) considered Neotrocholina Reichel as a synonym of Trocholina Paalzow. However, the two forms are different and should be sepa- rated. The outer wall of Neotrocholina is thick (approximately as thick as the spiral septum), perforate and the umbilicus is deeply fissured. On the other hand, the outer wall of Trocholina (= Coscinoconus alpinus Leupold) is thin (usually much thinner than the spiral septum, and often eroded), imperforate, and the test lacks the deeply fissured, canaliculated umbilical structure of Neotrocholina (Fig. 4.6). Late Middle Jurassic to Cretaceous (Bathonian to Cenomanian) (Plate 4.1, figs 2- 3; Fig. 4.8). Primary Characterisc Specific Characterisc Species Apical angle 130 - 150 Neotrocholina lenticularis Biconvex test degrees; pustulate / weakly (Cenomanian) pillared on both sides large, robust; apical angle ca. Perforated Andersenolina alpina Base convex 80 degrees; base pustular umbilical plate (Bathonian - Barremian) large, robust; apical angle ca. 80 degrees; base pustular T. arabica (Cenomanian) Conical tests Base flat Smaller more delicate; apical base strongly T. conica (Bajocian - angle obtuse pillared and Oxfordian) (Plate 4.1, fig. 6) deeply fissured Small, delicate; spiral T. multispira (Late Triassic) chamber <0.1mm high (Plate 3.1, figs 6-7) Large robust; spiral Base flat or weakly convex; Perforated A. elongata (Bathonian - chamber pustules or small pillars umbilical plate High cones tending to Barremian) (Plate 4.1, fig. 4) be parallel sided >0.1 mm high Base strongly convex, pitted and/or granular T. altispira (Cenomanian) Fig. 4.3. Key to main species of Trocholina (T.) and Andersenolina (A.). The Mesozoic Larger Benthic Foraminifera: The Jurassic 207 Primary Characteristics Specific Characteristics Species Aulotortus Wenschenk Streptospiral coiling (chamber - plane (Middle Triassic to Middle oscillates through ca. 45 degrees Jurassic) Involutina Terquem Planispiral coiling umbilici pillared incised (Upper Triassic to Cenomanian) Vidalina Schlumberger umbilici umbonate, smooth (Cenomanian to Santonian) Outer wall thick, perforate; umbilicus Neotrocholina Trochospiral coiling deeply fissured (Jurassic - Cretaceous) Trocholina Outer wall thin, imperforate (Upper Triassic to Cenomanian) Fig. 4.4. Key to main involutinides genera. • Septatrocholina BouDagher- Fadel 2016 (Type species:  Septatrocholina banneri BouDagher-F adel and Banner 2008, type species designated in BouDagher-F adel, 2016). The test is conical, consisting of a globular proloculus followed by a trocho- spirally enrolled, divided tubular second chamber with rudimentary septa around a solid core of pillars, filling the umbilical area. Secondary lamellar thickenings are on one or both umbilical regions. The aperture is terminal at the open end of the tube. This species is distinguished from other involutinides by the solid core of pil- lars filling the umbilicus and the rudimentary septa. Rigaud et al. 2013 included the genus Septatrocholina within the synonyms of Coscinoconus Leupold in Leupold and Bigler 1936, on the basis that Coscinoconus has “possibly slightly constricted endoskeletal structures or wall thickenings” similar to the rudimentary septa of Septatrocholina. However, the septa in the latter are almost non-e xistent, while those of Coscinoconus are complete (see Rigaud et al. 2013, Figs 7– 11, new illustrations of syntypes of Coscinoconus chouberti). In addition, Septatrocholina lacks the complex canal system which form polygonal nodes at the umbilical surface of Coscinoconus. Jurassic (Callovian to Oxfordian) (Plate 4.2, figs 1- 4; Plate 4.3, figs 1-6 ; Fig. 4.8). • Trocholina Paalzow, 1922 (Type species:  Involutina conica Schlumberger, 1898). Trocholina may be conical, plano-c onvex or lenticular in shape. It consists of a globular proloculus followed by a trochospirally enrolled undivided tubular second chamber around a solid core of pillars, filling the umbilical area. The aperture is at the end of a tubular chamber. Trocholina is distinguished from Neotrocholina by having a thinner, imperforate wall, and in lacking the deeply fissured, canaliculated umbilical structure of the latter (Fig. 4.6). Late Triassic to Late Cretaceous (Norian to Cenomanian) (Plate 3.1, figs 6- 7; Plate 4.1, figs 6- 9; Plate 4.4, fig. 1). 208 Evolution and Geological Significance of Larger Benthic Foraminifera Lamellar-perforate wall Coarse pores Fibrous wall Tubular chamber cavity without septa Protuberant laminar extensions Fig. 4.5. Involutina liassica (Jones), Early Jurassic, Italy, axial section, scale bar = 0.5mm. Thin imperforate wall Globular proloculus Solid core 0.55 mm. A B No fissures or canaliculaons Thick perforate oute r wall 0.50 mm. C D E Deeply fissured umbilicus Fig. 4.6. A) Trocholina conica (Schlumberger) Umm Shaif, core 3883 ft, Bajocian; B) Trocholina sp. sketch; C-E ) Neotrocholina friburgensis Guillaume and Reichel, type figures, late Barremian – early Aptian, west- central Switzerland. The Mesozoic Larger Benthic Foraminifera: The Jurassic 209 Family Ventrolaminidae Weynschenck, 1950 Members of this family have a lenticular test, which is planispiral to low trochospiral with numerous chambers in a rapidly enlarging whorl. The wall is calcareous with two layers, an inner microgranular one and an outer hyaline radial layer. Middle Jurassic to Early Cretaceous (late Bajocian to Berriasian). • Protopeneroplis Weynschenck, 1950 (Type species:  Peneroplis senoniensis Hofker, 1949). The test is enrolled in about two rapidly enlarging and loosely coiled whorls. The final whorl has twelve to sixteen chambers, and is involute with a rounded to angular periphery. The aperture is areal, in the early part of the test and slightly pro- truding. Middle Jurassic to Early Cretaceous (Berriasian) (Plate 4.5, fig. 7; Fig. 4.8). Order Textulariida Delage and Hérouard, 1896 The tests of these agglutinated foraminifera are made of foreign particles bound by organic cement. They range from Early Cambrian to Holocene. Superfamily Verneuilinoidea Cushman, 1911 Members of this family have a trochospiral early stage that is triserial or biserial, which later may be uniserial. Walls are non-c analiculate. Late Carboniferous to Holocene. Family Verneuilinidae Cushman, 1911 Members of this family have biserial tests, at least in early stages. Chambers are glob- ular with a terminal aperture. • Duotaxis Kristan, 1957 (Type species: Duotaxis metula Kristan, 1957). Originally regarded as a member of the Tetrataxidae, Duotaxis differs in having an aggluti- nated rather than a two-l ayered microgranular calcareous wall, so it was reassigned to the Family Verneuilinidae by Loeblich and Tappan (1988). The genus had earlier (Loeblich and Tappan 1964) been considered a synonym of Valvulina, but differs by not having an early triangular stage and in lacking a truly valvular tooth (Loeblich and Tappan 1988). Triassic to Early Jurassic (?Ladinian-R haetian to Pliensbachian) (Plate 3.4, figs 1- 2; Plate 3.8, fig. 5; Plate 4.6, fig. 1). Superfamily Pfenderinoidea Smout and Sugden, 1962 Members of this superfamily (see Fig. 4.7) have a trochospiral test throughout, or one that may become uncoiled. Some forms have a siphonal canal, others develop a central composite columella, with pillars between apertural plates and septa. Early Jurassic (Hettangian) to Cretaceous. Family Pfenderinidae Smout and Sugden, 1962 Members of this family have a loose trochospiral conical test with siphonal canals that connect successive apertures in primitive forms. Some forms develop a central compos- ite columella composed of thickened innermost septal ends (“septal buttons”) with or without additional pillars and a spiral canal between the columella and the thickened inner parts of the adjacent septa. A subcameral tunnel (simple or multiple) is present in advanced forms. The chamber interior of advanced taxa is subdivided by vertical 210 Evolution and Geological Significance of Larger Benthic Foraminifera or horizontal (or both) exoskeletal partitions, resulting in a reticulate subepidermal layer. In the trochospires, the spiral and intracameral septa are strongly oblique to the long (coiling) axis (resembling Arenobulimina), but rectilinear, uniserial developments (if present) produce peneropliform, conical (“coskinoliniform”) or lituoliform tests. The aperture is always cribrate, areal. Early Jurassic to Late Cretaceous (Sinemurian to Maastrichtian). Subfamily Pseudopfenderininae Septfontaine, 1988 The test is trochospirally coiled throughout with no subcameral tunnel. The middle of the test has a siphonal canal, or is filled in with a columella made by interseptal pillars and calcitic infilled material. No peripheral partitions or chamberlets are present. Early Jurassic (Hettangian) to Late Cretaceous. • Pseudopfenderina Hottinger, 1967 (Type species: Pfenderina butterlini Brun, 1962). The test is a high trochospiral with numerous chambers. The umbilical part of the chamber interior is filled with numerous pillars that are continuous from chamber to chamber. Early Jurassic (Sinemurian to Bathonian) (Plate 4.6, figs 10- 16). • Siphovalvulina Septfontaine, 1988 (Type species:  Siphovalvulina variabilis Septfontaine, 1988). The test is trochospirally coiled (high or low) in general with three chambers per whorl; the test wall is canaliculate, but rarely visible as such; the interiors of the chambers are free; a twisted siphonal canal connects successive apertures; the aperture is unique, interiomarginal, but may become cribrate in the last chambers of advanced forms. Early Jurassic to Early Cretaceous (Hettangian) to Late Cretaceous? (Plate 4.6, figs 2B, 3- 8, 9A). Subfamily Paleopfenderininae Septfontaine, 1988 The adult chambers of this subfamily may be uncoiled with various shapes. A simple or multiple spiral subcameral canal may be present. There are pillars between apertural plates and septa have calcitic fillings, giving the appearance of a columella. No periph- eral partitions or chamberlets are present. The aperture is terminal and multiple in the apertural plate. Jurassic (Bajocian to early Oxfordian). • Conicopfenderina Septfontaine, 1988 (Type species: Lituonella mesojurassica Maync, 1972). The genus includes conical forms, with a trochospiral early part followed by a rectilinear uniserial part with separated irregular pillars filling the centres of the chambers. The marginal zone of the chambers is not subdivided. Conicopfenderina differs from Parurgonina Cuvillier, Foury and Pignatti Morano, 1968, by the unise- rial arrangement of its chambers, which in the latter is trochospiral. Apertures occur multiply. Middle Jurassic (Bajocian to Callovian) (Plate 4.7, figs 8- 11). • Chablaisia Septfontaine, 1978 (Type species: Pfenderina? chablaisensis Septfontaine, 1977). This genus differs from other pfenderinoids in having a low trochospiral test and by the presence of a spiral canal and calcitic fillings (septal knobs) in the cham- bers. Jurassic (late Bathonian to early Oxfordian) (Plate 4.7, figs 12- 13). • Palaeopfenderina Septfontaine, 1988 (Type species:  Pfenderina salernitana Sarton and Crescenti, 1962). This genus includes forms with tight trochospiral coil- ing, which increase in diameter with spiral height, with a central microgranular The Mesozoic Larger Benthic Foraminifera: The Jurassic 211 columella twisted along the coiling direction. The apertures are multiple and crib- rate. Palaeopfenderina differs from Pfenderina Henson, 1948 (Plate 4.6, figs 6-7 ) by the superficial position of the tunnel, as a groove at the surface of the columella, and by the filling of the inter- pillar spaces with calcitic material. In Pfenderina the filling is less developed and the pillars are clearly visible. Jurassic (Bathonian to Callovian) (Fig. 4.7). • Pseudoeggerella Septfontaine, 1988 (Type species:  Pseudoeggerella elongata Septfontaine, 1988). This genus includes forms with high trochospiral coiling and a narrow columella in the axis of the test. The stalagmitic protuberance is deeply incised against the columella. This genus lacks the subcameral tunnel of Palaeopfenderina and differs from Pseudopfenderina by the presence of calcitic protuberances in the chambers. Jurassic (Bathonian) (Plate 4.7, fig. 14). • Sanderella Redmond 1964 (Type species:  Sanderella laynei Redmond, 1964). The test has long and narrow chambers without secondary septa. The initial portion of the test is a low trochospire with a single subcameral tunnel. The later portion of the test becomes discoid, with multiple subcameral tunnels. Jurassic (Bathonian to Oxfordian) (Plate 4.7, fig. 17). • Satorina Fourcade and Chorowicz, 1980 (Type species: Satorina apuliensis Fourcade and Chorowicz, 1980). This genus differs from Conicopfenderina by possessing radial pillars at the margin of the central mass formed by interseptal pillars. Jurassic (late Bathonian to early Oxfordian) (Plate 4.7, figs 15- 16). • Steinekella Redmond, 1964 (Type species:  Steinekella steinekei Redmond, 1964). This form is a pfenderinine with a massive, central, continuous columella of fused or coalescent pillars, as in Pfenderina, not the discontinuous end skeletons of thickened septal buttons, as in Kurnubia. It has a high trochospiral test with very long and nar- row chambers. Multiple subcameral tunnels combined with strong transverse parti- tions and very fine and weak subepidermal structure are present throughout the test. Jurassic (late Callovian to Oxfordian) (Plate 4.7, figs 18- 20). Subfamily Pfenderininae Smout and Sudgen, 1962 Members of this subfamily have a test with a single subcameral tunnel that is always buried in the columella, which is formed by pillars and calcitic deposits. No peripheral partitions or chamberlets are present. The aperture is multiple, cribrate, and on the apertural plate. Early Cretaceous (Valanginian to Barremian or?Aptian). • Pfenderella Redmond, 1964 (Type species: Pfenderella arabica Redmond, 1964). The test is an elongate cone consisting of a trochoid spiral of relatively short chambers arranged around the axis of coiling in such a manner that each chamber overlaps approximately one- half of its predecessor. Successive chambers are indirectly con- nected by a tunnel. Secondary septa are absent. The test is, without a solid core. Pfenderella differs from Kurnubia, Praekurnubia and Steinekella (Plate 4.7, figs 18- 20) by lacking secondary partitions in the chambers. It differs from Pfenderina in not having a solid central core, and its chambers are also broader in proportion to their length than are those of Pfenderina. Jurassic (Bajocian to Callovian) (Plate 4.7, figs 21- 24). 212 Evolution and Geological Significance of Larger Benthic Foraminifera Pfenderinoidea Period, Epoch and Stage Fig. 4.7. The phylogenetic development of the Pfenderinoidea through the Jurassic. 1 4 5 . 0 1 9 9 . 3 1 7 4 . 1 1 6 3 . 5 A g e ( M a ) J u r a s s i c L a t e E a r l y M i d d l e H e t t . S i n . P l i e n . T o a r c . A a l e . B a j o . B a t h o . C a l l . O x f . K i m . T i t h . S i p h o v a l v u l i n a P s e u d o p f e n d e r i n a P f e n d e r e l l a P f e n d e r i n a P a l a e o p f e n d e r i n a P s e u d o e g g e r e l l a S t e i n e k e l l a S a n d e r e l l a S a t o r i n a C o n i c o p f e n d e r i n a C h a b l a i s i a K u r n u b i a P r a e k u r n u b i a C o n i c o k u r n u b i a K i l i a n i n a P a r u r g o n i n a N e o k i l i a n i n a The Mesozoic Larger Benthic Foraminifera: The Jurassic 213 • Pfenderina Henson, 1948 (Type species:  Eorupertia neocomiensis Pfender, 1938). This genus differs from Pfenderella in having long narrow chambers communicating with each other, and with the exterior of the test, by means of a spiral subcameral tunnel which lies mostly beneath the surface of a solid core. Jurassic (Bajocian to Valanginian) (Plate 4.7, figs 6- 7). Subfamily Kurnubiinae Redmond, 1964 Members of this family have a test with no subcameral tunnel or tunnels, and may or may not have a solid core. The peripheral zone is divided by radial partitions. • Conicokurnubia Septfontaine, 1988 (Type species:  Conicokurnubia orbitolinifor- mis Septfontaine, 1988). The test is conical. Vertical partitions join the centre of the chambers and coalesce with the pillars. The apertures are multiple, cribrate. It d iffers from Dictyoconus Blanckenhorn, 1900 (see Chapter 5) by having a less com- plicated marginal zone of the chambers. Jurassic (late Bathonian to Kimmeridgian) (Plate 4.8, figs 1- 2). • Kurnubia Henson, 1948 (Type species: Kurnubia palastiniensis Henson, 1948). The test is conical, cylindrical or elongate fusiform, consisting of chambers arranged in an elongate trochoid spiral. The inner surfaces of individual chambers are divided into chamberlets by several sets of inward-p rojecting transverse and longitudinal secondary partitions. The aperture is represented by a number of closely- set pores on the base of the test. Jurassic (Bajocian to early Tithonian) (Plate 4.8, figs 3- 21; Plate 4.9, fig. 13C). • Praekurnubia Redmond, 1964 (Type species: Praekurnubia crusei Redmond, 1964). This differs from Kurnubia in having transverse partitions only. Jurassic (Bathonian to early Oxfordian) (Plate 4.8, figs 22- 24). Family Valvulinidae Berthelin, 1880 The test is trochospirally coiled, and generally triserial in the early stage. The wall is microgranular, and may be alveolar. The interior of chambers is simple. The aperture is interiomarginal, with a large valvular tooth (simple or complicated). The Valvulinidae show the first known example of a crosswise- oblique stolons system (the margino- poriform structure of Hottinger and Caus (1982)) among the Pfenderinoidea. This disposition of stolons appears as a morphological convergence in different groups of lituoloids: the valvulinids (trochospirally coiled), the orbitolinids (trocho- to uniserial) and the reniform-d iscoidal lituoloids (planispiral to uniserial). Subfamily Valvulininae Berthelin, 1880 This subfamily includes forms with a simple or complicated valvular tooth plate. Aperture: simple or cribrate. Jurassic (Pliensbachian) to Holocene • Kilianina Pfender, 1933 (Type species: Kilianina blancheti Pfender, 1933). This form is distinguished by its conical test, where the central zone septa thicken and coalesce into an almost solid mass (hummocks). In some forms the latter half of the test is formed by thin irregular plates intergrown by finer broadly- spaced pillars. The early chambers are trochospiral. They coil along a vertical axis and occupy almost half 214 Evolution and Geological Significance of Larger Benthic Foraminifera the test, later they become rectilinear, with the outer parts of chamberlets subdi- vided by numerous pillars. Jurassic (Bathonian to Kimmeridgian) (Plate 4.4, fig. 3; Plate 4.6, fig. 9B; Plate 4.10, figs 12, 14). • Valvulina d’Orbigny, 1826 (Type species: Valvulina triangularis d’Orbigny, 1826). The test is trochospiral throughout, and triserial. The valvular tooth plate is simple or perforated by supplementary apertures. The primary aperture is a basal slit. Jurassic (Pliensbachian) to Holocene (Plate 4.11, fig. 15). Subfamily Parurgonininae Septfontaine, 1988 This subfamily includes genera with eight or more chambers per whorl. The valvu- lar tooth plate is complicated by pillars. The apertures are multiple, with a crosswise- oblique stolons system. Late Jurassic. • Parurgonina Cuvillier, Foury and Pignatti Morano, 1968 (Type species:  Urgonina (Parurgonina) caelinensis Cuvillier et al., 1968). This genus has a highly conical test with numerous chamberlets, separated by curved septal extensions of the outer wall in a low trochospire. The septa break in the umbilical region to form sub-c onical pil- lars. Jurassic (late Oxfordian to Tithonian) (Plate 4.10, figs 17-1 8). • Neokilianina Septfontaine, 1988 (Type species: Paravalvulina complicata Septfontaine 1988). This genus differs from Kilianina in having a higher number of chambers per whorl in the adult stage (25 instead of 8), and a more pronounced development of pillars in the central part of the test. Jurassic (early Kimmeridgian) (Plate 4.10, figs 13, 15- 16). Superfamily Lituoloidea de Blainville, 1825 Members of this superfamily have conical, multilocular, rectilinear and uniserial tests. The early stage has plani- (strepto- ) or trochospiral coiling. The periphery of the cham- bers has radial partitions; but centrally they are with or without scattered, separated pillars. The septa are arched into hummocks (almost solid mass) between the apertures. The bases of the arches can be fused to the hummocks of previous septum, with the apertures then opening at the suture. The alignment of the apertures and thickening of the hummock walls produces the appearance of a series of “gutters”. No true pillars are formed. The walls are solid, non-a lveolar, non-c analiculate. The aperture is simple, with no internal tooth-p lates, areal or multiple, cribrate. Late Triassic (Carnian) to Holocene. Family Hauraniidae Septfontaine, 1988 The test is uncoiled, uniserial or planispirally coiled. The wall is microgranular with a hypodermic network. The septa are simple or with complicated microstructures. The interior of the chambers is simple or with pillars. The aperture is multiple. Jurassic (late Sinemurian) to Cretaceous. Subfamily Hauraniinae Septfontaine, 1988 The test is uncoiled, uniserial or planispirally coiled. The septa are simple or with com- plicated microstructures, and the interior of the chambers have fine pillars in the cen- tral zone. Jurassic (late Sinemurian to Bathonian). The Mesozoic Larger Benthic Foraminifera: The Jurassic 215 • Ataxella Bassoullet and Lorenz, 1988 (Paracoskinolina occitanica Peybernès, 1974). The test is conical to cylindrical with an early streptospiral stage followed by a recti- linear, uniserial stage. The marginal zone is divided by alternating radial partitions. The central zones is full with pillars which are often coalescent. Middle Jurassic (Bathonian) (Plate 4.4, fig. 10). • Cymbriaella Fugagnoli, 1999 (Type species: Cymbriaella lorigae Fugagnoli, 1999). The test is coarsely agglutinated with a coarse irregular subepidermal polygonal net- work, locally appearing as bifurcated alveolar cavities with a blind ending of polyg- onal outline below a thin epidermis. Megalospheric forms possessing a complex embryonic apparatus represented by a spherical proloculus enveloped by a subspher- ical deuteroconch, which is characterized by having short beams (exoskeletal parti- tions of the chamber lumen perpendicular to the septa) perpendicular to the wall. Cymbriaella is the oldest representative of the Hauraniidae that developed a com- plex megalospheric form. It differs from Amijiella by being larger, and having a more irregular, coarser subepidermal network. Jurassic (Hettangian to Pliensbachian) (Plate 4.11, fig. 14). • Gutnicella Moullade, Haman and Huddleston, 1981 (Type species:  Coskinolina (Meyendorffina) minoricensis Bourrouilh and Moullade, 1964). Gutnicella is a new name for Lucasella Gutnic and Moullade, 1967. The test is highly conical with a large spherical proloculus, enclosed in an early planispiral and involute coil, that is later briefly trochospiral, followed by being uniserial and rectilinear. Chambers are subdivided in the outer part by many short radial vertical partitions, creating a narrow peripheral zone of quadrate chamberlets; one row to each chamber. The broad central zone is filled with irregular scattered separated pillars of different sizes. Jurassic (Pliensbachian to Callovian) (Plate 4.11, figs 16- 17; Plate 4.12, figs 7- 9). • Haurania Henson, 1948 (Type species: Haurania deserta Henson, 1948). The radial partitions are delicate, often bifurcating vertically to form a partial tier of peripheral chamberlets. Jurassic (late Sinemurian to Bathonian) (Plate 4.11, figs 18-2 0; Plate 4.12, figs 1- 4). • Meyendorffina Aurouze and Bizon, 1958 (Type species:  Meyendorffina bathonica Aurouze and Bizon, 1958). The test is initially enrolled planispirally with a low, long initial trochospire, later becoming uniserial. Chambers are subdivided by vertical pillars that project a short distance inward from the outer wall, but lack horizontal rafters (exoskeletal partitions of the chamber lumen parallel to the septa and per- pendicular to the beams and the lateral chamber wall, see Hottinger (2006)). Central zone filled with pillars. Jurassic (Bathonian to early Oxfordian) (Plate 4.8, fig. 23B; Plate 4.12, figs 5- 6, 19). • Platyhaurania Bassoullet and Boutakiout, 1996 (Type species: type species Haurania (Platyhaurania) subcompressa Bassoullet and Boutakiout, 1996). This form differs from Haurania in possessing cylindrical chambers in the uncoiled part. Jurassic (late Sinemurian to Toarcian) (Plate 4.12, fig. 12). • Robustoconus Schlagintweit, Velić and Sokač 2013 (Type species: Robustoconus tis- ljari Schlagintweit et al., 2013). The test is conical with an early planispiral stage. Chambers are subdivided into a marginal zone formed of radial and intercalary beams with rafters or horizontal partitions forming a network of chamberlets. The 216 Evolution and Geological Significance of Larger Benthic Foraminifera central zone is large and complex, consisting of anastomising septal excrescences (with constrictions and swellings). Middle Jurassic (early Bajocian) (Plate 4.4, fig.12). • Socotraina Banner, Whittaker, BouDagher-F adel and Samuel, 1997 (Type spe- cies: Socotraina serpentina Banner et al., 1997). Tests are non-c analiculate, septate, initially coiled planispirally or in a low trochospire. Uniserial chambers are filled with near- vertical, subradial partitions with a sinuous serpentine form. In the cen- tral areas of the chambers the partitions fuse laterally. The aperture consists of many small pores between the partitions and is situated subterminally. Jurassic (Pliensbachian to Toarcian) (Plate 4.4, fig. 6; Plate 4.12, figs 10- 11). • Timidonella Bassoullet, Chabrier and Fourcade, 1974 (Type species:  Timidonella sarda Bassoullet et  al., 1974). Chambers have a subepidermal network of hori- zontal and vertical pillars connecting consecutive septa. This network is followed by a zone of quadrangular chamberlets formed by the pillars, an undivided zone termed the annular canal, and a zone of interseptal pillars. Timidonella differs from Orbitopsella in its alveolar hypodermis, which compare with the Cretaceous Loftusia (see Chapter 5). Jurassic (Aalenian to early Bathonian) (Plate 4.12, figs 13- 16). • Trochamijiella Athersuch, Banner and Simmons, 1992 (Type species: Trochamijiella gollesstanehi Athersuch, et al., 1992). This genus is similar to Socotraina, Haurania and Platyhaurania, but has an initial trochospiral coil. Late Jurassic (Bajocian to Callovian) (Plate 4.12, fig. 17). Subfamily Amijiellinae Septfontaine, 1988 Members of this subfamily have an uncoiled or planispirally coiled test. The septa are simple or have a complicated microstructure. The interior of the chambers is simple, however, some genera may develop pillars. Walls may have alveoles. The apertures are multiple. Jurassic (late Sinemurian) to Cretaceous. • Alzonella Banner and Neumann, 1970 (Type species: Alzonella cuvillieri Bernier and Neumann, 1970). A planispiral test with a hypodermis consisting of a coarse lattice of beams and rafters. Jurassic (Bathonian to Callovian) (Plate 4.13, figs 6- 8). • Alzonorbitopsella BouDagher- Fadel, 2008 (Type species:  Alzonorbitopsella arabia BouDagher- Fadel, 2008, genus and species validated in BouDagher-F adel, 2016). A planispiral, annular and discoidal test with no septulae. Any subepidermal retic- ulate mesh is absent in septa, which are thickened around cribrate apertures, but lacking the true pillars linking septum to septum as in Orbitopsella There is a deli- cate reticulate hypodermis of beams and rafters as in Alzonella, but there is a lack of continuation of this structure on to the septa. Chambers are annular, immedi- ately following the large megalospheric proloculus, as in Cyclorbitopsella, but not in the megalospheric Alzonella. The alveolar hypodermis and septa, and the annu- lar A- form differentiate Alzonorbitopsella from Orbitopsella with no (hypodermal) network, and from Timidonella with subepidermal network and partitions. The absence of medial pillars and the presence of subepidermal beams and rafters dif- ferentiate Alzonorbitopsella from Cyclorbitopsella. Jurassic (late Bathonian) (Plate 4.2, figs 5- 10; Plate 4.14, figs 1-6 ). The Mesozoic Larger Benthic Foraminifera: The Jurassic 217 • Amijiella Loeblich and Tappan, 1985 (Type species: Haurania amiji Henson, 1948). The test is straight and uniserial or planispiral, to uniserial in megalospheric gener- ations. The chambers have no pillars in the central zone. The radial partitions are strong, tending to become thickened towards the central zone. They bifurcate ver- tically to form a few, scattered chamberlets. Jurassic (late Sinemurian to Bathonian or?Callovian) (Plate 4.15, figs 2- 4). • Anchispirocyclina Jordan and Applin, 1952 (Type species: Anchispirocyclina henbesti Jordan and Applin, 1952). Tests are compressed and peneropliform. The central zone is filled by a complex reticulum of densely spaced pillars. Jurassic to Early Cretaceous (Oxfordian to early Campanian) (Plate 4.15, figs 6-7 , 8- 9). • Bostia Bassoullet, 1998 (Type species: Bostia irregularis Bassoullet, 1998). A dimor- phic test with clearly distinct microspheric and megalospheric generations. It is characterized by a complicated embryonic apparatus and a lack of pillars in the central zone. Walls have a subepidermal network formed by irregular radial and transverse partitions that constitute a superficial extension of the chambers. It is similar to the unpillared Amijiella and Alzonella, but differs in possessing an ammo- baculoid test morphology in both generations. It has a wall structure similar to Spiraloconulus, but differs by having an orbitoliniform test. Jurassic (Bathonian) (Plate 4.15, figs 10- 11). • Dhrumella Redmond, 1965 (Type species: Dhrumella evoluta Redmond, 1965). The initial part of the test is a low concavo- convex trochoid spiral, with chambers that are completely involute on the convex side and moderately evolute on the concave side. The later chambers become strongly evolute, becoming flattened and unise- rial. Chambers are subdivided into chamberlets by prominent transverse partitions that generally alternate in position from chamber to chamber. Jurassic (Bajocian to Bathonian) (Plate 4.5, fig. 4). • Ijdranella Bassoullet, Boutakiout and Echarfaoui, 1999 (Type species:  Ijdranella atlasica Bassoullet, et  al., 1999). An hauraniid with a compressed peneropliform uncoiled stage and an exoskeleton containing long radial pillars superficially united by a coarse network. Ijdranella differs from Amijiella in its compressed test and from Trochamijiella in its trochospiral initial coiling and canaliculated wall. Jurassic (Pliensbachian) (Plate 4.15, figs 12- 13). • Kastamonina Sirel, 1993 (Type species: Kastamonina abanica Sirel, 1993). Tests are elongated to conical, morphologically similar to Amijiella, but with a much reduced initial coiled portion and a more complex internal structure. The marginal zone of each chamber is subdivided by an irregular polygonal subepidermal network, con- sisting of vertical partitions (beams) and horizontal partitions (rafters), forming numerous irregular alveolar compartments. This genus differs from Haurania by the absence of a planispiral part and by lacking endoskeletal pillars in the centre of the test. It differs from Rectocyclammina (Plate 4.12, fig. 18) in its early chambers and cribrate aperture. Jurassic (Kimmeridgian to Tithonian) (Plate 4.10, fig. 1). • Pseudospirocyclina Hottinger, 1967 (Type species:  Pseudospirocyclina maynci Hottinger, 1967). Tests are planispiral, then rectilinear or flabelliform with an alve- olar hypodermis. There are thick septa with equally thick, scattered irregular pillars. Jurassic (Bajocian to Tithonian) (Plate 4.5, figs 9-1 5). 218 Evolution and Geological Significance of Larger Benthic Foraminifera • Spiraloconulus Allemann and Schroeder, 1980 (Type species:  Spiroconulus perco- nigi Alleman and Schroeder, 1972). The test is planispiral to low trochospiral, then uncoiling to become thick and conical, with a coarsely reticulate hypodermis. Thin septa, linked by thick heavy pillars, are irregularly and widely spaced. The central zone possesses endoskeletons of pillars from septum to septum. The septa are not alveolar. Spiraloconulus differs from Pseudospirocyclina by its thick, flat initial spire and its thick, conical rectilinear stage, by its reticulate hypodermis of “Alzonella- like” structure, and its thin, non-a lveolar septa. Spiraloconulus differs from Robustoconus by its exoskeleton, which consist of a narrow marginal zone of thin-w alled cham- berlets and the large central zone with septa typically agglutinating large grains (Schlagintweit et al., 2013). Jurassic (Bajocian to Callovian) (Plate 4.10, fig. 11). • Streptocyclammina Hottinger, 1967 (Type species:  Pseudocyclammina (Streptocyclammina) parvula Hottinger, 1967). A  peneropliform, low streptospiral test with an empty central zone, lacking pillars. It differs from Pseudocyclammina only in its streptospirality. Jurassic (Pliensbachian to Kimmeridgian) (Plate 4.16, fig. 7). Family Lituolidae de Blainville, 1827 The early stages of the tests are enrolled, but later they may become rectilinear. Walls are made from agglutinating foreign particles. Few chambers (less than 10) per whorl. Carboniferous to Holocene. Subfamily Ammomarginulininae Podobina, 1978 Tests in the early stage are coiled, but later show uncoiling. Aperture are single. Carboniferous (Early Mississippian) to Holocene. • Ammobaculites Cushman, 1910 (Type species:  Spirolina agglutinans d’Orbigny, 1846). The test is simple, not compressed and uncoils in the adult. Apertures are sin- gle and areal. Carboniferous (Mississippian) to Holocene (Plate 5.6, fig. 18). Subfamily Lituolinae de Blainville, 1827 Members differs from the Ammomarginulininae in having a multiple aperture. Late Triassic to Holocene. • Lituola Lamarck, 1804 (Type species: Lituolites nautiloidea Lamarck, 1804). These forms have no internal partitions and a multiple cribrate aperture. Late Triassic to Holocene (Plate 5.9, fig. 14; Plate 5.10, fig. 9). Superfamily Loftusioidea Bradey, 1884 The test is planispiral, but may uncoil in later stage. The wall is agglutinated with dif- ferentiated outer and inner alveolar layers. Late Triassic (Carnian) to Holocene. Family Choffatellidae Maync, 1958 This subfamily is characterised by having hypodermal alveoles. The test is planispiral, but the early part may be streptospiral, lacking continuously developed endoskeletal pillars. Early Jurassic (Pliensbachian) to Late Cretaceous (Coniacian). The Mesozoic Larger Benthic Foraminifera: The Jurassic 219 • Alveosepta Hottinger, 1967 (Type species:  Cyclammina jaccardi Schrodt, 1894). A planispiral test, that may have a streptospiral early stage. The wall is finely and complexly alveolar. The chambers are low with curved septa, perforated fine aper- tures and often have a clear line and/ or median lamella (or pillars) only in equatorial plane. Jurassic (Oxfordian to Kimmeridgian) (Plate 4.12, fig. 20; Plate 4.13, figs 1- 3). • Choffatella Schlumberger, 1905 (Type species: Choffatella decipiens Schlumberger, 1905). A planispiral test, with a wall that is finely and complexly alveolar. The septa have many fine apertures in the median line and are as complex and thick as the hypodermis of the wall. The chambers are high with the central zone being empty with no pillars. Late Jurassic to Late Cretaceous (Oxfordian to Santonian) (Plate 4.5, figs 1- 2; Fig. 5.3; Plate 5.5, Figs. 8–9 ). • Feurtillia Maync, 1958 (Type species: Feurtillia frequens Maync, 1958). The test is planispiral with an involute early stage and a rectilinear later stage with no basal layer. The chamber walls are with narrow, shallow alveoles surmounted by an Alzonella-l ike (Plate 4.12, figs 6- 8) reticulum of beams and rafters. The septa are weakly alveolar with single areal aperture possessing a thickened, invaginated rim. Early Jurassic to Late Cretaceous (Tithonian to Valanginian) (Plate 5.3, fig. 8) • Palaeocyclammina Bassoullet, Boutakiout and Echarfaoui, 1999 (Type spe- cies: Palaeocyclammina complanata Bassoullet et al., 1999). The test is planispiral, compressed and involute with long low chambers. It differs from Pseudocyclammina in having a reticulate subepidermal skeleton comprised of an irregular superficial network, made of short radial blades perpendicular to the septa. Early Jurassic (Pliensbachian) (Plate 4.15, fig. 1). • Pseudocyclammina Yabe and Hanzawa, 1926 (Type species:  Cyclammina lituus Yokoyama, 1890). A  planispiral nautiliform test, sometimes becoming uncoiled. The central zone of the test lacks the continuous endoskeletal pillars (discontinu- ous, columnar partitions on the inner surface of the chamber wall, see Hottinger (2006) for definition). Walls are coarsely alveolar and labyrinthic. Apertures are spread over the apertural face. Early Jurassic to Cretaceous (Pliensbachian to early Maastrichtian) (Plate 4.10, figs 3-1 0). • Redmondellina Banner and Whittaker, 1991 (Type species: Pseudocyclammina powersi Redmond, 1964). This form has a compressed test and differs from Alveosepta in hav- ing pillar-l ike hypodermal extensions linking the septal hypodermis to the anterior of the epidermis of each preceding septum. The division of the hypodermal alveoles distally produces finer and finer alveoles, as they bifurcate or trifurcate on approach- ing the epidermis. The pillars are present only in the median, equatorial plane, and are not spread throughout the chamber space as in Pseudospirocyclina. Late Jurassic (late Oxfordian to Kimmeridgian) (Plate 4.9, fig. 13A; Plate 4.13, figs 4- 5). • Torinosuella Maync, 1959 (Type species:  Choffatella peneropliformis Yabe and Hanzawa, 1926). The test is flabelliform, with simplified, thin septa. Late Jurassic to Early Cretaceous (Oxfordian to Barremian) (Plate 4.15, fig. 5). Family Everticyclamminidae Septfontaine, 1988 Tests are streptospiral or planispiral in early stages, but uncoiled in the adult or uni- serial throughout. Walls are microgranular with an alveolar microstructure. Interiors 220 Evolution and Geological Significance of Larger Benthic Foraminifera of chambers are simple. Apertures are unique and terminal. Early Jurassic to Early Cretaceous (Sinemurian to Aptian). • Everticyclammina Redmond, 1964 (Type species: Everticyclammina hensoni Redmond, 1964). A planispiral test with an alveolar wall. The septal aperture is single, and the initial coil is very short and missing in most random sections, which only show the elongate uniserial part. Loeblich and Tappan (1988) assigned Everticyclammina to the Family Cyclamminidae, but Septfontaine (1988) assigned it to his new, monoge- neric Family Everticyclamminidae. This proposal, seemingly overlooked by Banner and Highton (1990), when revising the genus to accommodate five named species other than the type, is adopted by Fugagnoli (2000) and herein. Fugagnoli’s amended phylogeny and BouDagher- Fadel et al. (2001) illustrate a known origin of the genus Everticyclammina from E. praevirguliana (Plate 4.16, fig. 3) in the Early Jurassic (late Sinemurian) rather than from species of the Middle or Late Jurassic (Callovian or Oxfordian) as generally accepted by previous authors, with a supposed ancestry in near coeval Ammobaculites. Therefore, the earliest known Everticyclammina is of middle Sinemurian age, and this suggests that by this time the genus was more widely dispersed in the Tethyan realm than recorded hitherto. Jurassic to Cretaceous (mid Sinemurian to Aptian) (Plate 4.4, fig. 11; Plate 4.16, figs 3- 6). • Buccicrenata Loeblich and Tappan, 1949 (Type species: Ammobaculites subgood- landensis Vanderpool, 1933). Buccicrenata was erected by Loeblich and Tappan, 1949 with Ammobaculites subgoodlandensis Vanderpool, 1933 as its type species. Loeblich and Tappan (1985, p.100; 1988, p.99) redefined the genus Buccicrenata to include an alveolar wall, which was first illustrated for B. libyca Gohrbandt (1966, p.67, pl.1, fig.11). However, they denied the presence of alveoles in the septum of Buccicrenata. Nevertheless, in examining numerous randomly thin sectioned specimens of Buccicrenata from different localities, some of them hypotypes of both B.  subgoodlandensis and B.  hedbergi, the alveoles are clearly seen to be present, and were used by Banner and Highton (1990) to distinguish this genus from Everticyclammina. Subsequently BouDagher- Fadel (2001) reviewed these forms and traced their evolution from a primitive form, B.  primitiva, from the Kimmeridgian of Lebanon. It is characterized by a planispiral test with septa that are a continuation of chamber walls, but with alveoles reduced and with no basal layer. Jurassic to Cretaceous (Kimmeridgian to Cenomanian) (Plate 4.5, fig.  8; Plate 5.6, fig.16). • Rectocyclammina Hottinger, 1967 (Type species:  Rectocyclammina chouberti Hottinger, 1967). The test is elongate, the early stage is planispiral, later uncoiling and becoming rectilinear. The septa are thick, and the aperture is terminal and cir- cular in the centre of the apertural face. Jurassic to Early Cretaceous (Pliensbachian to early Berriasian) (Plate 4.12, fig. 18). Family Mesoendothyridae Voloshinova, 1958 Members of this family have a strepto- or planispirally coiled test, that has involute initial chambers, and later is uncoiled. Adult chambers are cylindrical or flattened, falciform to cyclical. They are simple, with radial partitions or with pillars. The wall may have alveoles or a hypodermic network. Early Jurassic (Sinemurian) to Holocene. The Mesozoic Larger Benthic Foraminifera: The Jurassic 221 Subfamily Mesoendothyrinae Voloshinova, 1956 Tests are biumbilicate, or very compressed, strepto- or planispirally coiled, but later they are uncoiled to various degrees. Chambers increase rapidly in height with mono- lamellar septa. Adult chambers are cylindrical or flattened. Walls are microgranular to agglutinated, imperforate or alveolar, simple or with radial partitions, with or without pillars. Apertures are a single slit or made of numerous pores. Jurassic (Sinemurian) to Cretaceous. • Mesoendothyra Dain, 1958 (Type species:  Mesoendothyra izjumiana Dain, 1958). The test is involute, globose, and twisted in a tight streptospire. The wall has an outer imperforate layer and an inner alveolar layer. The aperture is a basal slit. Middle Triassic to Jurassic (Ladinian to Kimmeridgian) (Plate 4.5, figs 5-6 ). Subfamily Orbitopsellinae Hottinger and Caus, 1982 Tests have an early planispiral coil followed by a uniserial part in the adult, becoming cylindrical, or falciform to cyclical. The wall has an alveolar microstructure. The inte- rior of the chambers are divided by vertical radial partitions, with pillars in the central zone. Early Jurassic (late Sinemurian to Toarcian). • Cyclorbitopsella Cherchi, Schroeder and Zhang, 1984 (Type species: Cyclorbitopsella tibetica Cherchi et  al., 1984). An orbitopsellinine with adult chambers becoming annular. It has medial pillars in the central zone, but no subepidermal beams and rafters, in contrast to Alzonorbitopsella. Apertures are cribrate. Early Jurassic (late Sinemurian to Toarcian) (Plate 4.17, figs 8-9 ). • Lituosepta Cati, 1959 (Type species: Lituosepta recoarensis Cati, 1959). This species is characterized by having a simple proloculus followed by a planispiral stage, and a well- developed uncoiled part, which becomes fan shaped particularly in the micro- spheric forms. The test possesses multiple cribrate apertures. The apertural openings between the pillars penetrate the height of the chamber space. Peripheral partitions are seen in transverse sections. The flaring flabelliform test and the canaliculated wall distinguishes it from Labyrinthina. Early Jurassic (late Sinemurian to Aalenian) (Plate 4.16, figs 8-9 ). • Orbitopsella Munier- Chalmas, 1902 (Type species: Orbitulites praecursor Gümbel, 1872). An orbitopsellinine with a discoidal test, the first stage being planispiral followed by a flaring, flabelliform stage with 35 to 40 annular chambers. The wall has simple endoskeletal and exoskeletal pillars. Early Jurassic (late Sinemurian to Toarcian) (Plate 4.16, fig.10; Plate 4.17, figs 1- 6). Subfamily Planiseptinae Septfontaine, 1988 Tests are planispiral, compressed laterally and involute. Walls are microgranular with an alveolar microstructure. Interior of chambers are simple, or with vertical partitions and pillars in the median layer. Apertures are multiple. Early Jurassic (late Sinemurian to Pliensbachian). • Planisepta Septfontaine, 1988 (Type species:  Lituola compressa Hottinger, 1967). Tests have vertical radial partitions and pillars. Early Jurassic (Pliensbachian) (Plate 4.9, figs 1-3 ). 222 Evolution and Geological Significance of Larger Benthic Foraminifera • Palaeomayncina Septfontaine, 1988 (Type species:  Mayncina termieri Hottinger, 1967). Differs from Planisepta by the absence of the pillars and partitions. Septfontaine (1988) named Mayncina ternieri Hottinger, 1967 as the type species and proposed that this new genus is closely related to Lituolipora (Plate 4.11, fig. 7). Early Jurassic (Pliensbachian) (Plate 4.10, fig. 2). Subfamily Labyrinthininae Septfontaine, 1988 Tests are planispirally coiled in early stages, but uncoiled in the adult. Walls are thick and may have alveoles. Interior of chambers are subdivided by vertical, radial parti- tions. Pillars may be present in the central zone. Late Jurassic. • Labyrinthina Wenschenk, 1951 (Type species:  Labyrinthina mirabilis Wenschenk, 1951). A  labyrinthine with multiple apertures, with internal radial partitions and pillars. Late Jurassic (Plate 4.9, figs 4- 5). • Orbitammina Berthelin (Type species: Orbicula elliptica d’Archiac, 1843). The test is compressed, with a planispiral early stage. The early chambers are narrow and elon- gate, increasing rapidly in size to produce a reniform test. The aperture is cribrate and radially aligned. The peripheral margin lack the thickness of Orbitopsella, while the internal structure is similar as in Orbitopsella but with finer pillars. Hottinger (1967) and Septfontaine (1981) believe O. elliptica to be the microspheric form, and the senior synonym of Meyendorffina bathonica s.s. but no comparable B-f orm is known to match M. bathonica (Plate 4.11, figs 5-6 ). Also O. elliptica is known in late Bajocian, while M. bathonica is widely regarded as being Bathonian only. Middle Jurassic (late Bajocian to Bathonian) (Plate 4.12, figs 5- 6). Subfamily Levantinellinae Fourcade, Mouti and Teherani, 1997 Characterised by having a single layer of cyclical chambers, undivided by pillars. Jurassic. • Levantinella Fourcade, Mouti and Teherani, 1997 (Type species: Mangashtia? egyp- tiensis Fourcade, Arafa and Sigal, 1984). Peneropliform and compressed axially. A  simple proloculus followed by a planispiral evolute stage and a later uniserial stage. The first chambers are simple, and then subdivided by zigzag pillars with per- forations. Late Jurassic (Oxfordian to Kimmeridgian) (Plate 4.4, fig. 2). • Syriana Fourcade and Mouti, 1995 (Type species:  Syriana khouryi Fourcade and Mouti, 1995). This form is characterised by a compressed flaring test with bilat- eral symmetry, with multiple apertures successively disposed in rows. The numerous chambers are subdivided by many vertical radial subepidermal partitions arranged in a well- developed uniserial stage. Middle Jurassic (late Callovian) (Plate 4.9, fig. 6B). Superfamily Nezzazatoidea Hamaoui and Saint- Marc, 1970 The test is trochospiral or planispiral, later occasionally uncoiled. The wall is sim- ple nonlamellar, microgranular, and may possess internal plates or simple partitions. The aperture is simple or multiple with tooth plates. Middle Jurassic (Bajocian) to Holocene. The Mesozoic Larger Benthic Foraminifera: The Jurassic 223 Family Mayncinidae Loeblich and Tappan, 1985 The test is biumbilicate or very compressed, planispirally enrolled, and rarely uncoiling. Chambers increase rapidly in height. The septa are monolamellar. The wall is simple, solid and microgranular. Middle Jurassic to Late Cretaceous (Bajocian to Santonian). • Daxia Cuvillier and Szakall, 1949 (Type species:  Daxia orbignyi Cuvillier and Szakall, 1949). The test is planispiral, completely involute, lenticular, bilaterally sym- metrical with a moderately sharp margin. Chambers are very narrow. There is a single aperture, that is basal, rounded and central. Late Jurassic to Late Cretaceous (Kimmeridgian to Cenomanian) (Plate 5.3, fig. 11; Plate 5.4, fig. 7). • Freixialina Ramalho, 1969 (Type species: Freixialina planispiralis Ramalho, 1969). The test is biumbilicate, planispiral, with sutures that are slightly curved. The wall is finely agglutinated. The aperture is an areal slit in the middle of apertural face. Jurassic (Bajocian to Tithonian) (Plate 4.5, fig. 3). Superfamily Biokovinoidea Gušiç, 1977 Members of this superfamily have a free test with a trochospiral or planispiral early stage, that later may be uncoiled. The septa are homogenous and massive. The walls have an imperforate outer layer and a canaliculate inner layer. The aperture is basal to areal, single to multiple. Early Jurassic (Sinemurian) to Late Cretaceous (Maastrichtian). Family Biokovinidae Gušiç, 1977 The test is planispirally coiled, later it may be uncoiled. Endoskeletal pillars may be present. Walls are canaliculate, with alveoles that open both to the exterior or interior. Early Jurassic to Cretaceous. • Bosniella Gušiç, 1977 (Type species:  Bosniella oenensis Gušiç, 1977). This genus is characterised by having a well-d eveloped uncoiled later stage with thick, widely spaced and gently curved septa. The aperture is single in the coiled stage becom- ing cribrate in the uncoiled part. The presence or absence of alveoles in the wall is still debatable. Septfontaine (1988, p. 242) put the genus Bosniella in synonymy with Mesoendothyra Dain, 1958, as “the presence or absence of keriotheca which is not always visible due to diagenesis, is not a reliable criterion for the distinction between the taxa Mesoendothyra and Bosniella”. The type species of Mesoendothyra had been originally described as having alveolar within the wall structure. However, in my specimens the wall structure is solid and I still consider the two taxa as being separate. Early Jurassic (late Sinemurian to early Pliensbachian) (Plate 4.16, figs 1-2 ). Family Charentiidae Loeblich and Tappan, 1985 Early stages are planispiral or streptospiral. Walls are finely canaliculated. Apertures are single or multiple. Middle Jurassic to Late Cretaceous (Callovian to Maastrichtian). • Karaisella Kurbatov, 1971 (Type species: Karaisella uzbekistanica Kurbatov, 1971). Tests have streptospiral early coiling but later are planispiral. The base of sep- tum against the previous whorl is thickened and chomata- like as in Charentia (see Chapter 5). Late Jurassic (Oxfordian) (Plate 4.16, fig. 11). 224 Evolution and Geological Significance of Larger Benthic Foraminifera Family Lituoliporidae Gŭsić and Velić, 1978. Tests are planispirally coiled but later may be uncoiled. Walls have coarse alveoles. Early Jurassic. • Lituolipora Gŭsić and Velić, 1978 (Type species: Lituolipora polymorpha Gŭsić and Velić, 1978). Tests have a later stage that may be irregularly coiled, uncoiled and rec- tilinear. Walls are microgranular with coarse alveoles. Early Jurassic (Sinemurian to Pliensbachian) (Plate 4.17, fig. 7). Superfamily Spiroplectamminoidea Cushman, 1927 Tests are planispirally coiled or biserial in early stages, but later biserial. Walls are agglutinated, non- canaliculate. Carboniferous to Holocene. Family Textulariopsidae Loeblich and Tappan, 1982 Members of this family have a biserial early stage, later they may be loosely biserial or uniserial. Walls are agglutinated. Early Jurassic (Sinemurian) to Late Cretaceous (Maastrichtian). • Textulariopsis Banner and Pereira, 1981 (Type species: Textulariopsis portsdownen- sis Banner and Pereira, 1981). On the basis of a single (type) species, Banner and Pereira (1981, p. 98) defined their new genus as “Wall: agglutinated with calcareous cement, solid, imperforate, lacking canaliculi or pseudopores; proloculus succeeded by a rectilinear series of chambers, all biserially arranged; aperture: anteriomarginal, a simple basal, narrow slit. Differs from other small benthic Spirorutilus Hofker in its lack of a planispiral initial stage, from Textularia Defrance in its lack of cana- liculate, pseudoporous walls, and from Pseudobolivina Wiesner by its low aperture and structurally insignificant chitinous endoskeleton”. Two new Cretaceous spe- cies were assigned to the genus by Loeblich and Tappan (1982), plus three other Cretaceous species from the North America, and T. areoplecta from the Early Jurassic (Pliensbachian to Toarcian) of northern Alaska, which had previously been assigned to Textularia. BouDagher-F adel et  al. 2001 confirmed that this seemingly largely Cretaceous genus had originated by the Sinemurian. Early Jurassic to Cretaceous (Sinemurian to Maastrichtian) (Plate 4.8, fig. 25). Superfamily Textularioidea Ehrenberg, 1838 Tests are trochospiral, biserial or triserial in early stages but later may be uniserial or biserial. Walls are agglutinated and canaliculated. Early Jurassic (Sinemurian) to Holocene. Family Chrysalidinidae Neagu, 1968 Tests are high trochospiral, with quinqueserial, quadriserial, triserial or biserial coil- ing modes, or with certain consecutive pairs of these. The aperture is central along the axis of coiling. In quadriserial or quinqueserial forms, an umbilicus is present and the aperture is covered with a broad umbilical flap, which may be penetrated by multiple accessory apertures. Internal pillars may develop between successive The Mesozoic Larger Benthic Foraminifera: The Jurassic 225 intraumbilical flaps. Banner et al. (1991) divided this family into two subfamilies; the mainly Cretaceous Chrysalidininae (see Chapter 5), which evolved as a single lineage developing pillars in the centre of the test, and the mainly Jurassic Paravalvulininae (see below) which survived into the Valanginian and possibly the Hauterivian. This classification is followed here and in the Chapter 5. Subfamily Paravalvulininae Banner, Simmons and Whittaker, 1991 This subfamily was created by Banner et al. (1991) to include all initially quadriserial or quinqueserial forms, becoming quadriserial or quinqueserial in neanic growth (a postnepionic growth stage with the architecture of an adult shell, see Hottinger, 2006), and then quinqueserial, quadriserial or triserial in the adult. Jurassic (Sinemurian to Kimmeridgian). • Paravalvulina Septfontaine, 1988 (Type species:  Paravalvulina complicata Septfontaine, 1988). Internal pillars between successive umbilical flaps. The early test is quadriserial, later becoming uniserial. Jurassic to?Cretaceous (Bathonian to?Hauterivian) (Plate 4.11, fig. 1). • Pseudomarssonella Redmond, 1965 (Type species:  Pseudomarssonella maxima Redmond, 1965). Forms with cribrate apertures and an umbilicus that is concave with no internal umbilical pillars. Umbilical apertural flaps of successive whorls are broad and are axially separated by a narrow space. Adult tests are quadriserial or quinqueserial. Jurassic (Bathonian to Callovian) (Plate 4.8, fig.  26; Plate 4.11, figs 2- 4). • Redmondoides Banner, Simmons and Whittaker, 1991 (Type species: Pseudomarssonella media Redmond, 1965). The test is quadriserial in the adult form. The umbilicus is concave with no internal umbilical pillars. Apertural flaps of successive whorls are well separated. Septa are flattened with narrow umbilicus. Jurassic (Bajocian to Kimmeridgian) (Plate 4.11, figs 5- 7). • Riyadhella Redmond, 1965 (Type species:  Riyadhella regularis Redmond, 1965). Forms with septa and terminal faces that are highly convex. Redmond’s species of Riyadhella were known to him only as solid specimens extracted from their matrix. Banner et  al. (1991) revised the genus and its four assigned species partly on the basis of thin sections of type material, providing amended descriptions of the Redmond (1965) species R. arabica, R. elongata, R. inflata and R. regularis and plac- ing his species R. hemeri, R. intermedia and R. nana in synonymy with R. regularis. They reassigned the genus to the Family Chrysalidinidae rather than to the Family Prolixoplectidae (as by Loeblich and Tappan 1988). BouDagher-F adel et al. (2001) recorded a new, more primitive species R.  praeregularis in the Sinemurian of the western Mediterranean that reveals canaliculi within the test wall and extended the stratigraphic range of the genus down into the Early Jurassic, and its geographic range significantly westwards within Tethys. Jurassic (Sinemurian to Kimmeridgian) (Plate 4.11, figs 10- 13). • Riyadhoides Banner, Simmons and Whittaker, 1991 (Type species: Pseudomarssonella mcclurei Redmond, 1965). Tests are quadriserial in the adult with flattened septa, or with septa and terminal face concave, (late Bajocian  – Tithonian) (Plate 4.11, figs 8-9 ). 226 Evolution and Geological Significance of Larger Benthic Foraminifera Order Miliolida Delage and Hérouard, 1896 The miliolines have tests that are porcelaneous and imperforate, made of high Mg- calcite with fine randomly oriented crystals. They range from the Carboniferous to the Holocene. Superfamily Milioloidea Ehrenberg, 1839 Tests are coiled commonly with two or more chambers arranged in varying planes about the longitudinal axis, later they may become involute. Advanced forms may have secondary partitions within the chambers. Late Triassic (Norian) to Holocene. Family Nautiloculinidae Loeblich and Tappan, 1985 Tests are free, lenticular, planispiral, and involute with secondary thickening in the umbilical region. Apertures are equatorial. Middle Jurassic to Late Cretaceous. • Nautiloculina Mohler, 1938 (Type species: Nautiloculina oolithica Mohler, 1938) has been variously placed in previous systematics. It was originally described by Mohler (1938, p. 18) as calcareous and imperforate, compared to porcelaneous Planispirina. Loeblich and Tappan (1964b, p. 443) placed the genus into the miliolines. They sub- sequently (1985, p. 92; 1988, p. 71) reallocated this genus into the lituoloids. I do not agree with this classification as Nautiloculina does not possess a microgranu- lar agglutinated wall. Yet, it is not a member of the calcareous rotaliides with the hyaline calcitic perforate wall. The multi-l ayered imperforate wall of Nautiloculina is very difficult to separate from simple fusulinines such as the Carboniferous Millerella Thompson, 1942. Nonetheless, the considerable time gap between the Palaeozoic Loeblichiidae and the Jurassic- Cretaceous Nautiloculinidae would make it difficult to explain a relationship between them. Presently, Nautiloculina has been placed systematically in the Milioloidea (see Noujaim Clark and BouDagher- Fadel, 2001) because some of the miliolides (such as Austrotrillina, Quinqueloculina, etc.) display double layered walls like Nautiloculina, although the thin dark layer is inter- nal to the thick transparent layer of the wall of the normal miliolides, but it is exter- nal to the wall of Nautiloculina. Consequently, the thin dark layer may be basal to the wall of the normal miliolids, while the later part of the wall grow inwards toward the chamber lumen; in Nautiloculina the thin dark layer of the wall could equally be basal to the development of the thicker translucent part of the wall which also would have grown inwards towards the chamber lumen. Tests are biumbonate, nau- tiliform with chambers increasing slowly in height. Jurassic (Late Bajocian) to Early Cretaceous (Aptian) (Plate 4.9, figs 6A, 7- 12; see Chapter 5, Fig. 5.23). 4.3 Biostratigraphy and Phylogenetic Evolution 4.3.1 General Biostratigraphy Following the mass extinction in the marine realm at the end of the Triassic, fora- minifera did not make a full recovery until the middle of the Sinemurian. Hettangian The Mesozoic Larger Benthic Foraminifera: The Jurassic 227 foraminifera were rare and simple, and were predominately small Involutinoidea and Pfenderinoidea (see Figs 4.7, 4.8). Siphovalvulina and Textularia dominated the early Sinemurian (BouDagher- Fadel, 2000; BouDagher- Fadel et  al., 2001; Bosence et  al., 2009; Scherreiks et  al., 2010). However, from the Sinemurian onwards, the Early Jurassic witnessed the steady development of these textulariides from small simple forms to internally complicated forms, which became abundant from the Pliensbachian onwards, yielding a high biostratigraphic resolution for the Tethyan carbonate facies. The Pfenderinoidea and Textularioidea dominated assemblages in the Bathonian and Callovian, but the Lituoloidea/ Loftusioidea seem to have taken over during the Late Jurassic (see Fig. 4.9). The abundance of the agglutinated textulariides (see Chart 4.1) in the Jurassic and their short ranges makes them an invaluable biostratigraphical resource (e.g. Noujaim Clark and BouDagher- Fadel, 2004; BouDagher- Fadel and Bosence, 2007). Involutinoidea Period, Epoch and Stage Fig. 4.8. The phylogenetic development of the Involutinoidea through the Jurassic. 199.3 174.1 163.5 145.0 Age (Ma) Jurassic Early Middle Late Hett. Sine. Pliens. Toar. Aale. Baj. Bath. Call. Oxf. Kimm. Titho. Involutina Trocholina Neotrocholina Andersenolina Septatrocholina Protopeneroplis 228 Evolution and Geological Significance of Larger Benthic Foraminifera Period, Epoch, Involutinoidea Textularioidea Pfenderinoidea Lituoloidea/Loftusioidea Stage Triassic Fig. 4.9. Biostratigraphic ranges and diversity of the main Jurassic superfamilies of Tethys. Details of the genera and their ranges are shown in Chart 4.1 online. 1 9 9 . 3 1 7 4 . 1 1 6 3 . 5 1 4 5 . 0 A g e ( M a ) J u r a s s i c M i d d l e E a r l y L a t e R h a . H e t t . S i n e . P l i e n . T o a r . A a l . B a j . B a t h . C a l l . O x f . K i m . T i t h . The Mesozoic Larger Benthic Foraminifera: The Jurassic 229 Importantly in this epoch, dasyclad algae became abundant and flourished in a reefal environment. As foraminifera became large, they developed blind cham- berlets (alveoles) in their test walls, which housed these symbiotic algae (see Fig. 4.10). From the middle to late Sinemurian, forms with internal pillars, such as Pseudopfenderina (Plate 4.6, figs 10-1 6), and fine alveoles in the walls, as in Everticyclammina (Fig.  4.10), began to appear in the Western Mediterranean. Such features developed in later forms, with larger and more consistent alveoles, in the Pliensbachian to the Late Jurassic and Cretaceous, becoming very impor- tant components of Mesozoic rocks. Septfontaine (1981) had proposed a standard Jurassic biozonation on the basis of 56 foraminiferal generic ranges, which took account of work by earlier authors. His scheme began with an Orbitopsella range zone approximately equivalent to the Pliensbachian:  none of his genera were confidently recorded from the underlying Sinemurian. Later, Septfontaine (1984) recorded ‘Siphovalvulina’ from the Sinemurian- Pliensbachian (Domerian) of Morocco and distinguished a biozone of ‘Siphovalvulina’ and Mesoendothyra, characterized by a fauna of small, primitive lituoloids:  the two index genera plus Everticyclammina praevirguliana, Plate 4.16, fig.3), Glomospira sp. and Earlandia sp. These taxa occurred in all six biozones that he distinguished in the Moroccan Early Jurassic (Hettangian to Pliensbachian), but the lower boundary of this ‘interval zone’ was defined by the first appearance of Lituosepta recoarensis (Plate 4.16, figs 7- 8), the index fossil for the overlying zone. A similar scheme was applied to the Buffadero Member of the Gibraltar Limestone (BouDagher- Fadel et al., 2001). More recently, Fugagnoli (2004) recorded that the Hettangian-e arly Sinemurian inter- val belongs to his Dasyclades Zone, early most late Sinemurian to?Lituosepta recoa- rensis zone, while the lower part of the Sinemurian and early Pliensbachian belongs to the Orbitopsella zone and finally the late Pliensbachian belongs to the Planisepta compressa zone. On the other hand, the Toarcian is dominated by forms with internal radial partitions (e.g. Haurania). Later, BouDagher- Fadel and Bosence (2007), while systematically studying the Early Jurassic foraminifera, divided the Hettangian to Pliensbachian into four biozones, that correlate with strontium isotope dating. In this book, these biozones are revised and the entire Jurassic biozones are plotted (Fig.  4.11). These Jurassic biozones are: • the Siphovalvulina colomi biozone, corresponding to the early Sinemurian and includes S. colomi (Plate 4.1, figs 3- 6), S. gibraltarensis (Plate 4.1, figs 7- 8), Duotaxis metula (Plate 4.6, fig. 1), Riyadhella praeregularis (Plate 4.11, figs 10-1 1), Involutina liassica, Pseudopfenderina butterlini (Plate 4.6, fig. 10- 16); • the Everticyclammina praevirguliana biozone, which corresponds to the mid Sinemurian and coincides with the first appearance of E. praevirguliana (Plate 4.16, fig. 3) and includes Siphovalvulina colomi, S. gibraltarensis, Textulariopsis sinemuren- sis (Plate 4.8, Fig. 25), Riyadhella praeregularis, Duotaxis metula. Foraminifera of the Everticyclammina praevirguliana biozone occur in micritic limestones with asso- ciated algae/ cyanobacteria (Cayeuxia, Plate 4.18, figs 1-2 ; Thaumatoporella, Plate 4.18, figs 3- 4 and Palaeodasycladus mediterraneus, Plate 4.18, figs 5- 7); 230 Evolution and Geological Significance of Larger Benthic Foraminifera Mulple foramen/apertures Septa spread over the apertural face Mulple foramen/apertures Alveoles Short reclinear terminaon Foramen/single Small spherical protoconch B C A Enlargement showing Inclusion of small D connuity between foraminifera the labyrinthic hypodermis Median extension of beams and the septa Fine and complex alveoles Planispiral test with disconnuity between alveoles and septa Coarse alveoles Single areal aperture Uncoiling test Simple alveolar Early streptospiral test Alveolar chamber wall hypodermis Single aperture Thick septa N H E N S F G Very simple septa Single aperture Solid septum penetrated by a single large aperture The Mesozoic Larger Benthic Foraminifera: The Jurassic 231 Fig. 4.10. Alveolar exoskeleton and polygonal network. Scale bars = 0.5mm. A) simple alveolar layer in Everticyclammina virguliana (Koechlin), Zakum-1 , core 9536 ft, ARAB, Kimmeridgian; B) Pseudocyclammina lituus (Yokoyama), equatorial section, Hauterivian, Umm Shaif - 3, cuttings 6710 ft, THAMAMA IV; C) Enlargement of P. lituus (Yokoyama), from the Early Cretaceous of the Persian Gulf, showing continuity between the labyrinthic hypodermis and the septa, and the wholly areal position of the septal apertures. It shows the inclusion of small exotic benthic foraminifera in the labyrinthic part of the hypodermis; D) Alveosepta jaccardi (Schrodt), from early Kimmeridgian of the Persian Gulf, Dukhan-5 1, core 7444 ft, stratotype DARB (Darb 2/3 boundary), equatorial section of a microspheric specimen showing delicately labyrinthic hypodermis and discontinuity between the alveolae of successive septa; E) Everticyclammina virguliana (Koechlin), early Kimmeridgian, after Hottinger (1967), showing the alveolar wall, closer spacing of narrow alveolae like these would mimic the canaculate wall of Charentia (see Chap. 5). Note that the lower parts of the septa (S) often points forward (as in Spinoendothyra), producing very convex, almost conical septa, in which the aperture may be surrounded by a short, thick neck (N); these features are not typical of the Cretaceous Everticyclammina greigi (Henson) (see Chap. 5), the probable descendants of E. virguliana; F) Rectocyclammina cf. chouberti after Ramalho (1971), “Purbeckian”, Tithonian, Portugal, much larger and with elaborately structured hypodermis than in the type R. chouberti; G) Pseudocyclammina sphaeroidalis Hottinger, type specimen from the Kimmeridgian of east Morocco, after Hottinger (1967). It is distinguished from P. lituus (above) by having a simpler alveolar hypodermis and sphe- roidal test; note the very simple septa, often showing only one aperture in any one equatorial cut; H) Streptocyclammina liasica Hottinger, syntypes figured from the Early Jurassic of Morocco. Only the streptospirality readily distinguishes the heavily agglutinated forms of this species from Pseudocyclammina bukowiensis (Plate 4.10, figs 7- 10) and the exceptionally rectilinear forms of this species (see Hottinger, 1967) from P. vasconica (Plate 4.10, fig. 9). 232 Evolution and Geological Significance of Larger Benthic Foraminifera • Lituosepta recoarensis and Orbitopsella spp. biozone, corresponding to the late Sinemurian. It coincides with the first appearance of L. recoarensis (Plate 4.16, figs 8- 9) and includes Siphovalvulina sp., Haurania deserta (Plate 4.11, figs 18- 20; Plate 4.12, figs 1- 4), Orbitopsella praecursor (see Figs 4.18, 4.19; Plate 4.17, figs 2-3), Amijiella amiji (Plate 4.4, figs 8- 9), Pseudopfenderina sp., and Bosniella oenensis (Plate 4.16, figs 1- 2); • Planisepta compressa biozone which corresponds to the early Pliensbachian and coincides with the first appearance of P. compressa (Plate 4.9, figs 1-3 ). It includes Pseudocyclammina sp., Haurania deserta, Amijiella amiji, Riyadhella sp., Siphovalvulina sp., Siphovalvulina colomi, Siphovalvulina gibraltarensis, Textulariopsis sp., Duotaxis metula. Everticyclammina sp., Pseudocyclammina sp., Orbitopsella sp., Haurania sp., Lituosepta recoarensis, Orbitopsella “circumvulvata”, Textularia sp., Siphovalvulina sp., small miliolids, Orbitopsella praecursor, Bosniella oenensis, Amijiella amiji, Haurania deserta, Pseudopfenderina sp., and Buccicrenata sp. (see Chart 4.1); • Socotraina serpentina biozone which corresponds to the Toarcian. The beginning of this biozone coincides with the last appearance of Planisepta. It includes forms with pillared or partitioned central zone, e.g. Socotraina serpentina, Haurania deserta, Amijiella amiji, and Cyclorbitopsella tibetica. The top of this biozone marks the dis- appearance of Orbitopsella (see Fig. 4.11); • Gutnicella cayeuxi biozone which corresponds to the Aalenian. This biozone coin- cides with the first appearance of Timidonella (Plate 4.12, figs 13- 16), which replaced Orbitopsella and Gutnicella. It includes Gutnicella bizonorum, G. minoricensis, G. cay- euxi and Timidonella sarda; • Kurnubia palastiniensis biozone, corresponding to the Bajocian. This biozone coin- cides with the first appearance of Kurnubia, Pfenderina and Conicopfenderina. It includes Pseudopfenderina butterlini, Kurnubia palastiniensis, K.  jurassica, Pfenderina salernitana, P.  trochoidea, Conicopfenderina mesojurassica, Gutnicella cayeuxi, Timidonella sarda, Amijiella slingeri, Rectocyclammina ammobaculitiformis, Pseudocyclammina maynci and Pseudocyclammina bukowiensis; • Ataxella occitanica biozone, corresponding to the Early Bathonian. It coincides with appearance of Andersenolina elongata (Plate 4.1, fig. 4) and includes Trocholina granosa, T.  conica, Redmondoides lugeoni, Kurnubia palastiniensis, K.  jurassica, Pfenderina salernitana, P. trochoidea, Conicopfenderina mesojurassica, Paravalvulina complicata, Pseudocyclammina maynci, P. bukowiensis and Spiraloconulus perconigi. • Alzonorbitopsella arabia biozone, corresponding to the Late Bathonian. It coin- cides with the range of Alzonorbitopsella arabia (Plate 4.2, figs 5- 10) and it includes Redmondoides medius, R. inflatus, Pseudomarssonella maxima, P bipartita, Pseudocyclammina maynci and P. bukowiensis. The top of this biozone witnesses the disappearance of Haurania deserta; • Kilianina preblancheti biozone, corresponding to the Early Callovian. It includes Trocholina transversaria, Andersenolina alpina, Kurnubia palastiniensis, K.  jurassica, Pfenderina salernitana, P. trochoidea, Pseudospirocyclina smouti and Kilianina blancheti; • Pseudospirocyclina smouti biozone, corresponding to the Late Callovian. It includes Rectocyclammina ammobaculitiformis, Pseudocyclammina maynci, Pseudocyclammina bukowiensis, Kurnubia palastiniensis, K. jurassica, Pfenderina salernitana, P. trochoi- dea, Pseudospirocyclina smouti and Kilianina blancheti; The Mesozoic Larger Benthic Foraminifera: The Jurassic 233 Fig. 4.11. Jurassic larger benthic foraminifera biozones defined in this study with diagnostic first and last occurrences. Period, Diagnosis first occurence of Diagnosis last occurence of LBF Zones Epoch LBF LBF and Stage Parurgonina Freixialina planispiralis Freixialina Conicokurnubia Kilianina lata Buccicrenata Parurgonina caelinensis Septatrocholina Anchispirocyclina praelusitanica Anchispirocyclina Meyendorffina Pseudospirocyclina smouti Conicopfenderina Kilianina prebancheti Septatrocholina Alzonorbitopsella arabia Haurania deserta Ataxella occitanica Andersenolina Kurnubia palastiniensis Pfenderina/Kurnubia /Conicopfenderina Lituosepta Gutnicella cayeuxi Timidonella Orbitopsella Socotraina serpentina Planisepta Planisepta compressa Planisepta Lituosepta recoarensis Everticyclammina praevirguliana Haurania deserta Siphovalvulia sinemurensis Orbitopsella Siphovalvulina colomi Siphovalvulina 1 4 5 . 0 1 9 9 . 3 1 7 4 . 1 1 6 3 . 5 A g e ( M a ) J u r a s s i c M i d d l e L a t e E a r l y H e t t . S i n . P l i e n s . T o a r . A a l . B a j . B a t h . C a l l . O x f . K i m . T i t h . 234 Evolution and Geological Significance of Larger Benthic Foraminifera • Anchispirocyclina praelusitanica biozone, corresponding to the Early Oxfordian. It coincides with the first appearance of A. praelusitanica and includes Septatrocholina banneri, Rectocyclammina ammobaculitiformis, Pseudocyclammina maynci and Pseudocyclammina bukowiensis; • Parurgonina caelinensis biozone, corresponding to the Late Oxfordian. It includes Choffatella tingitana, Torinosuella peneroplliformis, Anchispirocyclina praelusitanica and Alveosepta jaccardi. The top of this biozone witnesses the disappearance of Septatrocholina; • Kilianina lata biozone, corresponding to the Kimmeridgian. It includes It includes Kilianina lata, Neokilianina rahonensis, Freixialina atlasica, Buccicrenata primi- tiva, Pseudocyclammina ukrainica, P.  sphaeroidalis, Rectocyclammina chouberti, Redmondellina powersi and Pseudospirocyclina mauretanica. The top of this biozone marks the disappearance of Conicokurnubia; • Freixialina planispiralis biozone, corresponding to the Tithonian. It includes Anchispirocyclina neumanni, Everticyclammina virguliana, E.  praekelleri, Pseudocyclammina lituus and Anchispirocyclina lusitanica. The early part of this biozone witnesses the disappearance of Kurnubia (Plate 4.8, figs 3-2 1), while the top is marked by the disappearance of Parurgonina, Pseudospirocyclina, Kastamonina and Labyrinthina. The Middle and Late Jurassic have been intensively studied from the Middle East (Henson, 1948), Eastern Mediterranean (Noujaim Clark and BouDagher-Fadel, 2002, BouDagher- Fadel and Lord, 2002; Noujaim and BouDagher- Fadel, 2004) and the Western Mediterranean (Septfontaine, 1988, BouDagher-F adel and Bosence, 2007; Bosence et al., 2009), and phylogenetic evolutions have been traced for dif- ferent groups, such as the Textularioidea (Banner et al., 1991) and the Lituoloidea (Septfontaine, 1988). The Middle Jurassic is rich with foraminifera filled with inter- nal pillars (e.g. Haurania), while for assemblages in the Late Jurassic, foraminifera with narrow internal alveoles dominate (see Fig. 4.12). Chart 4.1 summarises the ranges of most important species in the Jurassic, and Chart 4.2 shows the range of the major textulariine superfamilies. Below are discussed the evolutionary lineages and revised phylogenetic evolutions of the most important superfamilies in the Jurassic, namely the Pfenderinoidea, the Lituoloidea, the Textularioidea, and the Involutinoidea. 4.3.2 The Pfenderinoidea of the Jurassic In the Hettangian, small simple textulariides developed a twisted siphonal canal con- necting the successive apertures (Figs, 4.13, 4.14, 4.15). These forms evolved gradually from the Sinemurian to the Bathonian into Pseudopfenderina, a form with high, loose spires with separate pillars filling the narrow central zone (see Fig. 4.7), which in turn evolved forms, such as Pfenderella, with single subcameral tunnels, short chambers with secondarily deposited material but without secondary septa (Plate 4.7, figs 21- 24). These forms are probably the ancestors of Pfenderina, which developed a solid central core in the Bajocian. The pillars in the centre of Pfenderina fuse and coalesce in a strong central zone (Fig. 4.13A, Plate 4.7, figs 6-7 ). Pfenderina persists into the Early Cretaceous. In the Bathonian, Pfenderina evolved into Palaeopfenderina, with a The Mesozoic Larger Benthic Foraminifera: The Jurassic 235 Lituoloidea/Loftusioidea Period, Epoch and Stage Fig. 4.12. The biostratigraphic ranges of the Lituoloidea in the Jurassic. 1 9 9 . 3 1 7 4 . 1 1 6 3 . 5 1 4 5 . 0 A g e ( M a ) J u r a s s i c E a r l y M i d d l e L a t e H e t t . S i n . P l i e n . T o a r c . A a l e . B a j . B a t h . C a l l . O x f . K i m m . T i t h . C y m b r i a e l l a O r b i t o p s e l l a C y c l o r b i t o p s e l l a B o s n i e l l a E v e r t i c y c l a m m i n a H a u r a n i a P l a t y h a u r a n i a A m i j i e l l a S o c o t r a i n a P s e u d o c y c l a m m i n a I j d r a n e l l a P a l a e o c y c l a m m i n a L i t u o s e p t a P l a n i s e p t a P a l a e o m a y n c i n a G u t n i c e l l a T i m i d o n e l l a S p i r a l o c o n u l u s M e y e n d o r f f i n a A l z o n e l l a B o s t i a A n c h i s p i r o c y c l i n a A l v e o s e p t a R e d m o n d e l l i n a L a b y r i n t h i n a L e v a n t i n e l l a K a s t a m o n i n a 236 Evolution and Geological Significance of Larger Benthic Foraminifera Peripheral rectangular chamberlets A B Columella made of coalescent pillars Convex arching of thickened pillars in central zone Fig. 4.13. Difference in morphological structure between: A) Pfenderina trochoidea Smout and Sudgen, Umm Shaif-4 , core 9835 ft; B) Meyendorffina bathonica Aurouze and Bizon, Umm Shaif, core 9370 ft. Scale bar = 0.5mm. superficial position of the tunnel, with a groove at the surface of the columella, and by the filling of the inter-p illar spaces with calcitic material. In the Bathonian, Sanderella (Plate 4.7, fig. 17), the ancestral form of the terminally uncoiled forms, evolved from Pfenderina by developing a flaring, flattened, peneropliform test. In Sanderella, the spiral canal may bifurcate in the rectilinear portion forming multiple subcameral tun- nels. In the late Bathonian, forms with a subcylindrical, lituoliform test evolved from Satorina (Plate 4.7, figs 15- 16), with the tunnel partly or wholly surrounding the rectilin- ear columella, and other forms appear at the same time, with conical, coskinoliniform tests, but with pillars becoming separated in a broad central zone and the tunnel of the rectilinear part being reduced and discontinuous. In the late Bathonian, Chablaisia (Plate 4.7, figs 12- 13) may have evolved directly from Pfenderina by developing a spiral canal and calcitic fillings in the chambers. In the Bathonian, Pseudoeggerella (Plate 4.7, fig.  14) evolved from Pfenderella (Plate 4.7, figs 21- 24) by developing a narrow columella and calcitic protuberances in the chambers. In the Oxfordian, Steinekella (Plate 4.7, figs 18-2 0) replaces Pfenderella by developing a massive, central, continuous columella and multiple subcameral tunnels. Pfenderella (Plate 4.7, figs 21-2 4) probably is also ancestral to Kurnubia (Plate 4.8, figs 3- 8), which developed a peripheral zone divided by radial partitions in the Bajocian (Fig. 4.14). The high, loose, slender trochospiral test with a columella made of thickened innermost septal buttons of Kurnubia developed transverse partitions to form Praekurnubia in the Bathonian to Callovian. This latter in turn developed a broad central zone of a rectilinear cone, with scattered pillars and thickened inner ends of radial partitions, to form Conicokurnubia (Plate 4.8, figs 1- 2) in the Oxfordian to Kimmeridgian. Pfenderina, with secondary infillings and internal partitions, appeared in the Late Jurassic (Oxfordian) and ranged into the Early Cretaceous (Valanginian) of southern Europe, but in the Middle East, survived until the Late Cretaceous (Chart 4.2). The Mesozoic Larger Benthic Foraminifera: The Jurassic 237 Tunnel Septum Raer Beam A B Raer Columella Exoskeleton C D Trochoid spiral of relavely short chambers arranged around the axis of coiling Fig. 4.14. A) Kurnubia wellingsi (Henson), Umm Shaif, core 8886 ft; B) Kurnubia jurassica (Henson), Umm Shaif 3, cutting 10.072feet; C-D ) Pfenderina salernitana Sartonia and Crescenti, Bathonian, Uwainat, Qatar, NHM P43720- 21. Scale bars: 0.4mm. The Valvulinidae show the first known example of a crosswise- oblique stolons sys- tem (the marginoporiform structure of Hottinger and Caus, 1982) (Fig. 4.16). They evolved a trochospiral test, with a valvular tooth plate, Valvulina. This form gives rise to conical tests, where central septa are thickened in a broad central zone (Kilianina, Plate 4.1, fig.9B) in the Bathonian to Callovian. In the Oxfordian to Tithonian they are replaced with highly conical forms, possessing septa breaking in the umbilical region to form sub- conical pillars (Parurgonina, Plate 4.10, figs 17- 18), which in turn gives rise in the early Kimmeridgian to forms with a high number of chambers per whorl and a strongly pillared centre (Neokilianina, Plate 4.10, Figs 13, 15- 16), while com- pressed tests with cyclical chambers with numerous pillars evolved in the Oxfordian to Kimmeridgian (Levantinella, Plate 4.10, fig. 19). 238 Evolution and Geological Significance of Larger Benthic Foraminifera Princple Characteristics Generic characteristics and forms Species No pillars: a low trochospire. Wall possibly alveolar (canaliculate). C. chablaisensis SeptfontaineChablaisia Septfontaine (Bath. - Call. -? Oxf.) High, loose Pillars separated, Loose coil, strong pillars: "arenobuliminid" spires discrete, P. butterlini (Brun) in narrow central Pseudopfenderina Hottingerwith broad chamber spaces (Sin. - Bath.) zone Low trochospire: P. trochoidea High or low trochospires Pillars fuse, Smout and Sudgen (Baj. - Cal.) with reduced chamber coalescent High trochospire, shallow sutures, spaces in strong central smooth outlines: P. salernitana zone Pfenderina Henson Sartoni and Crescenti (Baj. - Oxf.) Deep sutures, lobulate outline: P. neocomensis (Ber. Val.) Test flaring, flattened, Tunnel may bifurcate Sanderella Redmond S. laynei (Cal.) perneropliform Pillars coalescent in recitlinear portion Test subcylindrical, into massive Tunnel partly or wholly S. apuliensis Fourcade and Chorowicz lituoliform columella surrounds rectilinear Satorina Fourcade and Chorowicz Bath? - Cal. columella Pillars discrete Test conical, Tunnel of rectilinear part C. mesojurassica (Maync)separated, in a broad coskinoliniform is reduced and Conicopfenderina Septfontaine, 1988 (Baj. - Cal.)central zone discontinuous radial peripheral partitions Praekurnubia Redmond, 1964 P. crusei Redmond only, no chamberlets in series Bath. - E. Oxf. Columella made of thickened High loose, innermost septal ends "buttons" 6/8/ ch/whorl: K. jurassica (when so slender without pillars with lateral chamberlets loosely coiled last chambers are in trochospires alternating in several tiers Kurnubia Henson uniseries, called K. palastiniensis) Baj.-Kim - ?Tith ca. 20 ch/whorl: K. wellingsi (Baj. -Kim.-?Tith.) Trochospire followed by Broad central zone of rectilinear cone radial partitions produce one Conicokurnubia orbitoliniformis conical, "coskinoline" with scattered pillars and thickened tier of peripheral chamberlets Conicokurnubia Septfontaine Septfontaine, 1988 rectilinear low chambers inner ends of radial partitions in each rectilinear chamber Oxf. - Kim. Fig. 4.15. Morphological characteristics of the key species of the pfenderinids. P e r i p h e r a l z o n e d i v i d e d b y r a d i a l p a r ti o n s N o p e r i p h e r a l p a r ti ti o n s o r c h a m b e r l e t s K u r n u b i i n a e P a l e o p f e n d e r i n i n a e / P f e n d e r i n i n a e W i t h p i l l a r s i n c o l u m e l l a b e t w e e n " s e p t a l b u tt o n s " o f s u c c e s s i v e w h o r l s T e r m i n a l l y u n c o i l i n g T r o c h o s p i r a l t h r o u g h o u t r e c ti l i n e a r The Mesozoic Larger Benthic Foraminifera: The Jurassic 239 Princple Characteristics Generic Characteristics Specific Characteriscs Inial high trochospire Walls of hummocks not thickened Parurgonina Cuvillier et al., 1968 Urgonina (Parurgonina) caelinensis Cuvillier et al., 1968 of 2 to 4 whorls or coalescent: no massive central zone late Oxfordian to Tithonian Test conical; central zone K. rahonensis Foury and Cincent arches not thickened into an almost (Early Kim.) solid mass Inial spire reduced Wall of hummocks are thickened in Kilianina Pfender, 1933 to one whorl only a broad central zone Test flaring, concavo=convex; K. lata Oberhauser central zone sll wth open (Kim.) passage ways; chambers highly arched Test conical, central zone K. blanche Pfender arches thickened and coalescent (Bath-Early Oxf.) into an almost solid mass Fig. 4.16. Morphological characteristics of key species of the valvulinids. H u m m o c k s a n d a r c h e s c o n v e x l y c u r v e d ; n o r e c t a n g u l a r p e r i p h e r a l c h a m b e r l e t s 240 Evolution and Geological Significance of Larger Benthic Foraminifera 4.3.3 The Lituoloidea of the Jurassic Jurassic lituoloids (see Figs 4.12, 4.17) evolved rapidly, exhibiting a marked dimor- phism between microspheric and megalospheric generations. Their rapid evolution, combined with their short ranges, gives the group a very useful biostratigraphic role in the Jurassic. In the Hettangian, the ancestor form of the lituoloids evolved from a simple form, Lituola, with early planispirally enrolled whorls, which later became uncoiled and rectilinear with a solid wall. Lituola evolved into Cymbriaella at the beginning of the Jurassic by developing a coarse irregular subepidermal polygonal network. In the late Sinemurian, Cymbriaella evolved into a form with radial parti- tions, Haurania. This form evolved into Socotraina in the Pliensbachian by filling the uniserial chambers with vertical subradial partitions. Haurania in turn gives rise to Platyhaurania in the latest Sinemurian, which possesses cylindrical chambers in the uncoiled part. Socotraina gives rise, in the Middle Jurassic, to forms with open flabel- liform chambers filled centrally with pillars, with subepidermal nets of alternating horizontal and vertical pillars in Timidonella, and a central zone with irregular pillars of different sizes in Gutnicella. Meyendorffina replaced Timidonella in the Bathonian by losing the horizontal pillars. Early forms, such as Haurania, Amijiella had a comparatively coarse structure with no clear differentiation of an epiderm, just a thin outer wall covering the polygonal network (Fig. 4.20). They also have radial partitions and pillars in the central zone with no orderly differentiation of beams and rafters. This group evolved into forms such as Orbitopsella, which developed an alveolar microstructure in the Sinemurian (Fig. 4.18). Large, complex, internally complicated agglutinating benthic foraminifera with pillars and/ or intramural alveoles, such as Orbitopsella and Cyclorbitopsella, did not appear until latest Sinemurian to Pliensbachian times (Fig.  4.19). Orbitopsella have a flaring second stage, that becomes a well- developed uncoiled part in Lituosepta, and completely annular in Cyclorbitopsella. Lituosepta became planispiral and later- ally compressed in Palaeomayncina, and developed vertical partitions and pillars in Planisepta. In the Late Jurassic, Labyrinthina developed a more important spiral stage than that in Lituosepta. Amijiella evolved in parallel to Haurania in the late Sinemurian by having strong radial partitions, but unlike Haurania it has no pillars in the central zone (Fig. 4.20). Other planispiral forms became uncoiled and developed pillars superficially united by a coarse network (Ijdranella), or in the central zone with thin septa and uncoiled test (Spiraloconulus), or with a central zone filled by complex reticulum of densely spaced pillars (e.g. Pseudospirocyclina, Anchispirocyclina). There appear to be a grade through Pseudospirocyclina smouti (Plate 4.5, figs 9- 10), and/ or P. mauretanica (Plate 4.5, figs 11- 12), and/ or P. maynci (Plate 4.5, figs 13, 15), into Anchispirocyclina lusitanica (Plate 4.15, figs 6- 9) in Early Kimmeridgian. Other uncoiled forms are found without pillars, but develop a subepidermal reticulate mesh, which continues in the septa (Alzonella), or have irregular transverse radial partitions (Bostia). Kastamonina replaces Amijiella in the Kimmeridgian to Tithonian by developing a much reduced early coil and a more complex internal structure. The Mesozoic Larger Benthic Foraminifera: The Jurassic 241 Lituoloidea/Loftusioidea Period, Epoch and Stage Lituola Ammobaculites sp. Fig. 4.17. The phylogenetic development of the Lituoloidea through the Jurassic. 1 4 5 . 0 1 9 9 . 3 1 7 4 . 1 1 6 6 . 1 A g e ( M a ) J u r a s s i c M i d d l e L a t e E a r l y H e t t . S i n . P l i e n s . T o a r . A a l e n . B a j . B a t h . C a l l o . O x f . K i m . T i t h . C y m b r i a e l l a H a u r a n i a P l a t y h a u r a n i a S o c o t r a i n a G u t n i c e l l a T i m i d o n e l l a M e y e n d o r f f i n a A m i j i e l l a P s e u d o c y c l a m m i n a P a l a e o c y c l a m m i n a A l v e o s p e t a R e d m o n d e l l i n a C h o f f a t e l l a A n c h i s p i r o c y c l i n a I j d r a n e l l a S p i r a l o c o n u l u s P s e u d o s p i r o c y c l i n a A l z o n e l l a B o s t i a K a s t a m o n i n a B o s n i e l l a E v e r t i c y c l a m m i n a R e c t o c y c l a m m i n a L i t u o l a B u c c i c r e n a t a O r b i t o p s e l l a C y c l o r b i t o p s e l l a L i t u o s e p t a P l a n i s e p t a P a l a e o m a y n c i n a L a b y r i n t h i n a L e v a n t i n e l l a 242 Evolution and Geological Significance of Larger Benthic Foraminifera sphaeroconch Protoconch Pillars A Pillars Annular passage B Septa Mulple cribVrate a apertures Megalospheric protoconch C Pillars Annular passages Foramen Fig.  4.18. Scale bars  =  0.5mm. The structure of Orbitopsella, having a discoidal test formed by a sim- ple exoskeleton and a pillared endoskeleton: A) Oblique section of a megalospheric Orbitopsella primaeva (Henson) (= Coskinolinopsis primaevus Henson, 1948), type species figured by Henson (1948) from the Early Jurassic of the Musandan limestone, Oman. B) Oblique section of a microspheric Orbitopsella praecursor (Gümbel), Milhala, Oman; C) An enlargement of B showing (a) the alternating pattern in the disposition of the foramen, which are the openings that allow communication between the consecutive chambers, providing passages for functional endoplasm, and (b) oblique centered section of a megalospheric specimen showing the spherical protoconch. The Mesozoic Larger Benthic Foraminifera: The Jurassic 243 Orbitopsella beca Cherchi, Schroeder and Zhang Orbitopsella africana Honger ? Orbitopsella praecursor (Gumbel) Orbitopsella primaeva (Henson) Fig. 4.19. Example of evolutionary lineages of some Orbitopsellinae species (Hottinger, 1967). Microspheric forms: Number of spiral/annular chambers: Orbitopsella primaeva = 35; O. praecursor = +/- 12; O. africana = +/- 12; Cyclorbitopsella tibetica = 10-15 (Initial peneropline spire) + 10-12 (neanic stage of the test) + 35-45 (annular chambers). Diameter of the equatorial section of the test (mm): O. primaeva = (+/-) 5; O. praecursor: 8-10; O. africana = +/-10; C. tibetica = 6-8. Megalospheric forms: Diameter of protoconch (mm): O. primaeva = 0.24-0.32; O. praecursor = 0.3-0.45; O. africana = 0.5-0.6; C. tibetica = 0.32-0.5. Maximum observed number of spiral/annular chambers: O. primaeva = 31; O. praecursor = 8-9; O. africana = 3-5; C. tibetica = 20-25. Diameter of the equatorial section of the test (mm): O. primaeva = 1.2-2.4; O. praecursor = 2-4; O. africana = 2.4-3.6; C. tibetica = 1.2-2. 4.3.4 The Loftusioidea of the Jurassic The Jurassic planispiral forms with no pillars in the central zone, similar to Lituola, but with an alveolar wall, first appeared in the Sinemurian (Everticyclammina prae- virguliana, Plate 4.16, fig. 3). This was immediately followed in the Pliensbachian by Pseudocyclammina, forms with areal cribrate apertures spread over the apertural face, a labyrinthic hypodermis, and a reticulate subepidermal skeleton. The early primitive 244 Evolution and Geological Significance of Larger Benthic Foraminifera Princple Generic Characteristics Specific Characteriscs Characteriscs Radial parons delicate, oen With fine pillars in central zone bifurcang vercally to form a paral Haurania Henson H. deserta Henson er of peripheral chamberlets Chambers high, A. amiji (Henson) Radial parons strong, tending to septa very arched, Late Sin. - Bath. become thickened in transverse secon Amijiella Loeblich and Tappan inial coil very small No pillars in central zone towards central zone, only rarely bifurcate (late Sin. to Bath. or ?Cal.) Chambers low, vercally to form few, scaered chamberlets A. sp. = "Iraniaca slingeri" (Gollesstaneh)septa thick and weakly (Baj.-Bath-?Cal.) arched; inial coil larger Septa arched convexly in Myendorffina n. sp. aff. bathonica Vercal pillars that project Vercal parons thickened in central zone to form a solid or vesicular central mass, Meyendorffina central zone, but quadrately Cal. /Oxf. a short distance inward from the outer wall arching becoming rectangular peripherally Arouze and Bizon, 1958 peripherally, producing Meyendorffina bathonica (Bath) rectangular peripheral Arouze and Bizon chamberlets (Bath.) Fig. 4.20. Morphological characteristics of key species of the hauraniids. The Mesozoic Larger Benthic Foraminifera: The Jurassic 245 Pseudocyclammina has coarse simple alveoles, but soon several independent evolution- ary modifications of the test occur, such as an increased complexity of the hypoder- mis which reached acme in the Santonian with Martiguesia (see Chapter 5), and to increased test compression and coiling rate (e.g. Torinosuella, Plate 4.15, fig. 5). Other lineages deriving from Pseudocyclammina in the Oxfordian to Kimmeridgian are forms with walls with fine and complex alveoles and septa with many apertures (Alveosepta), and with pillar- like hypodermal extensions (Redmondellina). A very few forms from the Jurassic survived into the Early Cretaceous, e.g. Everticyclammina, Pseudocyclammina. Rectocyclammina in the Kimmeridgian has the apertural characters of Everticyclammina and probably grades into it; R. chouberti (Plate 4.12, fig. 18) is very close to E. gr. virguliana (Plate 4.16, fig.  6), however, its internally thickened aper- ture is close to that of Feurtillia. The latter evolved, in latest Jurassic possibly from Ammobaculites by the development of an alveolar hypodermis and thick septa, but with the retention of the simple aperture with the strong apertural neck (see Chapter 5). 4.3.5 The Biokovinoidea of the Jurassic Quite independently a group of foraminifera with canaliculate walls appear in the Sinemurian, the uncoiled Early Jurassic biokovinids and lituoliporids and the coiled planispiral to streptospiral charentiids. The charentiids that appear later in the Callovian have planispiral to streptospiral walls and are finely canaliculate. 4.3.6 The Nezzazatoidea of the Jurassic In the Kimmeridgian, the biumbilicate nezzazatoids made their first appearance. They are distinguished by their simple nonlamellar, microgranular walls and by the simple internal partitions and their apertural tooth plates. Their oldest representative, Freixialina is essen- tially a Jurassic form, ranging from the Bajocian to the Tithonian, while Daxia that appear much later in the Kimmeridgian has a completely involute test and more curved sutures. It survived the Jurassic- Cretaceous boundary, only to die out in the Late Cretaceous. 4.3.7 The Spiroplectamminoidea of the Jurassic The representative of the Jurassic spiroplectamminoids, Textulariopsis first appeared in the Sinemurian (BouDagher-F adel et al, 2001). This essentially Cretaceous genus lacks the early planispiral coil and canaliculi. 4.3.8. The Textularioidea of the Jurassic Within the Textularioidea, the chrysalidinoids evolved in the Jurassic from simple tex- tulariide forms with triserial, quadriserial or quinqueserial tests (with simple interiors) into forms with convex septa and canaliculi in the walls (Riyadhella) (see Fig. 4.21). 246 Evolution and Geological Significance of Larger Benthic Foraminifera Textularioidea Period, Epoch and Stage Fig. 4.21. The phylogenetic development of the Textularioidea through the Jurassic, modified from Banner et al. (1991). These forms developed umbilical apertural flaps with a narrow umbilicus in the Bajocian to Kimmeridgian (Redmondoides), and flattened septa in the late Bajocian to Tithonian (Riyadhoides). Internal pillars between successive umbilical flaps appeared in the Bathonian (Paravalvulina), while in the Bathonian to Kimmeridgian forms with no internal pillars, but with apertural flaps separated only by a narrow space, thrived (Pseudomarssonella). 199.3 174.1 163.5 145.0 Age (Ma) Jurassic Early Middle Late Hett. Sin. Pliens. Toar. Aal. Baj. Bath. Call. Oxf. Kim. Tith. Riyadhella Redmondoides Riyadhoides Paravalvulina Pseudomarssonella The Mesozoic Larger Benthic Foraminifera: The Jurassic 247 4.3.9. The Involutinoidea of the Jurassic After their first appearance in the Late Triassic, Trocholina (Figs. 4.3, 4.6) and Involutina (Figs 4.4, 4.5) continue to thrive in the Jurassic. Forms, with rapidly enlarging and loosely coiled planispiral with many chambers (Protopeneroplis), first appeared in the Middle Jurassic, at the beginning of the Aalenian. In the Bathonian, Trocholina gave rise to forms with a thick outer wall and a deeply fissured umbilicus, Neotrocholina, and forms with a perforated plate covering the umbilicus. In the Callovian, Septatrocholina developed rudimentary septa and survived into the Oxfordian. Most of the involuti- nids survived the Jurassic- Cretaceous boundary, but this order completely disappeared at the end of the Cenomanian. 4.3.10. The Milioloidea of the Jurassic The large miliolids are rare in the Jurassic, however, in the Late Bajocian, a lenticular, planispiral, and involute form, Nautiloculina with secondary thickening in the umbil- ical region made its first appearance. This form has been variously placed in previous systematics, but it is placed here because it displays double layered walls like the mili- olides (see explanation above). Nautiloculina survives the Jurassic - Cretaceous bound- ary and is commonly found in the backreef/ reefal environments of the Cretaceous. 4.4 Palaeoecology of the Jurassic Larger Foraminifera The Jurassic period saw warm tropical greenhouse conditions world-w ide. The sea level gradually rose (O’Dogherty et al., 2000) and the shallow warm waters of Tethys and the Proto- Atlantic flooded large portions of the continents and spread across Europe. The level of oxygen in the atmosphere was recovering gradually through the Jurassic (see Fig.  3.9). The Jurassic sedimentary sequences around the Mediterranean were dominated by warm-w ater, shallow-m arine carbonates that are of crucial importance both as a record of climatic/o ceanic conditions but also as hydrocarbon reservoirs. These deposits are dominantly biogenic in origin, consisting primarily of larger ben- thic foraminifera and algae, with hermatypic corals. Coral reefs were common in the Jurassic, just as they are today. In the Hettangian, the diversity of Tethyan foraminifera was poor, and was made up of small agglutinated forms such as Textularia, Siphovalvulina, Duotaxis (Plate 4.6, fig. 1), and small microgranular and porcelaneous forms such as Glomospira. These small foraminifera were widespread on the platform and have been considered as very tolerant (Septfontaine, 1984; Fugagnoli, 2004). They occur in marginal environments (oligohaline, with terrestrial influx) and in deeper marine environments that reflect ele- vated levels of organic carbon influx (Rettori, 1995), and with high rates of micritic production (Fugagnoli, 2004). The presence of these small foraminifera, with the com- plete absence of larger forms, points to a stressed environment or an ecosystem that suffered severe environmental fluctuations. 248 Evolution and Geological Significance of Larger Benthic Foraminifera In the early Sinemurian, larger benthic foraminifera were mostly textulariides. The biota as a whole is characteristic of inner carbonate platform environments that were widespread along the rifted western margins of the Early Jurassic Tethys. They are more primitive than species well- known from the later Early Jurassic (Pliensbachian). These Sinemurian assemblages still included distinctive smaller foraminifera, such as Siphovalvulina (with depressed chambers). Microflora are present as the probable cyano- bacterium Cayeuxia piae (Plate 4.18, fig 1-2 ), the dasyclad green alga Palaeodasycladus mediterraneus (Plate 18, figs 5- 7), and the disputed alga Thaumatoporella parvovesicu- lifera (Plate 4.18, figs 3-4 , 9). Thaumatoporella were in general widespread in Early Jurassic platform carbonate of Tethys. Palaeodasycladus was abundant and is well- preserved, consistent with deposition in shallow marine, inner platform conditions. These small foraminifera and dasycladacean algae are all found within limestones that show a range of shallow carbonate platform lithologies, largely packstones and grainstones that were subjected to periodic emergence, calcrete formation, and ero- sion along the margin of Tethys (BouDagher- Fadel et al., 2001; BouDagher- Fadel and Bosence, 2007). The late Sinemurian is characterized by the presence of larger benthic foramin- ifera, with large, test- wall surface area, as in Orbitopsella, with many small cham- berlets, which may have been used for hosting endosymbionts (Hottinger, 1982). Such larger foraminifera are highly adapted to mesotrophic and oligotrophic (nutrient- deficient) conditions (Fugagnoli, 2004). The Orbitopsella assemblages are present in peloidal wackestone/p ackstone deposits of the inner platform, together with Lituosepta, Pseudopfenderina, Everticyclammina and Haurania in Morocco and Spain (BouDagher-F adel and Bosence, 2007). Such an assemblage is not very tolerant of environmental change (Fugagnoli, 2004), and is encountered in shal- low water carbonate facies all along the southern Tethyan margin (Septfontaine, 1984). These complex and highly diverse faunas point to the establishment of stable ecosystems. Orbitopsella is found throughout the Pliensbachian, together with the appearance of new forms with a so- called subepidermal network, which is described by Hottinger (1996) as a shell architecture adapted to avoid photoinhibition in the lowermost photic zone. Forms with coarsely alveolar tests, such as Pseudocyclammina, made their first appearance. These large hypodermal alveoles may well have had a function of har- bouring photosymbionts, which would have thrived in the shallow palaeoenvironment populated by these foraminifera (Banner and Whittaker, 1991). These forms such as Haurania, Amijiella (which appeared in the late Sinemurian), Socotraina and Lituosepta (which first appeared in the Pliensbachian) indicate established shallow warm waters along Tethys in that period of the Early Jurassic. The early Toarcian transgression is marked by a brief period of global warming (Jenkyns, 2003)  and the occurrence of organic carbon rich shales in large parts of western Europe and other parts of the world. There is a positive carbon isotope excur- sion of pelagic limestones in several of the Tethyan sections. The widespread occur- rence of the early Toarcian shales is explained by an Oceanic Anoxic Event (OAE) (Jenkyns, 1988; Jenkyns and Clayton, 1997). Despite these highly unfavourable living conditions for benthic foraminifera, the number of forms going extinct is low, with new forms appearing at the Pliensbachian-T oarcian boundary possessing the so- called The Mesozoic Larger Benthic Foraminifera: The Jurassic 249 subepidermal network and alveolar walls. Extinctions may have occurred regionally or locally, however, where relative sea level low stands resulted in enclosed stagnated basins with adverse environments. The Aalenian saw the gradual recovery for the shallow carbonate environment. However, it was not before the Bajocian to Bathonian that an explosion of these large complicated forms became prominent in the shallow carbonate facies of Tethys, extending from the shallow carbonates of Japan to those of the Middle East, Europe, and Tanzania. The “cold snap” at the Callovian-O xfordian (Jenkyns, 2003)  might have trig- gered the extinction of many elongated, internally complicated forms such as the pfenderinoids. Shallow water forms with large intramural alveoles persisted into the Tithonian and were in association with green algae (e.g. Clypeina jurassica, Plate 4.18, fig.  11) and stromatoporoids (e.g. Cladocoropsis mirabilis, Plate 4.18, fig.  12). Forms with narrow alveoles and a regularly labyrinthic hypodermis (e.g. Alveosepta, Redmondellina, Choffatella) occurred from Portugal through North Africa and southern Europe to the Middle East, ranging from the late Oxfordian to Kimmeridgian and occupied a deeper water palaeoenvironment than the contem- poraneous forms with larger alveoles and irregularly labyrinthic hypodermis (e.g. Everticyclammina). Their appearance seems to have extended the distribution of these foraminifera further into outer neritic waters, as they appear to have inhab- ited deeper waters (outer neritic) than Pseudocyclammina (with large hypodermal alveoles), or to have tolerated water richer in argillaceous suspensions. According to Banner and Whittaker (1991), they seem to have thrived under reduced illumi- nation, in which conditions codiacean and dasyclad algae were rare or absent. It is possible that the narrow hypodermal alveoles allowed ionic exchange between inter- nal cytoplasm and surrounding seawater, through the extremely thin hypodermis. Hughes (2004) in analyzing the significance of alveoles in Pseudocyclammina, and in the light of its occurrence with deep water foraminifera in the Hanifa Formation, Saudi Arabia, argued that the interpreted function of broader alveoles needs further consideration. While agreeing with Banner and Whittaker (1991) that the presence of alveoles may have enabled the organism to construct a test of the required size in a muddy environment, it is also possible low oxygen availability may have been their main rationale. The alveoles gave the organism greater protection from anoxic, hostile sulphide-e nriched bottom waters. Low oxygen availability has also been dis- cussed by Preece et al. (2000), who considered that the complex wall structure and presence of alveoles are a means to increase the surface area to volume ratio for gaseous exchange under conditions of low oxygen availability. On the other hand, the increased internal surface area provided by the alveoles may have increased the efficiency of the symbiotic algae, by sheltering them from abnormal water chemistry within the photic zone. Pseudospirocyclina and Anchispirocyclina are not recorded from the Kimmeridge (at the Alveosepta jaccardi horizon) of the eastern Mid East Gulf; probably both these genera were restricted to shallower, more inner shelf envi- ronmental than Alveosepta jaccardi. By the end of the Kimmeridgian crisis the short lived forms with narrow alveoles had disappeared (except for Choffatella, Fig. 4.22), and only robust forms characteris- tic of shallow clear waters survived the Tithonian and crossed over to the Cretaceous. 250 Evolution and Geological Significance of Larger Benthic Foraminifera Raers Narrow chambers Beams increasing slowly in height foramen septum Fine and complex alveoles Fig.  4.22. Choffatella decipiens Schlumberger, megalospheric form, Aptian of the Persian Gulf. Scale bar = 0.3mm. 4.5 Palaeogeographic Distribution of the Jurassic Larger Foraminifera During the Jurassic, Pangea continued to disintegrate and the extent of the oceans was far more widespread than in the Triassic. The supercontinent fragments began to drift in different directions forming rift valleys, and one of these opened to form the southern part of the North Atlantic Ocean (Fig. 4.23). Polar ice caps were still lacking throughout the Jurassic period (Hallam, 1995), and larger foraminiferal distribution indicates that warm conditions extended to much higher latitudes than today. The Jurassic, which according to Hallam (1978) marks the end of the ancient stages of Earth evolution, was the period where newly evolved characters in the foraminiferal communities became established, thrived and went on to dominate the Cretaceous shal- low marine environment. The development of this new biota in the Jurassic occurred against the background of significant events in the Earth’s history:  the termination of the early Cimmerian orogeny, the opening of the North Atlantic, the Triassic- Jurassic reef destruction of the Tethyan carbonate platform, the pre- Cretaceous late Cimmerian uplifts, and climatic changes that resulted in increased differences between the microfaunas in the different palaeogeographic provinces (Basov and Kuznetsova, 2000). Nonetheless, there is no clear evidence of a catastrophic extinction event during the Jurassic. On the contrary, as is evident from larger benthic foraminifera evolution, many forms experienced expansions in distribution, and forms appeared in new niches, such as the appearance of deeper- water larger benthic foraminifera (Choffatella) in The Mesozoic Larger Benthic Foraminifera: The Jurassic 251 Fig. 4.23. Palaeogeographic and tectonic reconstruction of the Early Jurassic, (by R. Blakey http:// jan.ucc. nau.edu/ ~rcb7/p aleogeographic.html). the Late Jurassic, and most notably the first occurrence (see Fig. 4.24) in the Middle Jurassic of planktonic foraminifera (see BouDagher-F adel et al., 1997; 2015). The similarity of the Mediterranean Jurassic assemblages with those recorded from Southern Turkey, Iran, Saudi Arabia, Iraq and the Gulf States, Syria and Israel is remarkable (Noujaim Clark and BouDagher- Fadel, 2001). This commonality may be surprising given the presence in the Tethys of palaeo- highs, and vertical tectonic movements affecting differentially subsided and uplifted platform blocks, but obvi- ously oceanic circulation must have allowed a cosmopolitan distribution of larger ben- thic foraminifera. Despite their widespread occurrence, the larger Jurassic forms had a significant evolutionary history. The factors driving their evolution involved climatic, paleo- oceanographic, tectonic (or impact) processes, and below is outlined how all of these factors affected their test structure, phylogenetic evolution and distribution. As discussed in the previous Chapter, the Triassic-J urassic boundary marks one of the five largest mass extinctions in the past 500 Ma. However, there is still debate as to 252 Evolution and Geological Significance of Larger Benthic Foraminifera Fig. 4.24. The parallel evolution of the early planktonic foraminifera compared with the lituoloids in the Jurassic Planktonic foraminifera Lituoloidea/Loftusioidea/Pfenderinoidea Period, Epoch and Stage 1 4 5 . 0 1 9 9 . 3 1 7 4 . 1 1 6 3 . 5 A g e ( M a ) J u r a s s i c E a r l y L a t e M i d d l e H e t t . S i n . P l i e n s . T o a r . A a l e . B a j . B a t h . C a l l . O x f . K i m . T i t h . C o n o g l o b i g e r n a G l o b u l i g e r i n a S p i r a l o c o n u l u s M e y e n d o r f f i n a A l z o n e l l a B o s t i a K i l i a n i n a A n c h i s p i r o c y c l i n a A l v e o s p e t a R e d m o n d e l l i n a The Mesozoic Larger Benthic Foraminifera: The Jurassic 253 the cause of this extinction. This might have happened, as explained in Chapter 3, as a response to an impact at the Triassic-J urassic boundary (e.g. Boslough et al., 1996), which in turn might have triggered the eruption of a large igneous flood basalt prov- ince (e.g. Jones et al., 2005). The Tr- J boundary, is contemporaneous with the erup- tion of one of the world’s largest known continental igneous province, the Central Atlantic Magmatic Province (Marzoli et al., 2004). This event would have caused a climatic and biotic crisis at the Tr- J boundary, triggered by the emission volcanic gases and would have been responsible for the abrupt turnover of faunas that occurred at the Rhaetian- Hettangian boundary (Fowell and Olsen, 1993; Fowell and Traverse, 1995; McElwain et al., 1999; Olsen et al., 2002a, 2002b; Percival et al., 2017). These gases could have also caused the short-l ived global warming, of possibly 2–4 oC, that has been implicated as a cause of the Tr-J mass extinction (McElwain et  al., 1999; Beerling and Berner, 2002), and that has been inferred from a marked negative carbon isotope (δ13C) anomaly observed in marine and terrestrial Tr- J boundary strata from Hungary, Canada, and England (Ward et al., 2001; Hesselbo et al., 2002). This crisis at the Tr- J boundary would have inhibited photosynthesis in the shallow seas of Tethys. According to Vermeij (2004) starvation and habitat loss triggered by these conditions would lead to widespread collateral extinction of species. Vermeij argues that condi- tions traditionally identified by palaeontologists as initiating mass extinction, such as oxygen deprivation, oversupply of nutrients and poisoning (by carbon dioxide, meth- ane and sulphides) are manifestations of the ecological avalanche triggered by a crisis among primary producers, and are therefore considered consequences rather than pri- mary causes of extinction. In the aftermath of the Tr-J event, the small involutinides (see Fig. 4.8) with a plani- spiral to trochospiral enrolled tubular second chamber survived and continued into the Jurassic as “disaster forms”. Common in pre-e xtinction shallow marine assemblages, Involutina and Trocholina are rare and small in the immediate post- extinction after- math. Urbanek (1993) introduced the term “Lilliput effect”, which describes a tem- porary, within lineage size decrease of the surviving organisms through an extinction event. This effect, which is the morphological manifestation of a post-c risis ecological stress, explains the pattern of size change which is seen through the Hettangian and early Sinemurian. Following the Kauffman and Erwin (1995) post- extinction event recovery model, Twitchett (2006) divided the post-e xtinction “repopulation interval” into an initial “survival interval” and a later “recovery interval”. In the Hettangian, blooms of small opportunistic genera such as Duotaxis dominate sparse assemblages of new, small tex- tulariine forms, such as Siphovalvulina, that represent the “survival stage”. By the mid- dle Sinemurian, there was a recovery, with an increase in new genera, as new forms evolved to fill the niches which had been vacated by the end Triassic extinction event. This recovery continued through the Sinemurian to the Pliensbachian with no major extinction at the boundary, but with more evolved taxa becoming established through the Pliensbachian. The filling of the larger benthic foraminiferal niches continued through the Pliensbachian, where new forms of lituoloids appeared that went on to play an important role in the evolutionary development of this group throughout the Jurassic and Cretaceous. 254 Evolution and Geological Significance of Larger Benthic Foraminifera The development of the carbonate facies of the western Mediterranean, from the late Pliensbachian onwards, was affected by events which led to the drowning of the Trento Platform (Zempolich, 1993). This stopped the further development of the ben- thic foraminifera in the Jurassic carbonates of the western Mediterranean. However, despite well documented mass extinctions of other faunas, the Pliensbachian and Toarcian show only modestly enhanced (but seemingly not catastrophic) levels of extinction amongst the larger foraminifera. Towards the end of the Pliensbachian, 33% of the larger foraminifera became extinct (see Fig. 4.25). This extinction was fol- lowed by rapid diversification in the late Toarcian. Hallam (1986) proposed that to speak of an end Pliensbachian extinction was misleading, and that it was in fact a low- level event, particularly among benthic marine invertebrates, and not focused at the Pliensbachian- Toarcian boundary, but spreading over into the early Toarcian. This was, he suggested, caused by ocean bottom- water anoxia in Western Europe, evidenced by the development of widespread units of laminated organic-r ich shale (see Vermeij, 2004). In addition, Hallam (1986) argued that there is no evidence for contemporary organic- rich shale sequences (reflecting low bottom-w ater oxygenation) or extinc- tions, in South America. Thus, he concluded that the early Toarcian extinction was a regional European event only and that global explanations were irrelevant, although even in Western Europe organic-r ich shale facies is not universal. However, there were Jurassic Genera 45 40 35 30 25 20 15 10 5 Total Genera 0 New genera Number of extinctions Fig. 4.25. The total number of genera, new appearances and extinctions of larger foraminifera throughout the Jurassic. The extinctions correspond to the end of each stage and the appearances of new genera with the beginning of the following stage. Hettangian Sinemurian Pliensbachian Toarcian Aalenian Bajocian Bathonian Callovian Oxfordian Kimmeridgian Tithonian The Mesozoic Larger Benthic Foraminifera: The Jurassic 255 major global extinction events in the early Toarcian, for example in benthic ostracods (Boomer et al. 2008). Little and Benton (1995) in studying the distribution of the macrofauna in Europe at the Pliensbachian-T oarcian boundary argued that there is no evidence for a sin- gle family- level mass- extinction event at the end of the Pliensbachian stage (Sepkoski, 1989, 1990). Rather, there is a five- zone phase of extinction from the late Pliensbachian to early Toarcian. The event has a global distribution because, although the majority of the family extinctions occurred within Boreal north- western Europe, there were also extinctions in the Tethyan and Austral realms. Of the many events that happened during the Pliensbachian, one of the drivers for possible ecological stress, which may have caused the observed, but diffuse, enhanced extinction rates, is the Karoo- Ferrar flood basalt event (Palfy and Smith, 2000). The Karoo province in South Africa and the Ferrar province in Antarctica (Fig.  4.26) are disjunct parts of a once contiguous Gondwanan large igneous province. It ranks among the most voluminous flood basalt provinces of the Phanerozoic (Rampino and Stothers, 1988) and extends for 4000 km2. The vast majority of the lava volume appears to have been extruded at 183 Ma in about 1 Ma. The anoxic event identified by Hallam (1961) and Wignall (2001) in NW European marine sections and in South America, Fig. 4.26. The Pliensbachian world showing the position of the Karoo- Ferrar flood basalts. 256 Evolution and Geological Significance of Larger Benthic Foraminifera may have been triggered by the eruption of volcanic CO2 and ensuing global warming (Jenkyns, 1999; Wignall, 2005). Unlike the Tr- J boundary, where the CAMP volcan- ism seems to have triggered a major catastrophic extinction, the Pliensbachian event is more diffuse. The reduced impact of the Karoo-F errar large igneous province (LIP) eruption may be related to the fact that it occurred at high paleolatitudes, and so may not have globally affected the paleoclimate so rapidly. As conditions ameliorated during the Aalenian, the biota became more diverse and gradually began to resemble pre- extinction biotas (Harries and Little, 1999). Many forms which survived the Pliensbachian- Toarcian event flourished in established niches during the Toarcian and only very few new forms appeared. The Aalenian, through to the Bathonian, shows a major expansion of larger benthic genera. This may be associ- ated with the global recovery in O2 levels (Fig. 3.10), or the opening up of new habitats as the incipient North Atlantic began to widen as a result of the rifting induced by the CAMP volcanism at the end of the Triassic (see Figure 4.27). Fig. 4.27. Palaeogeographic and tectonic reconstruction of the Middle Jurassic (by R. Blakey http:// jan. ucc.nau.edu/ ~rcb7/ paleogeographic.html). The Mesozoic Larger Benthic Foraminifera: The Jurassic 257 Following the Aalenian- Bajocian regional anoxic event, which is recorded in the Carpathian part of the Western Tethys, the most intense foraminiferal turnover took place at the Bajocian- Bathonian boundary, and coincides with a maximum diversity of Ammonites (O’Dogherty et al., 2000, see here Plate 4.18, fig. 8). In the Bajocian- Bathonian, the diversity of the larger benthic foraminifera is also at its maximum in the carbonate platforms and reefs of Tethyan margins. There was a major influx of new genera (~30%, see Fig. 4.25) at the beginning of the Bajocian and continued towards the Bathonian (~42% of new genera). These foraminifera were mainly agglu- tinated and were characterized by the noticeable development of multiple alveoles in their walls (Basov and Kuznetsova, 2000). The Bajocian saw also the most impor- tant event in the history of foraminiferal evolution, namely the appearance of plank- tonic foraminifera (BouDagher-F adel et  al., 1997; BouDagher-F adel, 2015). These foraminifera, for unknown reasons, began a meroplanktonic mode of life (benthic in the early stage, becoming planktonic in the last stage). These foraminifera were represented by Conoglobigerina (Fig. 4.24) which had a restricted geographic occur- rence; they all occurred in present- day South-C entral and Eastern Europe. They did not become cosmopolitan and holoplanktonic (fully planktonic) until the late Bathonian, with the occurrence of Globuligerina (see BouDagher-F adel et al., 1997; BouDagher- Fadel, 2015). The end Bathonian and Callovian were also associated with enhanced extinction rates (see Fig. 4.25 and Chart 4.2). This might just reflect the vigorous increase in the number of genera to be found, which reached a maximum in the Bathonian, and which would have produced a more competitive evolutionary environment, and hence a higher background rate of extinctions. However, this Middle Jurassic epoch also coincides with at least two major impact events that gave rise to the 80km diameter Puchezh- Katunki crater in Russia and the 20 km diameter Obolon crater in the Ukraine. These events might also have contributed to enhanced environmental stress that could have been responsible for part of the enhanced extinction rate in these stages. The transition from the Middle to the Late Jurassic was characterized by significant changes in oceanography and climate. These changes were accompanied by modifica- tions in the global carbon cycle as shown in the carbon isotope record (Louis-S chmid et  al., 2007). They were triggered by the opening and/o r widening of the Tethys– Atlantic– Pacific seaway and a massive spread of shallow- marine carbonate production leading to higher PCO2, and according to Louis- Schmid et al. (2007) this increase in PCO2 may have triggered changes in the biological carbon pump and in organic carbon bur- ial in the mid Oxfordian. The Oxfordian and Kimmeridgian show another burst of larger foraminifera spe- ciation (Fig. 4.25) that maintained the overall number of genera at a high level, but increased extinction rates at the end of the Kimmeridgian and Tithonian saw numbers of genera decline as the Jurassic came to an end. This general decline may be related to the final opening of the proto- North Atlantic, and a consequent change in global cir- culation patterns. The larger foraminifera in the Oxfordian developed special features, such as narrow alveoles and a regularly labyrinthic hypodermis (e.g. Alveosepta), which helped them to occupy deeper waters than the contemporaneous forms with larger alveoles and irregularly labyrinthic hypodermis (e.g. Everticyclammina). They became 258 Evolution and Geological Significance of Larger Benthic Foraminifera cosmopolitan and can be found from Portugal through North Africa and southern Europe to the Middle East, and range from the late Oxfordian to Kimmeridgian. In parallel to their evolution, planktonic foraminifera occupying the upper waters of the oceans became more established, wholly planktonic and cosmopolitan. The end Kimmeridgian saw only a minor extinction, and these were of forms which colonized deeper waters, while no new larger foraminifera appeared to occupy these empty niches in the Tithonian. Around 30% of the larger benthic foraminifera became extinct towards the end of the Tithonian. The end Jurassic (~145Ma) coincides with two major events: (a) a series of large terrestrial impact events (see Glikson, 2005) includ- ing Moroweng (70 km), Mjolnir (40 km) and Gosses Bluff (24 km), and (b) a major sub-m arine flood basalt event that created the Shatsky Rise, which is the oldest of the great Pacific plateaus with an estimated flood basalt volume of 4.3x106 km3. Mahoney et al. (2005) suggest that this feature is consistent with an impact origin. These events might have been the reasons for the disappearance of long ranging, well established Jurassic larger foraminifera, such as Pseudospirocyclina. However, a number of Jurassic agglutinated foraminifera continued through to the Cretaceous where they flourished and thrived, before their final extinctions within the early Cretaceous. 1 2 3 4 5 6 7 8 9 Plate 4.1 Scale bars: Figs 1, 9 = 0.5mm; Figs 2-3 , 6- 8 = 0.25mm; Figs 4- 5 = 1mm. Fig. 1. Involutina lias- sica (Jones), axial section, Early Jurassic, Italy, UCL coll. Fig. 2. Neotrocholina sp., Callovian/O xfordian, Saudi Arabia, UCL coll. Fig.  3. Neotrocholina valdensis Reichel, Callovian, Saudi Arabia, UCL coll. Fig. 4. Andersenolina elongata (Leupold), late Bathonian, Saudi Arabia, UCL coll. Fig. 5. Andersenolina alpina (Leupold), Kimmeridgian/B erriasian, Lebanon, UCL coll. Fig. 6. Trocholina conica (Schlumberger), Callovian/ Oxfordian, Saudi Arabia, UCL coll. Fig. 7. Trocholina cf. granosa (Frentzen), Bajocian/ Bathonian, Saudi Arabia, UCL coll. Fig. 8. Trocholina palastiniensis Henson, holotype, Jurassic, Kurnub, South Israel, NHM P38477. Fig. 9. Haplophragmoides sp., Trocholina palastiniensis Henson, holotype, Kurnubia jurassica (Henson), Jurassic, Kurnub, South Israel, NHM P38477. 1 2 3 4 5 6 7 8 9 10 Plate 4.2 Scale bars: Figs 1-4  = 0.25; Figs 5- 10 = 0.5mm. Septatrocholina banneri, BouDagher- Fadel, first described by BouDagher-F adel (2008) and validated in BouDagher-F adel (2016); 1- 3) NHM coll., BP 7702, from 8172 ft in Juh- 1 core, Qatar; Callovian-O xfordian, upper Araej Formation; fig. 4, NHM coll., BP 7701, from 9880 ft in Um-S haif- 4 core; Upper Bathonian, basal Uweinat Formation; Abu Dhabi. 1) Paratype, tan- gential section showing rudimentary Septa; 2) Holotype, equatorial section showing the globular proloculus followed by a trochospirally enroled divided tubular second chamber with rudimentary septa; 3) Paratype, tangential axial section showing the rudimentary septa; 4) Paratype axial section in which septa are not vis- ible. Figs 5- 10. Alzonorbitopsella arabia, BouDagher- Fadel, first described by BouDagher-F adel (2008) and validated in BouDagher- Fadel (2016), 5) holotype NHM BP 6626, equatorial section of the annular holo- type with large megalospheric proloculus; 6, 9, 10) NHM BP 6627, from 9879 ¼ ft, 6) Paratype, equatorial section showing the annular test with no septulae, 9) Paratype, oblique axial section showing the delicate reticulate hypodermis of beams and rafters, 10) Paratype, enlargement of the axial section to show that the reticulate hypodermis of beams and rafters does not continue onto the septa; 7, 8) NHM BP 6623, from 9880 ft; Umm-S haif-4 core; Upper Bathonian, basal Uweinat Formation, Abu Dhabi, 7) Paratype, equato- rial section showing the large megalospheric proloculus, 8) Paratype, oblique equatorial section showing the annular chambers immediately following the large megalospheric proloculus. Plate 4.3 Scale bars:  Figs 1-6   =  0.6mm. Figs 1- 6. Septatrocholina banneri BouDagher- Fadel, 1- 3, 5-6 ) Callovian- Oxfordian, Upper Araej Formation, from 8172 ft in Juh- 1 core, Qatar, NHM BP 7702; 4)  late Bathonian, basal Uweinat Formation, from 9880 ft in Um- Shaif- 4 core Abu Dhabi, BP 7701. 1 2 3 4 5 6 9 2 1 3 4 10 6 7 5 9 8 A 11 B 14 12 Plate 4.4 Scale bars: Fig. 1 = 0.25mm; Figs 3- 4 = 1mm; Figs 1, 5-7 , 10- 12 = 0.5mm; Figs 8- 9 = 0.3mm. Fig. 1. Trocholina palastiniensis Henson, paratype, Late Callovian, Kurnub Anticline, Israel, NHM P38477. Fig. 2. Levantinella egyptiensis Fourcade, Arafa and Sigal, type figured by Fourcade et al. (1997), Oxfordian, Jibal As Sahilyeh, Syria. Fig. 3. Kilianina lata Oberhauser, paratype figured by Oberhauser (1956), Kimmeridgian, Karadag, West Taurus, Turkey. Fig. 4. Sanderella laynei Redmond, holotype figured by Redmond (1964), Bathonian/C allovian (probably Early Callovian), basal Upper Dhruma, ARAMCO T- 60- 60A, 40- 50ft. Fig. 5. Haurania deserta Henson, paratype, figured by Henson (1948), Bathonian, Muhaiwir Formation, West Iraq, NHM P35859. Fig. 6. Socotraina serpentina Banner et al., identified wrongly as Milahaina tor- tuosa Smout unpublished species and genus by Smout MS, probably Bajocian, Wadi Milaha, Oman. Fig. 7. Amijiella sp., identified wrongly as Iranica slingeri Gollestaneh MS (1965), probably Bajocian, Wadi Milaha, Oman. Gollestaneh wrongly described it as possessing an initial high trochospire, but is undoubtedly an Amijiella. The species name (nomen nudum) was published by Gollestaneh (1974), but the species has never been validly named. Figs 8-9 . Amijiella amiji (Henson), paratypes, Bathonian, Muhaiwir Formation, Wadi Amij well, West Iraq, NHM M/ 3869-3 870. Fig. 10. Ataxella occitanica (Peybernés), figured by Pelissié et al. (1984) as “Paracoskinolina occitana”, late Bathonian, Pyrénées, France. Fig. 11. Everticyclammina praekelleri Banner and Highton, paratype, Kimmeridgian to Tithonian, Broumana, Lebanon, NHM P52255. Fig. 12. Robustoconus tisljari Schlagintweit, Velić and Solač, 3 axial, oblique sections showing microspheric speci- mens (courtesy of Dr Schlagintweit), figured by Schlagintweit (2013), early Bajocian, Croatia. Plate 4.5 Scale bars: Figs 1- 6, 8-1 6 = 0.5mm; Fig. 7 = 0.3mm. Figs 1- 2. Choffatella tingitana Hottinger, type specimens figured by Hottinger (1967), Kimmeridgian-T ithonian, Morocco, 1) holotype, equatorial B- form; 2) paratype, off-c entered axial B-f orm. Fig. 3. Freixialina planispiralis Ramalho, holotype, figured by Ramalho (1969), Kimmeridgian-T ithonian,NNw of Freixial, Portugal. Fig. 4. Dhrumella evoluta Redmond, paratype figured by Loeblich and Tappan (1986), Bathonian, Saudi Arabia. Figs 5-6 . Mesoendothyra izu- miana Dain, type specimen figured by Dain (1958), Kimmeridgian, Russia, 5) equatorial section; 6) axial section. Fig. 7. Protopeneroplis striata Weynschenk, holotype, figured by Weynschenk (1950), Middle/ Late Jurassic, Austria. Fig.8. Buccicrenata primitiva BouDagher- Fadel, holotype, equatorial section of micro- spheric form, NHM 66907. Fig. 9- 10. Pseudospirocyclina smouti (Banner), late Callovian- early Oxfordian, 9) holotype figured by Banner (1970), Zakum 1, Upper Araej, Umm Shaif, Persian Gulf showing irregu- lar, sporadic extensions of the hypodermis; 10) oblique vertical section, Lebanon, UCL coll. Figs 11-1 2. Pseudospirocyclina mauretanica Hottinger, types specimens figured by Hottinger (1967), Kimmeridgian, Morocco, 11) axial section; 12) equatorial section. Fig. 14. Pseudospirocyclina muluchensis (Hottinger), types figured by Hottinger (1967), Kimmeridgian, Morocco. Figs 13, 15. Pseudospirocyclina maynci Hottinger, type specimens figured by Hottinger (1967), Kimmeridgian, Morocco, 13) paratype; 15) holotype. Fig. 16. Pseudocyclammina sphaeroidalis Hottinger, type specimen figured by Hottinger (1967), Kimmeridgian, East Morocco (distinguished from P. lituus by simpler alveolar hypodermis and sub- spheroidal test). 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 A B 1 2 3 4 5 6 7 A B 8 9 10 11 12 13 14 15 16 17 18 19 20 21 Plate 4.6 Scale bars: Figs 1-8 , 16 = 0.15mm; Figs 9-1 5, 17- 21 = 0.5mm. Fig. 1. Duotaxis metula Kristan, vertical section showing a conical multi- serial test, Sinemurian, NHM P.66938, Sample L7, Gibraltar, UCL coll. Fig. 2. A). Thaumatoporella?parvovesiculifera (Raineri), B) Siphovalvulina sp., Sinemurian, showing the columellar- siphon, Sample CQ 87,UCL coll. Figs 3-6 . Siphovalvulina colomi BouDagher-F adel et al. 2001, Sinemurian, Apennines, Mt Bove, 3) megalospheric form showing the initial coiled part of the test, Sample MB 63; 4) microspheric form showing the nearly parallel sides in later growth; 5) holotype, vertical sec- tion, NHM P66910, Sample G27; 6)  paratype, transverse section, Sample L6, NHM P66911b. Figs 7-8 . Siphovalvulina gibraltarensis BouDagher- Fadel et al., 2001 Sinemurian; 7) figured holotype, NHMP66912, Sample G8; 8), figured paratypes, NHM P66930, Sample D20. Fig. 9. A) Siphovalvulina beydouni BouDagher- Fadel, holotype, B) Kilianina blancheti Pfender, figured by Noujaim Clark and BouDagher- Fadel (2004), Kesrouane Limestone Formation, Lebanon, UCL coll. Fig.  10. Pseudopfenderina butterlini (Brun), Sinemurian, Sample Ad56g, High Atlas, Jebel Rat Formation, Morocco, UCL coll.. Fig. 11. A specimen sup- posed by Sepfontaine (1967) to be intermediate between Siphovalvulina and Pseudopfenderina, Bathonian, Cevennes, France. Figs 12- 14. Pseudopfenderina butterlini (Brun), type species figured by Hottinger (1967), Sinemurian- Pliensbachian, Morocco. Fig. 15. A reconstruction by Hottinger (1967) of Pseudopfenderina. The columella reconstructed by Hottinger seems to be too elaborate for typical P. butterlini, and the incip- ient tunnels of Pseudopfenderina are not drawn in the reconstruction. Fig.  16. Pseudopfenderina butter- lini (Brun), oblique vertical section figured by Noujaim Clark and BouDagher- Fadel (2004), Bathonian, Kesrouane Limestone Formation, Lebanon, UCL coll. Figs 17- 21. Palaeopfenderina salernitana (Sartonia and Crescenti), Bathonian, Uwainat, Qatar, NHM P44639. Plate 4.7 Scale bars: Figs 1, 2 = 0.15mm; Fig. 3- 5, 12- 13, 20, 21 = 0.5mm; Figs 6- 11, 14 = 0.3mm; Figs 18- 19 = 1mm. Figs 1- 2. Palaeopfenderina salernitana (Sartonia and Crescenti), Bathonian, Uwainat, Qatar, NHM P43712, 1) vertical section; 2) transverse section. Figs 3- 5. Palaeopfenderina trochoidea (Smout and Sugden), 3)  type figure, Bathonian, Uwainat Limestone, Qatar, showing sub-c ameral tunnel and the pil- lared coalescent septal structure of the pfenderid columella; 4) paratype, NHM P43715; 5) paratype, NHM P42967. Figs 6- 7. Pfenderina neocomiensis (Pfender), syntypes of Eorupertia neocomiensis Pfender (1938), Valanginian near Toulon, France, 6)  vertical section; 7)  transverse section. Figs 8- 11 . Conicopfenderina mesojurassica (Maync), 8)  type figure, late Bathonian, Switzerland, figured by Septfontaine (1981); 9)  late Bathonian of Upper Kesrouane Limestone Formation, Lebanon,UCL coll; 10 -1 1) Bathonian, Switzerland, 10) figured by Septfontaine (1981); 11) figured by Septfontaine (1978). Figs 12- 13. Chablaisia chablaisensis (Septfontaine), types figured by Septfontaine (1977), Bathonian?-C allovian, French Pre- Alps. Fig. 14. Pseudoeggerella elongata Septfontaine, type figured by Septfontaine (1988), Bathonian, Pre-A lps, Switzerland. Figs 15- 16. Satorina apuliensis Fourcade and Chorowicz, type figures, Bathonian- Callovian, Yugoslavia, 15) holotype. Fig. 17. Sanderella sp. Section showing bifurcating sub-c ameral tunnel in the flar- ing, rectilinear growth stage, figured by Altiner and Sepfontaine (1979), Callovian, Tauras, Turkey. Figs 18- 20. Steinekella steinekei Redmond, Oxfordian, Tuwaiq Mountain Formation, figured by Redmond (1964), 18- 19) holotype; 20) superficially eroded paratype showing exoskeletal partitions and peripheral chamberlets. Figs 21- 24. Pfenderella arabica Redmond, 21) type species from Middle Jurassic (Bathonian or Callovian), Saudia Arabia; 22-2 4) sketches showing: 22) high trochospiral, test with somewhat inflated chambers; 23) the aperture covered by a finely perforate hemispherical apertural plate; 24) secondary intercameral foramina connected by subcameral tunnel that spirals around the axis of coiling. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 A B 21 22 23 24 25 26 Plate 4.8 Scale bars: Figs 1-2 ,8- 12, 14- 25 = 0.5mm; Figs 3-7 , 13, 26 = 0.3mm. Fig. 1. Conicokurnubia orbi- toliniformis Septfontaine, figured by Septfontaine (1988), Oxfordian-K immeridgian, Pre- Alps, Switzerland. Fig.  2. Conicokurnubia sp., figured by Septfontaine (1981), Kimmeridgian, Turkey. Figs 3- 8. Kurnubia jurassica (Henson), late Oxfordian, Shuqraia Beds, Kurnub, Israel, 3)  NHM P39087; 4)  NHM P39129; 5) NHM P39086; 6) paratype, Jurassic, Israel,NHM M38/4 0;7- 8) Kimmeridgian, Saudi Arabia, UCL coll. Figs 9- 14. Kurnubia palastiniensis Henson, 9)  paratype, NHM M/ 3836; 10)  holotype, revised by Maync (1966), late Oxfordian, Kurnub, Israel, NHM P39089; 11- 12) oblique vertical section of “B- form”, fig- ured by Noujaim Clark and BouDagher- Fadel (2004), Oxfordian- early Kimmeridgian, Bhannes Complex, Lebanon; 13) transverse section of a paratype; 14) figured by Noujaim Clark and BouDagher-F adel (2004), Oxfordian, Kesrouane Limestone Formation, Lebanon. Figs 15-2 1. Kurnubia wellingsi Henson. 15-1 7), fig- ured paratypes, late Oxfordian, Shuqraia Beds, Kurnub, Israel; 15) solid specimen, NHM P39083; 16) trans- verse section, NHM P43718; 17)  vertical section, NHM P43718; 18-1 9) figured by Noujaim Clark and BouDagher- Fadel (2004), Oxfordian, Kesrouane Limestone Formation, Lebanon;18) oblique vertical sec- tions; 19) transverse section; 20-2 1) figured by Hottinger (1967), Oxfordian, Morocco, 20) transverse section; 21) vertical section. Figs 22, 24. Praekurnubia crusei Redmond, figured by Noujaim Clark and BouDagher- Fadel (2004), Oxfordian, Kesrouane Limestone Formation, Lebanon. Fig.  23. A) Praekurnubia crusei Redmond, B) Meyendorffina bathonica Aurouze and Bizon, figured by Noujaim Clark and BouDagher- Fadel (2004), Oxfordian, Kesrouane Limestone Formation, Lebanon. Fig. 25. Textulariopsis sinemurensis BouDagher- Fadel and Bosence, vertical section, holotype, Sample G8, Gibraltar, NHM P66936. Fig. 26. Pseudomarssonella maxima Redmond, late Bathonian-e arly Callovian, paratype, solid specimen, American Museum of Natural History (AMNH) FT- 1270. Plate 4.9 Scale bars:  Figs 1- 6, 11- 14  =  0.5mm; Figs 7- 10  =  0.15mm. Figs 1- 3. Planisepta compressa (Hottinger), 1- 2) types figured by Hottinger (1967), Sinemurian, Morocco, microspheric specimens, 1) hol- otype, 2) paratype; 3) megalospheric specimen from Betics, Gavillan Formation, Southern Spain, UCL coll. Figs 4- 5. Labyrinthina mirabilis Weynschenk, figured by Fourcade and Neumann (1966), Kimmeridgian, Spain, 4)  vertical section; 5)  oblique transverse section. Fig.  6. A) Nautiloculina circularis (Said and Barakat), B) Syriana khouri Fourcade et  al., figured by Noujaim Clark and BouDagher-F adel (2004), Callovian, Lebanon. Figs 7- 8. Nautiloculina oolithca (Mohler), Bajocian- Bathonian, Persian Gulf, NHM coll; 7) oblique axial section, Um Shaif-4 , core, 9961ft, Lower ARAEJ; 8) equatorial section showing double septa, Umm Shaif- 4, core, 9969ft, Lower ARAEJ. Figs 9-1 0. Nautiloculina circularis (Said and Barakat), Callovian, Persian Gulf, NHM coll, Um Shaif-4 , 9) core, 9705ft; 9586ft. Figs 11,13. Nautiloculina cretacea Arnaud- Vanneau and Peybernés, types figures from Arnaud- Vanneau and Peybernés (1978), Berriasian- Aptian, France. Fig.  12. Otbitammina elliptica (d’Archaic), figured by Hottinger (1967), late Bajocian, Chaumont, France. Fig. 14. A) Redmondellina powersi (Redmond), B, D) Nautiloculina oolithica (Mohler), C) Kurnubia wellingsi (Henson), early Kimmeridgian, Lebanon, UCL coll. 1 2 3 A B 4 5 6 7 8 9 10 11 12 B A D 13 C 14 1 2 3 4 5 6 7 8 9 10 11 12 13 A B 14 15 16 17 18 19 Plate 4.10 Scale bars: Figs 1-2  = 0.25mm; Figs 3, 5, 7-1 0, 12-1 3; Figs 4, 11, 14-1 9 = 0.5mm. Fig. 6 = 0.15mm. Fig. 1. Kastamonina abanica Sirrel, type specimens figured by Sirel (1993), Kimmeridgian- Tithonian, Turkey. Fig.  2. Palaeomayncina ternieri (Hottinger), figured by Septfontaine (1988), Sinemurian- Pliensbachian, Swiss Pre- Alps. Figs 3, 5, 6. Pseudocyclammina lituus (Yokoyama), 3) equatorial section with the beginnings of a rectilinear terminal chamber series; this is rare in P. lituus but common in P. vasconica (see Chapter 5), Hauterivian, Saudi Arabia; 5) axial section of a topotype figured by Maync (1959), Kimmeridgian, Japan; 6) Early Cretaceous, Iran, NHM P52300. Fig. 4. Pseudocyclammina kelleri Henson, syntype, Awasil no.5, Ramadi, Iraq, Middle Jurassic, NHM P35968. Figs 7- 10. Pseudocyclammina bukowiensis Cushman and Glazewski, Kimmeridgian, Saudi Arabia, 7) equatorial section; 8) axial section; 8- 9) late Valanginian, upper- most Zangura Formation, NHM P52301; 9) axial section; 10) equatorial section. The thick-w alled, coarsely agglutinated, inner hypodermal alveolae distinguish this form from P. lituus. Fig. 9 Pseudocyclammina vas- conica Maync, Kuwait, Mutriba- 2, cuttings, 12,540ft, MINAGISH- D, Valanginian, Murectilinear growth stages and initial coil in tangential section. Note the smaller test size and less complex alveolar wall than in P. lituus. Note the range of P. vasconica in the Tethyan carbonate shelf seems to be demonstrated to be throughout the Valanginian to Aptian, although it occurs in floods as a local Valanginian index in Saudi Arabia and SE Mid East Gulf. Fig. 11. Spiraloconulus perconigi Alleman and Shroeder, holotype figured by Allemann and Shroeder (1972), Bathonian, Spain. Fig. 12. Kilianina blancheti Pfender, Oxfordian, Upper Kesrouane Limestone, Lebanon, UCL coll. Figs 13, 15- 16. Neokilianina rahonensis Foury and Vincent, 13, 15) types figured, Kimmeridgian, Chaussin, 13) holotype, axial section; 15) transverse section; 16) vertical section, basal Kimmeridgian, Lebanon, UCL coll. Fig.  14. Kilianina preblancheti BouDagher- Fadel and Noujaim Clark, types figured by Noujaim Clark and BouDagher-F adel (2004), Early Callovian;14A) holo- type, axial section; 14B) oblique equatorial section. Figs 17- 18. Parurgonina coelinensis Cuvillier, Foury and Romano, 17) type specimen figured by Cuvillier et al. (1968), Kimmeridgian, Italy; 18) an oblique central section showing the septa breaking in the umbilical region to form pillars, Oxfordian, Lebanon, UCL coll. Fig. 19. Levantinella egyptiensis (Fourcade et al.), oblique equatorial section, Oxfordian, Lebanon, UCL coll. Plate 4.11 Scale bars:  Figs 1- 4, 6, 8- 9, 11- 13, 20  =  0.3mm; Figs 7, 10, 17- 19  =  0.5mm; Figs 5, 15- 16 = 0.15mm; Fig. 14 = 0.25mm. Fig. 1. Paravalvulina sp., Bathonian, Uwainat Formation, United Arab Emirates, NHM P52625. Fig. 2. Pseudomarssonella maxima Redmond, paratype sectioned and figured by Banner et al. (1991), late Bathonian-e arly Callovian, Saudi Arabia, American Museum of Natural History (AMNH) FT- 1270. Fig. 3. Pseudomarssonella plicata Redmond, Bajocian- Bathonian, Persian Gulf, Umm Shail-3 , cuttings, 10,084ft (caved from Lower ARAEJ). Fig. 4. Pseudomarssonella bipartia Redmond grad- ing into P.  inflata Redmond, Callovian, Idd-e l- Shargi-1 , core, 8740ft, Upper Uwainat Formation, United Arab Emirates, NHM coll. Fig.  5. Redmondoides rotundatus (Redmond), paratype figured and sectioned by Banner et  al. (1991), mid/ late Bathonian, Saudi Arabia, AMNH FT- 1293A. Fig.  6. Redmondoides medius (Redmond), note chamber walls straight, perpendicular to the septa, Callovian-O xfordian, Persian Gulf, Umm Shaif- 4, cores, basal Upper Araej, NHM coll. Fig.  7. Redmondoides lugeoni (Septfontaine), metatypic topotypes, axial sectioning showing broad, plate-l ike apertural lips, Bathonian- Callovian, near Chablais, France, NHM P52616. Figs 8-9 . Riyadhoides mcclurei (Redmond), late most Bajocian, 8) para- type, AMNH FT- 1272, 9)  sectioned and figured by Banner et  al. (1991), Saudi Arabia, Dhruma, NHM coll. Figs 10- 11. Riyadhella praeregularis BouDagher- Fadel et  al., 2001, figured types, Sinemurian- Early Pliensbachian, Gibraltar, 10)  holotype, axial section showing thin, convexly curved septa and canalicu- lated test wall, NHM P66947, Sample D20; 11)  paratype, NHM P66916. Fig.  12. Riyhadella regularis Redmond, mid/ late Bathonian, paratype of the synonymous R. nana Redmond, figured by Banner et al. (1991), Aramco well, Saudi Arabia. Fig. 13. Riyadhella sp., Callovian, Upper Uweinat, Persian Gulf, NHM coll. Fig. 14. Cymbriaella lorigae Fugagnoli, figured by Fugagnoli (1999), Sinemurian- Pliensbachian, Italy. Fig. 15. A sketch showing Valvulina sp. and the tooth- like aperture from Banner’s collection, UCL. Figs 16-1 7. Gutnicella cayeuxi (Lucas), 16)  type specimen figured by Lucas (1939), Aalenian, Algeria; 17) fig- ured by Gutnic and Moullade (1967), Aalenian, West Taurus, Turkey. Figs 18- 20. Haurania deserta Henson, Bathonian, 18) paratype from Muhaiwir Formation, NHM P35859; 19) paratype, Jurassic, Wadi Amij well, West Iraq M/ 3846, NHM P35863; 20) paratype, transverse thin section, NHM P3856. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 Plate 4.12 Scale bars:  Figs 1-3 , 5, 7, 9, 14, 17-2 0  =  0.25mm; Fig.  4  =  1mm; Figs 6, 8, 11  =  0.15mm; Figs 10, 12-1 3; Fig. 15 = 3mm; Fig. 16 = 2mm. Figs 1- 4. Haurania deserta Henson, 1948, 1- 2) figured by BouDagher- Fadel and Bosence (2007), Middle Jurassic, 1)  Betics, Gavillan Formation, Southern Spain, Sample RA- 01- 203; 2) High Atlas, Jebel Rat Formation, Morocco; 3) longitudinal section, Toarcian, Yemen, NHM P53894; 4) vertical section figured by Hottinger (1967), Morocco. Fig. 5- 6 Meyendorffina bathonica Arouze and Bizon, 5) figured by Furrer and Septfontaine (1977), late Bathonian, Swiss Pre- Alps; 6) fig- ured by Noujaim Clark and BouDagher- Fadel (2004), late Bathonian, Lebanon. Fig. 7. Gutnicella oxitan- ica (Peybernes), late Bathonian- late Callovian, Upper Kessrouane Limestone, Lebanon, UCL coll. Fig. 8. Gutnicella bizonorum (Bourrouilf and Moullade), figured paratype, Bathonian, Minorca. Fig. 9. Gutnicella minoricensis (Bourrouilh and Moullade), type figured, Bathonian, Algeria. Figs 10- 11. Socotraina serpentina Banner et al., Early Jurassic, Socotra, Yemen, 10) holotype, microspheric specimen, vertical section, NHM P53883; 11)  paratype, transverse section, NHM P53892. Fig.  12. Platyhaurania subcompressa Bassoullet and Boutakiout, type specimen figured by Bassoullet and Boutakiout (1996), late Sinemurian, Morocco. Figs 13- 16. Timidonella sarda Bassoulet, Chabrier and Fourcade, type figures, Bajocian-B athonian, Sardinia, 13-1 4) A- forms; 15- 16) B-f orms. Fig. 17. Trochamijiella gollesstanehi Athersuch et al., type specimens fig- ured, Bathonian, Oman. Fig. 18. Rectocyclammina chouberti Hottinger, type specimen figured by Hottinger (1967), early Kimmeridgian, Morocco. Fig. 19. Meyendorffina sp., Bathonian, France, UCL coll. Fig. 20. Alveosepta jaccardi (Schrodt), equatorial section of a B-f orm, early Kimmeridgian, Qatar, Dukhan-5 1, 74441/ 2ft, NHM coll. Plate 4.13 Scale bars: Figs 1-8  = 0.5mm Figs 1- 3. Alveosepta jaccardi (Schrodt), late Oxfordian, 1) topo- type figured by Hottinger (1967), Switzerland; 2-3 ) figured by Noujaim Clark and BouDagher-F adel (2002) from Bhannes Complex, Lebanon, 2) axial section; 3) oblique equatorial section. Figs 4- 5. Redmondellina powersi (Redmond), Early Kimmeridgian, figured by Noujaim Clark and BouDagher-F adel (2002) from Bhannes Complex, Lebanon, 4) oblique equatorial section showing the “clear line” in septa, and median lamella cut lengthwise in last chamber; 5) transverse section of a microspheric form showing pillar- like hypo- dermal extensions, UCL coll. Figs 6-8 . Alzonella cuvillier Bernier and Neumann, type figures, 6) showing partitions “beams” which subdivide the chamberlets, Bathonian, Alzon, France; 7) Bathonian, Mas- del-P ont Well, France; 8) Callovian, Uweinat, Umm Shaif- 3, Persian Gulf, NHM coll. 1 2 3 4 6 5 8 7 1 2 4 3 5 6 7 Plate 4.14 Scale bars: Figs 1-6  = 1mm. Figs 1-6 . Alzonorbitopsella arabia BouDagher- Fadel, 5) BP 6626; 6,9,10) BP 6627, from 9879 ¼ ft; 7, 8) BP 6623, from 9880 ft; Um-S haif-4 core; Upper Bathonian, basal Uweinat Formation, Abu Dhabi, NHM coll.. 1 2 3 4 5 11 6 7 10 12 8 9 13 Plate 4.15 Scale bars:  Figs 1, 7- 13  =  0.5mm; Figs 2, 5  =  0.3mm; Figs 3-4 =0.15mm; Fig.  6  =  1mm. Fig.  1. Palaeocyclammina complanata Bassoullet, Boutakiout and Echarfaoui, holotype F62025, fig- ured by Bassoullet et al. (1999), Pliensbachian, Morocco. Figs 2-4 . Amijiella amiji (Henson),, type speci- mens, Bathonian, Muhaiwir Formation, Wadi Amij well, West Iraq, 2)  solid specimen, holotype, NHM P35869; 3- 4) paratypes, 3)  transverse section, 4)  axial section, NHM P35866. Fig.  5. Torinosuella pen- eropliformis (Yabe and Hanzawa), Thamama Formation, Oman, Umm Shaif- 3, cuttings, 7255ft, NHM coll. Figs 6- 7. Anchispirocyclina lusitanica (Egger), 6)  figured by Ramalho (1971), equatorial section of B- form, Kimmeridgian, Portugal; 7)  figured by Hottinger (1967), Kimmeridgian, Morocco. Figs 8-9 . Anchispirocyclina praelusitanica (Maync), near topotypes, B- forms figured by Hottinger (1967), Oxfordian, Israel (note: Hottinger referred praelusitanica to Alveosepta (=Redmondellina powersi), but praelusitanica lacks the septal “clear line” of Redmondellina and has pillars spread throughout the whole chamber lumen, not merely in the median plane). Figs 10-1 1. Bostia irregularis Bassoullet, type figured by Bassoullet et al. (1999), late Bathonian, France, 10) holotype, axial section; 11) paratype, axial section. Figs 12- 13. Ijdranella altasica Bassoullet et al., type specimens figured by Bassoullet et al. (1999), Pliensbachian, Morocco, 12) hol- otype, equatorial section; 13) oblique vertical section. Plate 4.16 Scale bars:  Figs 1- 6, 8, 10  =  0.5mm; Figs 7, 9, 11mm. Figs 1- 2. Bosniella oenensis Gušiç, Sinemurian- Pliensbachian, Jebel Rat Formation, southern High Atlas, Morocco, UCL coll. Fig.  3. Everticyclammina praevirguliana Fugagnoli, Sinemurian-P liensbachian, Gibraltar, UCL coll.. Fig.  4. Everticyclammina kelleri (Henson), paralectotype, equatorial section, Berriasian-V alanginian, Zangura Formation, Iraq, NHM P35968. Fig.  5. Everticyclammina greigi (Henson), paratype, basal Cretaceous, Qatar, NHM P35795. Fig. 6 . Everticyclammina virguliana (Koechlin), Kimmeridgian, Persian Gulf, Zakum- 1, core, 9536ft, NHM coll. Fig.  7. Streptocyclammina parvula (Hottinger), figured by Hottinger (1967), Kimmeridgian, Morocco. Figs 8-9 . Lituosepta recoarensis (Cati), 8) microspheric form figured by Hottinger (1967), Middle Jurassic, Morocco; 9)  megalospheric form, Betics, Gavillan Formation, Spain. Fig.  10. Orbitopsella primaeva (Henson, 1948), microspheric forms, Betics, Gavillan Formation, Spain,UCL coll.. Fig. 11. Karaisella uzbekistanica Kurbatov, figured from Kurbatov (1971), Oxfordian, Uzbekistan. 1 2 5 3 7 4 6 8 9 10 11 1 2 3 4 7 5 6 8 9 Plate 4.17 Scale bars: Fig. 1 = 1mm; Figs 2-3  = 0.3mm; Figs 4- 9 = 0.5mm. Fig. 1. Orbitopsella dubari Hottinger, megalospheric form, figured by Hottinger (1967), Middle Jurassic, Morocco. Fig. 2- 3. Orbitopsella praecursor (Gümbel), syntypes, Sinemurian, Jabal Milaha, Oman, NHM P35780. Figs 4- 6 . Orbitopsella primaeva, (Henson), syntypes, Middle Jurassic, Jabal Milhaha Oman, Davies collection, 4) NHM 35791; 5) NHM P35789; 6) NHM 35788. Fig. 7. Lituolipora polymorpha Gŭsić and Velić, figured by Gŭsić and Velić (1978), Early Jurassic, Yugoslavia. Fig. 8-9 . Cyclorbitopsella tibetica Cherchi, Schroeder and Zhang, Pliensbachian, 8)  holotype, figured by Cherchi, Schroeder and Zhang (1984), Pupuga Formation, South Tibet; 9) Betics, Gavillan Formation, Spain, UCL coll 1 2 3 4 5 6 7 8 9 11 10 12 Plate 4.18 Scale bars:  Figs 1- 12  =  0.5mm. Figs 1- 2. Cayeuxia?piae Frollo, Sinemurian- Pliensbachian, 1)  Lebanon, UCL coll.; 2)  figured by BouDagher- Fadel and Bosence (2007), High Atlas, Jebel Rat Formation, Morocco. Figs 3-4 , 9.  Thaumatoporella?parvovesiculifera (Raineri), Sinemurian, 3)  Betics, Gavillan Formation, Spain, figured by BouDagher-F adel and Bosence (2007); 4)  Gibraltar Limestone Formation, NHM P66949; 9)  sample MB 49, UCL coll. Figs 5-7 . Palaeodasycladus?mediterraneus (Pia), figured by BouDagher- Fadel et al. (2001), Sinemurian, Gibraltar Limestone Formation 5) NHM P66945; 6- 7) NHM P66931- 32. Fig. 8. Ammonites, Pliensbachian, Italy, UCL coll. Fig. 10. Orbitopsella spp., Early Jurassic, Pakistan, UCL coll. Fig. 11. Clypeina jurassica Favre, early Kimmeridgian, Kesrouane Limestone, UCL coll. Fig. 12. Cladocoropsis mirabilis Felix, Kimmeridgian, Greece,UCL coll. 285 Chapter 5 The Mesozoic Larger Benthic Foraminifera: The Cretaceous 5.1 Introduction Although more than a third of the larger benthic foraminifera became extinct towards the end of the Tithonian, the complex and partitioned lituolids of the order Textulariida continued to dominate the biota of the inner carbonate platform environments that were widespread along the rifted western and eastern margins of the Cretaceous Tethys. Many of the Cretaceous lituolid species have quite limited stratigraphic ranges and are valuable as index fossils. They were joined by a new group, which made its first appearance in the late Early Cretaceous, the alveolinids. These were large mili- olides that showed spectacular expansion in the Middle Cretaceous, proliferating in mid- latitudes, and often becoming annular or discoid and subdivided by partitions. Many of them resembled the planispiral-f usiform fusulinides (see Chapter 2) of the Permian, attaining approximately the same range of sizes, but differing fundamentally in their imperforate, porcelaneous wall structure. In the Late Cretaceous new, simple rotaliides evolved into forms with complicated three-l ayered structures, the orbitoids. While the previous two groups had their main breeding ground in the Tethyan realm, the orbitoids showed provincialism and some were only found in the Caribbean. Cretaceous larger benthic foraminifera have been closely studied by many work- ers (for example, Maync, 1952; Brönnimann, 1954a, b, 1958; Schroeder, 1962, 1964, 1975; Banner, 1970; Bergquist, 1971; Decrouez and Moullade, 1974; van Gorsel, 1975, 1978; Caus, 1988; Hottinger and Drobne, 1989; Hottinger et al., 1989; Hughes, 2004; Kaminski, 2010; Matsumaru, 1991; Özcan, 1995; Görög and Arnaud Vanneau, 1996; Noujaim Clark and Boudagher- Fadel, 2001; BouDagher- Fadel and Lord, 2001; Schroeder et al., 2010; Albrich et al., 2014; Scherreiks et al., 2015a, 2015b; BouDagher- Fadel et al., 2015; Sun et al., 2015; Schlagintweit et al., 2016, BouDagher- Fadel., 2008, 2015; BouDagher-F adel and Price, 2017; Schlagintweit and Rashidi, 2017). In this chapter, the taxonomy of the major Cretaceous larger benthic foraminifera is summa- rized, and then their biostratigraphic, paleoenvironmental and paleogeographic signif- icance is discussed. 5.2 Morphology and Taxonomy of Cretaceous Larger Benthic Foraminifera As in the Jurassic, the dominant larger foraminifera of the Early Cretaceous were the agglutinated Textulariida. The miliolides became abundant in the Middle Cretaceous, but it was not before the Cenomanian that larger miliolides played an important role in the benthic assemblages of carbonate platforms. The rotaliides became common in the Late Cretaceous. 286 Evolution and Geological Significance of Larger Benthic Foraminifera Figure 5.1. The evolution of the Cretaceous orders (thick lines) and superfamilies (thin lines) of the larger benthic foraminifera. In this section, the following four orders will be discussed: • Involutinida • Textulariida • Miliolida • Rotaliida The development and evolution of the superfamilies of these orders is schematically shown in Fig. 5.1. Below are given the morphological characteristics and taxonomic relationships of these major Cretaceous forms. In this chapter only the foraminifera which appeared in the Cretaceous are dealt with, those which made their first appear- ance in the Jurassic have been described in Chapter 4. ORDER INVOLUTINIDA Hohenegger and Piller, 1977 This order includes all forms with an enrolled second chamber. They have walls that are aragonitic, but commonly they are recrystallised to give a homogenous microgranular structure. The umbilical region has pillar- like structures on one or both sides of the test. They range from Triassic to Late Cretaceous (Cenomanian). The Mesozoic Larger Benthic Foraminifera: The Cretaceous 287 Superfamily INVOLUTINOIDEA Bütschi, 1880 This superfamily consists of forms with a first chamber that is followed by a plani- spiral to trochospiral enrolled tubular second chamber. Triassic to Late Cretaceous (Cenomanian). All genera to be found in the Cretaceous originated in the Jurassic or earlier, and have been described in Chapter 4. ORDER TEXTULARIIDA Delage and Hérouard, 1896 The tests of these agglutinated foraminifera are made of foreign particles bound by organic cement. They range from Early Cambrian to Holocene. Superfamily NEZZAZATOIDEA Hamaoui and Saint-M arc, 1970 The test is trochospiral or planispiral, later occasionally uncoiled. The wall is s imple, non- lamellar, and microgranular, and may possess internal plates or simple partitions (Fig. 5.2). The aperture is simple or multiple. Middle Jurassic (Bajocian) to Holocene. Family Nezzazatidae Hamaoui and Saint-M arc, 1970 The test is trochospiral or planispiral, later it may be uncoiled. The interior of each chamber has an internal plate and may be basally digitate. The aperture is single or multiple with an internal tooth plate. Cretaceous (Barremian to Maastrichtian). Subfamily Nezzazatinae Hamaoui and Saint-M arc, 1970 The test is trochospiral or planispiral, later it may be uncoiled. The interior of each chamber has an internal plate and is basally digitate. The aperture is single or multiple. Cretaceous (Barremian to Turonian) (see Fig. 5.2). • Nezzazata Omara, 1956 (Type species: Nezzazata simplex Omara, 1956). The test is trochospiral with a closed umbilicus. A narrow plate extends backwards from one septum to the previous one. The aperture is single, extending from the umbilicus to the periphery, then bending sharply with an apertural tooth. Cretaceous (Albian to Turonian) (Plate 5.1, fig. 12). • Nezzazatinella Darmoian, 1976 (Type species: Nezzazatinella adhami Darmoian, 1976). The test is a low trochospiral, with a flattened spiral side, and the opposite side convex and involute. The chambers in the final whorl are narrow, elongate, with the final chamber flaring backwards. The aperture is a large curved slit, and may be accompanied by secondary apertures either in a curved row or s cattered over the apertural face. Cretaceous (Barremian to Turonian) (Plate 5.1, figs 10-1 1). • Lupertosinnia Farinacci, 1996 (Type species:  Lupertosinnia pallinii Farinacci, 1996). The test is conical and trochospirally enrolled. The central portion of the test is occupied by alveolar thin layers, communicating with the chambers by perpendicular canals. The wall is simple and finely agglutinated. Cretaceous (Cenomanian). 288 Evolution and Geological Significance of Larger Benthic Foraminifera Figure 5.2. Key morphological characteristics of the Lituoloidea, the Biokovinoidea, the Nezzazatoidea as compared to the Haplophragmioidea (representatives of the small benthic foraminifera, see Loeblich and Tappan, 1988). Lituoloidea Planispiral, Planispiral with Aperture multiple, Lituola sometimes solid, areal; test sometimes with last few non- alveolar, uncoiling chambers non-c analiculate uncoiling; walls and coiled test no internal nautiliform toothplates Biokovinoidea to globose Canaliculate Aperture cribrate at Debarina walls base of apertural face Haplophragmioidea Aperture a basal slit Haplophragmoides Nezzazatoidea Planispiral, Aperture single, rounded, central Daxia involute or interiomarginal basal equatorial slit Stomatostoecha nearly so; (basal) biumbilicate, Aperture single, in middle of apertural Friexialina compressed; areal face many (15+) slit at top of apertural Phenacophragma chambers face per wholrl, Aperture test thickness/d iameter Mayncina usually multiple approximately 1/ 3, rapidly apertures scattered increasing in over apertural face height test thickness/ diameter Gendrotella approximately 1/ 7, apertures in vertical row in face Subfamily Coxitinae Hamaoui and Saint- Marc, 1970 The test is trochospiral or planispiral with an internal median plane between consec- utive septa, the plates being terminally bifurcate or digitate. Apertures are multiple. Cretaceous (Cenomanian to Maastrichtian). • Coxites Smout, 1956 (Type species:  Coxites zubairensis Smout, 1956). The test is trochospiral with a bilocular embryonal stage, followed by numerous low chambers, that are oblique on the spiral side and strongly arched on the umbilical side around a central depression. There are incomplete partitions within the chambers parallel to the septa. The aperture is multiple. Cretaceous (late Cenomanian to early Turonian) (Plate 5.2, figs 6- 7). • Rabanitina Smout, 1956 (Type species:  Rabanitina basraenesis Smout, 1956). The test is planoconvex, trochospiral to completely involute. A complex perforated plate within the adult chamber parallels the spiral wall, but twists and is attached to both chamber floor and roof. The aperture is interiomarginal and multiple. Cretaceous (early Cenomanian) (Plate 5.4, fig. 3). The Mesozoic Larger Benthic Foraminifera: The Cretaceous 289 • Demirina Özcan, 1994 (Type species: Demirina meridionalis Özcan, 1994). The test is biumbilicate, peneropliform with strongly thickened low and broad adult chambers, and has a simple inner structure with one set of partitions. Cretaceous (Cenomanian) (Plate 5.3, fig. 4). Family Mayncinidae Loeblich and Tappan, 1985 The test is biumbilicate or very compressed, planispirally enrolled, but rarely uncoiling. The chambers increase rapidly in height. The septa are monolamellar. The wall is simple, solid and microgranular. Late Jurassic to Late Cretaceous (Kimmeridgian to Santonian). • Gendrotella Maync, 1972 (Type species: Choffatella rugoretis Gendrot, 1968). The aperture is multiple in vertical rows. The absence of an alveolar hypodermis dif- ferentiates this form from homeomorph such as Choffatella or Torinosuella. Late Cretaceous (Santonian) (Plate 5.4, fig. 9). • Mayncina Meumann 1965. (Type species: Daxia orbignyi Cuvillier and Szakall, 1949). The aperture is multiple and scattered. Cretaceous (Valanginian to Cenomanian) (Plate 5.4, fig. 8). • Phenacophragma Applin, Loeblich and Tappan, 1950 (Type species: Phenacophragma assurgens Applin, Loeblich and Tappan, 1950). Similar to Freixialina (see Chapter 4) but with an aperture at the top of apertural face. Cretaceous Early (Albian). • Stomatostoecha Loeblich and Tappan, 1950 (Type species:  Stomatostoecha plum- merae Applin et al., 1950). This form is similar to Daxia (see Chapter 4) but with an extensive slit-l ike aperture (close to the rounded, basal aperture of Daxia). Early Cretaceous (Albian) (Fig. 5.3; Plate 5.4, fig. 10). Superfamily BIOKOVINOIDEA Gušiç, 1977 Members of this superfamily have a free test with a trochospiral or planispiral early stage, that later may become uncoiled. Septa are homogeneous and massive. The aper- ture is basal to areal, and single to multiple (Fig. 5.2). Early Jurassic (Sinemurian) to Late Cretaceous (Maastrichtian). Family Biokovinidae Gušiç, 1977 The test is planispirally coiled, later it may uncoil. Endoskeletal pillars may be present. Walls are canaliculate with alveoles that open both to the exterior or interior. Early Jurassic to Cretaceous. • Zagrosella Schlagintweit and Rashidi, 2017 (Type species:  Zagrosella rigaudii Schlagintweit and Rashidi, 2017). The test has an oscillating coiling plane in the early stage, later it is planispiral, and finally may be rectilinear. Few and irregularly distributed endoskeletal pillars are present in the chamber interior. Cretaceous (late Maastrichtian). Family Charentiidae Loeblich and Tappan, 1985 Early stage tests are planispiral. Walls are finely canaliculated. Apertures are single or multiple. Middle Jurassic (Callovian) to Late Cretaceous (Maastrichtian). • Charentia Neumann, 1965 (Type species:  Charentia cuyillieri Neumann, 1965). Planispiral throughout with a single aperture, becoming enlarged with growth. 290 Evolution and Geological Significance of Larger Benthic Foraminifera A B C D E F Figure  5.3. A, D) axial section/o blique axial section; B, E) equatorial sections; C, F) part of same equatorial section (scale bar  =  0.5mm). A- C are Stomatostoecha sp. from the Santonian of France s howing gross “choffatelloid” form of tests, chambers and apertures, but lack of a labyrinthic or alveolar hypodermis of D- F Choffatella decipiens Schlumberger, from the Aptian of the Persian Gulf. Note: The original description of Stomatostoecha is misleading. It has a single, basal, slit- like aperture only and no areal apertures. The finely canaliculate wall differs from the alveolar wall of Hemicyclammina and from Everticyclammina in that the alveoles are thicker, more separately spaced, branch and dichotomize. Early Cretaceous (Barremian) to earliest Late Cretaceous (Cenomanian) (Plate 5.5, fig. 12). • Debarina Fourcade, Raoult and Vila, 1972 (Type species:  Debarina hahounerensis Fourcade et al., 1972). Planispiral in later whorls at least. Apertures are multiple, cribrate, and basal. Debarina differs from Charentia by its cribrate aperture and pos- sible intitial streptospirality. Early Cretaceous (late Barremian to Aptian) (Plate 5.5, fig. 13). Superfamily LITUOLOIDEA de Blainville, 1825 Members of this superfamily have multilocular, rectilinear and uniserial tests. The wall has no alveolar or reticulate hypodermis. The early stage has plani- (strepto- ) or tro- chospiral coiling. The periphery of the chambers have radial partitions; centrally with The Mesozoic Larger Benthic Foraminifera: The Cretaceous 291 or without scattered, separated pillars (Fig. 5.2). The aperture is simple or multiple cribrate, with no tooth plate. Early Jurassic (Sinemurian) to Holocene. Family Hauraniidae Septfontaine, 1988 Tests may be uncoiled, uniserial or planispirally coiled. Walls are microgranular, with a hypodermic network. Septa are simple or with complicated microstructures. The interiors of chambers are simple or with pillars. Apertures are multiple. Jurassic (late Sinemurian) to Cretaceous. Subfamily Amijiellinae Septfontaine, 1988 Members of this subfamily have an uncoiled or planispirally coiled test. The septa are simple or with a complicated microstructure. The interiors of the chambers are simple, however some genera may develop pillars. Walls may have alveoles. The aperture is multiple. Jurassic (late Sinemurian) to Cretaceous. • Spirocyclina Munier-C halmas, 1887 (Type species:  Spirocyclina choffati Munier- Chalmas, 1887). Tests are flat coils, becoming peneropliform. There is a coarse early epidermis, with rapid development of subepidermal alveolae. Central pillars are few and scattered. Late Cretaceous (late Cenomanian to Santonian) (Plate 5.3, Fig. 4). • Martiguesia Maync, 1959 (Type species:  Martiguesia cyclamminiformis Maync, 1959). Coiled, nautiliform, with a central zone having delicate pillars bridging from septum to septum. Late Cretaceous (Santonian) (Plate 5.4, fig. 4). • Montsechiana Aubert, Coustau and Gendrot, 1963 (Type species:  Montsechiana martiguae Aubert et  al., 1963). Strong central pillars span flabelliform chambers. There is a finely reticulate hypodermal mesh, with beams below. The aperture is c ribrate. Early to Late Cretaceous (Valanginian to Santonian) (Plate 5.2, fig. 1). • Qataria Henson, 1948 (Type species: Qataria dukhani Henson,1948). The early stage is planispiral, later chambers are cyclical. The outer parts of the chambers are sub- divided into numerous small chamberlets. Late Cretaceous (late Cenomanian to Turonian) (Plate 5.4, figs 5- 6). • Sornayina Marié, 1960 (Type species:  Sornayina foissacensis Marie, 1960). The wall has a subepidermal meshwork. Chambers are subdivided into small chamber- lets by bifurcating septula perpendicular to the septa. Late Cretaceous (Coniacian) (Plate 5.5, Fig. 11). Family Everticyclamminidae Septfontaine, 1988 Tests are streptospiral or planispiral in the early stage, uncoiled in the adult or unise- rial throughout. Walls are microgranular, with an alveolar microstructure. Interiors of chambers are simple. Apertures are areal. Early Jurassic (Sinemurian) to Cretaceous (early Cenomanian). • Hemicyclammina Maync, 1953 (Hemicyclammina sigali Maync, 1953). Tests are pla- nispiral, with a single aperture occupying all, or nearly all, of the total height of the apertural face in equatorial section, reducing the solid septa which are clearly differ- ent in structure from spiral wall. The basal layer is deposited over the chamber floors. Cretaceous (late Aptian to early Cenomanian) (Fig. 5.4; Plate 5.6, figs 12- 15). 292 Evolution and Geological Significance of Larger Benthic Foraminifera Much reduced later septa Strong basal layer Single aperture occupying nearly all of the total height of the apertural face Straight septa Figure  5.4 Equatorial section of an advanced form of Hemicyclammina sigali Maync from the Early Cenomanian of the Persian Gulf. Scale bar = 0.25mm. • Hottingerita Loeblich and Tappan, 1985 (Type species: Mesoendothyra complanata Hottinger, 1967). Tests are evolute, biumbilicate, loosely coiled in a low, almost pla- nar streptospire. It differs from Alveosepta (see Chapter 4) by having a single aper- ture rather than a multiple one, and differs from Mesoendothyra (see Chapter 4) by having complex septa and smaller and more numerous chambers, and from both genera in being evolute. Early Cretaceous (Barremian) (Plate 5.2, fig. 2). Family Lituolidae de Blainville, 1827 The early stages of the tests are enrolled, but later they may become rectilinear. Walls are made from agglutinated foreign particles. There are few chambers (less than 10) per whorl. Carboniferous to Holocene. Subfamily Ammomarginulininae Podobina, 1978 Test have early stages that are coiled, but later uncoiling. Apertures are single. Carboniferous (Early Mississippian) to Holocene. • Ammobaculites Cushman, 1910 (Type species:  Spirolina agglutinans d’Orbigny, 1846). The test is simple, not compressed and uncoils in the adult. Aperture is single, areal. Carboniferous (Mississippian) to Holocene (Plate 5.6, fig. 18). Subfamily Lituolinae de Blainville, 1827 Members differs from the Ammomarginulininae in having a multiple apertures. Late Triassic to Holocene. The Mesozoic Larger Benthic Foraminifera: The Cretaceous 293 • Lituola Lamarck, 1804 (Type species: Lituolites nautiloidea Lamarck, 1804). These forms have no internal partitions and a multiple cribrate aperture. Late Triassic to Holocene (Plate 5.5, fig. 14; Plate 5.6, fig. 9). Superfamily PFENDERINOIDEA Smout and Sugden, 1962 Members of this superfamily have a trochospiral test throughout, or may become uncoiled. Some forms have a siphonal canal, others develop a central composite colu- mella with pillars between the apertural plates and septa. Early Jurassic (Hettangian) to Cretaceous. Family Pfenderenidae Smout & Sugden, 1962 Members of this subfamily have a loose trochospiral test with siphonal canals that con- nect successive apertures in primitive forms. Some forms develop a central composite columella, composed of thickened innermost septal ends (“septal buttons”) with or without additional pillars and a spiral canal between the columella and the thickened septa. A subcameral tunnel (either simple or multiple) is present in advanced forms. The chamber interior of advanced taxa is subdivided by vertical or horizontal (or both) exoskeletal partitions, resulting in a reticulate subepidermal layer. The apertures are always cribrate, areal. Early Jurassic (Sinemurian) to Late Cretaceous (Maastrichtian). Subfamily Kurnubiinae Redmond, 1964 Members of this family have a test without subcameral tunnels, and may or may not have a solid core. The peripheral zone is divided by radial partitions. Late Cretaceous (Maastrichtian). • Gyroconulina Schroeder and Darmoian, 1977 (Type species: Gyroconulina columel- lifera Schroeder and Darmoian, 1977). The early stage is conical with a globular pro- loculus, followed by a cylindrical stage. The proloculus is globular and followed by chambers subdivided by vertical beams and horizontal rafters, both perpendicular to the outer surface of the test, forming a honeycomb, subepidermal network. The central zone has irregularly distributed pillars. Late Cretaceous (late Maastrichtian) (Plate 5.2, fig. 5). Subfamily Pfenderininae Smout and Sudgen, 1962 Members of this subfamily have a test with a single subcameral tunnel that is always buried in the columella, which is formed by pillars and calcitic deposits. No peripheral partitions or chamberlets are present. The aperture is multiple, cribrate, and on the apertural plate. Early Cretaceous (Valanginian to Barremian or?Aptian). • Banatia Schlagintweit and Bucur, 2017 (Type species: Banatia aninensis Schlagintweit and Bucur, 2017). The test is low, asymmetric conical to almost planispiral, formed by a short low trochospiral followed by flaring chambers. The marginal part of the chambers is undivided. The axial region has pillars that are continuous between successive chambers, grading into a system of numerous fine endoskeletal horizon- tal plates with irregularly arranged fine vertical elements (pillars or buttresses). In transverse sections this zone displays a reticulate-l abyrinthic network. Cretaceous (late Barremian). 294 Evolution and Geological Significance of Larger Benthic Foraminifera Superfamily LOFTUSIOIDEA Brady, 1884 The test is planispiral, but may uncoil in later stages. The wall is agglutinated with a dif- ferentiated outer layer and an inner alveolar layer. Late Triassic (Carnian) to Holocene. Family Choffatellidae Maync, 1958 This family is characterised by having hypodermal alveoles. The tests are planispiral, lacking continuously developed endoskeletal pillars. Walls are finely and complexly alveolar. The septa have many fine apertures. Early Jurassic (Pliensbachian) to Late Cretaceous (Maastrichtian). • Balkhania Mamontova, 1966 (Type species:  Balkhania balkhanica Mamontova, 1966). Annular test with simplified, thin septa. Early Cretaceous (Barremian to Aptian) (Plate 5.7, fig. 12). • Broeckinella Henson, 1948 (Type species: Broeckinella arabica Henson, 1948). The test is flabelliform, with a large undivided globular proloculus followed by a planispi- ral evolute stage. There are subepidermal vertical and transverse partitions resulting in a polygonal meshwork near the interior margins of the chambers. Late Cretaceous (Santonian to Maastrichtian) (Plate 5.7, fig. 13). • Pseudochoffatella Deloffre, 1961 (Type species: Pseudochoffatella cuvillieri Deloffre, 1961). Flabelliform test with high chambers and septa that are as complex and thick as the hypodermis of the wall. It has a more coarsely agglutinated epidermis and less delicately alveolar epidermis than in the more choffatelloid B- forms (see Choffatella). The septa of Pseudochoffatella are much thicker than their narrow “tubular” aper- tures, whereas in Balkhania the septa are thinner than the broad apertures. Early Cretaceous (Aptian to Albian) (Fig. 5.5; Plate 5.3, 13; Plate 5.5, fig. 10). Family Loftusiidae Brady, 1884 The test is globose to fusiform with septula perpendicular to septa, and parallel to the coiling direction. An alveolar hypodermis is weakly or fully developed. A subepider- mal reticulate mesh is absent in the septa. Late Cretaceous (Maastrichtian) to?Eocene. The following genera have a globose, spherical test with interseptal horizontal lamel- lae, and a very weakly developed alveolar hypodermis: • Feurtillia Maync, 1958 (Type species: Feurtillia frequens Maync, 1958). This form is characterised by thick septa and a strong apertural neck even in small, wholly coiled specimens. The wall has a distinct subepidermal network of shallow alveoles that appear polygonal in tangential section. Late Jurassic to Early Cretaceous (Tithonian to Valanginian) (Plate 5.3, fig. 8) • Praereticulinella Deloffre and Hamaoui, 1970 (Type species: Praereticulinella cuvil- lieri Deloffre and Hamaoui, 1970). Septula aligned from chamber to chamber, but without a basal pre- septal canal. There is a pre- septal tunnel at the top of the sep- tum. Early Cretaceous (Barremian) (Plate 5.4, figs 1- 2). • Reticulinella Cuvillier, Bonnefous, Hamaoui and Tixier, 1970 (Type species: Reticulina reicheli Cuvillier et al., 1969). Secondary septula are intercalated between primary septula in later chambers and pre- septal canals are present at the base of the septum. Late Cretaceous (Cenomanian to Maastrichtian) (Plate 5.4, fig. 3). The Mesozoic Larger Benthic Foraminifera: The Cretaceous 295 Alveolar hypodermis B Rectilinear stage reticulate hypodermis A Deuteroconch Protoconch Figure 5.5. A) Spiroconulus perconigi Allemann and Schroeder, holotype figured by Allemann and Schroeder (1972), the Bathonian of Cadiz, South Spain; B) Pseudochoffatella cuvillieri Deloffre, latest Aptian to earliest Albian, North Croatia, after Guşič (1975), transverse section of a microspheric form (B-f orm). Note: The more coarsely agglutinated epidermis and less delicately alveolar epidermis than in other megalospheric choffatelloid (A- forms) (e.g. Plate 5.5, fig. 10) While a fusiform test with interseptal pillars and a fully developed alveolar hypoder- mis (but with no pre- septal canal or tunnel) is shown by: • Loftusia Brady, 1870 (Type species: Loftusia persica Brady, in Carpenter and Brady, 1870). The test is irregularly fusiform or cylindrical in shape and may have either rounded or pointed ends. It is made of a series of chambers that are planispirally coiled around an elongate axis. The front end of each chamber slopes gently down to meet the previous whorl. The portion behind the frontal slope merges smoothly with the previous chamber at the point where its own frontal slope merges smoothly into the area of equal elevation. There is, thus, no part of the exterior wall possess- ing a configuration that would set it off as an apertural face in the normal sense, although there are a number of pores in the surface of the chamber adjacent to the previous whorl. The interiors of the chambers are partially filled by networks of irregular projections that cover the inner surface of the exterior walls, some of the projections reach the surface of the preceding whorl thus forming interseptal pillars or partial septa, depending on their shape. The spiral wall is made up of a thin outer calcareous layer and a thick arenaceous layer, the forward extension of the latter forms the gently- inclined frontal slope of each chamber. Late Cretaceous (Maastrichtian) to?Eocene (Fig. 5.6; Plate 5.8, figs 1- 8; Plate 5.9, figs 6- 7, with fig. 7 296 Evolution and Geological Significance of Larger Benthic Foraminifera Figure 5.6. A) Oblique equatorial section of Loftusia persica Brady, Maastrichtian of Iran; B) Loftusia morgani Douville, Maastrichtian of Turkey. Scale bar = 0.5mm. showing Turborotalia sp. trapped in a chamber, hence the suggested?Eocene age, see BouDagher- Fadel and Price, 2009). Superfamily ATAXOPHRAGMIOIDEA Schwager, 1877 Members of this family have multilocular, high trochospiral tests that become biserial or uniserial in later stages. Middle Triassic to Holocene. Family Ataxophragmiidae Schwager, 1877 Tests are initially trochospirally enrolled, and often triserial or biserial with no contin- uous composite central columella. Apertures are high and terminal (Figs 5.7 and 5.8). Late Triassic to Palaeocene. • Ataxophragmium Reuss, 1860 (Type species:  Bulimina variabilis d’Orbigny). The chambers are formed in a low trochospiral with a simple undivided interior. Late Cretaceous (Cenomanian) to Palaeocene. • Opertum Voloshina, 1972 (Type species:  Ataxophragmium (Opertum) incognitum Voloshina, 1972). The shape of the test is similar to Ataxophragmium but the i nterior of the chambers are subdivided by numerous buttresses of internal partial partitions, that form distinct chamberlets. Cretaceous (Aptian to Maastrichtian) (Plate 5.10, figs 1-2 ). • Voloshinoides Barnard and Banner, 1980 (Type species: Arenobulimina labirynthica Voloshina, 1961). The test is trochospirally enrolled with whorls enlarging rapidly. It has thick walls with the interior of the chambers being subdivided by numer- ous arched peripheral partitions. Often, they have intercalary buttresses attached to the outer wall that may meet to form irregular chamberlets. Cretaceous (Albian to Maastrichtian) (Plate 5.10, figs 3- 4). newgenrtpdf Figure 5.7. Key characteristics of Mesozoic genera with high comes. Trochospiral, often triserial to biserial with no continuous composite col- umella (the Ataxophragmiidae); with continuous composite columella (the Pfenderenidae); high trochospiral, with quinqueserial or quadriserial or triserial or biserial coiling modes, or with certain consecutive pairs of these (the Chrysalidinidae). Spiral suture and septa Initially 3 Tightly coiled, No peripheral partitions Ataxophragmium strongly oblique to the long chambers/ whorl low spire with peripheral Opertum axis then 4 or more partitions No peripheral partitions Arenobulimina with peripheral Voloshinoides partitions Spiral suture almost at right 3 chambers/w horl Low chambers, bo Verneuilinoides angles to long axis (septa from throughout umbilical tube oblique to parallel to long High chambers, with Siphovalvulina axis) umbilical tube 4 ch/ wh in adult 6/ 7 ch/ wh to 4 ch/w h Orientalia 4 ch/ wh throughout Verneuilinella 4 to 3 ch/ wh 4 ch/ wh to 3 ch/ wh Riyadhella 4 ch/ wh reducing Septa convex No peripheral partitions Dorothia to biserial With peripheral Cuneolina partitions Septa flat or No peripheral partitions Marssonella concave With peripheral Pseudotextulariella partitions adult with 3 to 4 Imperforate No pillars Pseudomarssonella ch/ wh apertural plate adult with 3 ch/ Praechrysalidina wh Scattered pillars on Dukhania apertural plate Abundant pillars Accordiella crowded centrally on broad plate I n t e r i o m a r g i n a l a p e r t u r a l l i p I n t e r i o m a r g i n a l a p e r t u r e s h a v e l i p   - o n l y a t h i n r i m e x t e n d e d a c r o s s u m b i l i c u s a s a b r o a d p l a t e I n i t i a l l y 4 o r m o r e c h a m b e r s p e r w h o r l , o f t e n r e d u c i n g l o o s e l y c o i l e d , h i g h s p i r e P e r f o r a t e a p e r t u r a l p l a t e ( c r i b r a t e a p e r t u r e s ) 298 Evolution and Geological Significance of Larger Benthic Foraminifera Ataxophragmium (Cen. - Pal.) Orbignyna (Camp. - Maast.) Opertum (Camp.-Maast.) Voloshinovella (Camp.) Arenobulimina Voloshinoides (Albian - Maastrichtian) Figure 5.8. The evolution of the internal structure of the simple Cretaceous ataxophragmids. Family Cuneolinidae Saidova, 1981 Tests are conical to sub- flabelliform, and trochospiral in the early stage with as many as five chambers per whorl, rapidly reduced to biserial. Chambers are subdivided by radial partitions and may have horizontal partitions (see Fig. 5.9). Apertures are sim- ple slits or series of pores along the base of the final chamber face. Middle Triassic (Anisian) to Late Cretaceous (Coniacian). • Cuneolina d’Orbigny, 1839 (Type species: Cuneolina pavonia d’Orbigny, 1846). The test is compressed. The interiors of the chambers are divided into nearly rectangular chamberlets by both vertical and horizontal partitions perpendicular to the outer wall. Apertures form an aligned series of pores, along the base of the final chambers. Cretaceous (Valanginian to Maastrichtian) (Plate 5.10, figs 5-1 3, 15). • Pseudotextulariella Barnard, 1953 (Type species:  Textulariella cretosa Cushman, 1932). The initial stage is trochospiral, becoming triserial, then biserial. The cham- bers are subdivided by vertical and horizontal partitions, resulting in up to six tiers of chamberlets per chamber. The aperture is a single low slit at the base of the flat- tened apertural face. Cretaceous (Berriasian to Cenomanian) (Plate 5.10, figs 16, 17). • Sabaudia Charollais and Brönnimann, 1965 (Type species:  Textulariella minuta Hofker Jr., 1965). The test has a flattened base, a small initial trochospiral of three to four subglobular chambers, followed by later triserial and biserial final stages. Biserial chambers are subdivided by vertical and horizontal partitions resulting in a subepidermal network of chamberlets. The aperture is a single slit. Sabaudia differs from Pseudotextulariella in having a thick, hyaline calcareous layer enveloping the embryonts, which is often found isolated, and by the lack of foreign, agglutinated grains in the wall of species with tiered chamberlets. Cretaceous (Valanginian to early Cenomanian) (Plate 5.10, figs 18- 19). • Vercorsella Arnaud- Vanneau, 1980 (Type species:  Vercorsella arenata Arnaud- Vanneau, 1980). The test is flaring, with an early trochospiral stage followed by a biserial stage. Chambers are subdivided by radial partitions that increase in number in successive chambers. Less well developed horizontal partitions are present in later chambers. The aperture is a single, simple slit in the basal groove. Early Cretaceous (Valanginian to Albian) (Plate 5.10, fig. 14). The Mesozoic Larger Benthic Foraminifera: The Cretaceous 299 Rectangular chamberlets Protoconch A B Beams Septa Protoconch C D Figure 5.9. A- B) Cuneolina parva Henson, Tibet, A) Vertical section, showing the interior of the chambers which is divided into nearly rectangular chamberlets by both vertical and horizontal partitions perpendicu- lar to the outer wall, B) horizontal section; C- D) Vercorsella arenata Arnaud- Vanneau, C) Vertical section, D) horizontal section. Family Dicyclinidae Loeblich and Tappan, 1964 The test is discoidal with cyclical chambers subdivided by transverse and radial parti- tions that form numerous small chamberlets. Cretaceous (Albian to Santonian). • Dicyclina Munier-C halmas, 1887 (Type species:  Dicyclina schlumbergeri Munier- Chalmas, 1887). Forms have a flattened initial stage, that is often formed with one proloculus, with later chambers becoming annular, and alternately added on the two sides of the test. The interiors of the chambers are divided by numerous thin radial partitions. The aperture is made of multiple pores at the periphery. Cretaceous (Albian to Santonian) (Plate 5.11, figs 1-3 ). Superfamily ORBITOLINOIDEA Martin, 1890 Tests are conical with numerous chambers, partially subdivided by radial or transverse partitions or with pillars. Early Cretaceous to Oligocene. Family Orbitolinidae Martin, 1890 Forms have an initial low trochospire, that is usually very much reduced and later is rectilinear, broad and conical (Fig. 5.10). They have low uniserial chambers, subdivided 300 Evolution and Geological Significance of Larger Benthic Foraminifera Valdanchella Valanginian Coskinolinoides Urgonina Falsurgonina (Aptian - Albian) Barremian (l. Barremian - e. Aptian) Orbitolina (upper Albian to upper Cenomanian) Dictyoconus (Aptian - Oligocene) Periembryonic chambers Proloculus Deuteroconch Figure 5.10. The gradual evolution from Coskinolinoides to Orbitolina. The Mesozoic Larger Benthic Foraminifera: The Cretaceous 301 with pillars or peripheral vertical partitions. The aperture is cribrate. Early Cretaceous to Oligocene. (i) No complex central zone: • Abrardia Neumann and Damotte, 1960 (Type species:  Dictyoconus mosae Hofker, 1955). Occasional radial partitions extending across the central region of the test. Late Cretaceous (Campanian to Maastrichtian) (Plate 5.7, fig. 9). • Campanellula De Castro, 1964 (Type species:  Campanellula capuensis De Castro, 1964). The test is narrow, lacking thick radial partitions but having chamberlets in the uniserial part, separated by strongly undulating septa. Early Cretaceous (Valanginian to Barremian) (Plate 5.7, fig. 10). • Coskinolinoides Keijzer, 1942 (Coskinolinoides texanus Keijzer, 1942). Radial partitions alternate in successive chambers and extend from the periphery to the centre, with an areal aperture. Early Cretaceous (Aptian to Albian) (Plate 5.12, fig.10). • Cribellopsis Arnaud- Vanneau, 1980 (Orbitolinopsis? neoelongata Cherchi and Schroeder, 1978). With peripheral rectangular chamberlets made by alternat- ing longer and shorter radial partitions. Early Cretaceous (late Hauterivian to early Aptian) (Plate 5.7, fig. 7). • Falsurgonina Arnaud- Vanneau and Argot, 1973 (Type species:  Falsurgonina pileola Arnaud-V anneau and Argot, 1973). Occurs without tiers of periph- eral chamberlets and no true pillars in the central zone. Early Cretaceous (late Barremian to early Aptian) (Fig. 5.10). • Pseudorbitolina Douvillé, 1910 (Type species:  Pseudorbitolina marthae H. Douvillé, 1910). The cyclic chambers in the median zone are subdivided into chamberlets by long radial partitions, thickening towards the lower side of the test, but leaving a large undivided zone. Late Cretaceous (Campanian to Maastrichtian) (Plate 5.7, figs 8, 14). (ii) A complex central zone and radial partitions thickening away from the periph- ery, that become broken up into pillars in the central zone: • Dictyoconus Blanckenhorn, 1900 (Type species: Dictyoconus egyptiensis (Chapman) = Patellina egyptiensis (Chapman). Aptian to Oligocene (Plate 5.13, figs 11-12). • Montseciella Cherchi and Schroeder, 1999 (Type species: Paleodictyoconus glanensis Foury, 1968). Cretaceous (late Hauterivian-early Aptian) (Plate 5.9, figs 2-4; Plate 5.13, figs 7-10). • Paracoskinolina Moullade, 1965 (Type species:  Coskinolina sunnilandensis Mayne, 1955). Without tiers of peripheral rectangular chamberlets. Early Cretaceous (Barremian to Albian) (Plate 5.2, fig. 3; Plate 5.12, fig. 11). • Paleodictyoconus Moullade, 1965 (Type species: Dictyoconus cuvillieri Foury, 1963). With peripheral tiered rectangular chamberlets in a narrow marginal zone subdivided by many radial beams. Main partitions are numerous and concentrated more towards the base of the test, those of adjacent rows alter- nating in position. Early Cretaceous (Valanginian to early Aptian) (Fig. 5.11; Plate 5.13, fig. 13). 302 Evolution and Geological Significance of Larger Benthic Foraminifera • Urgonina Foury and Moullade, 1966 (Type species:  Urgonina protuberans Foury and Moullade, 1966). Apex twisted and outer parts of the chambers lack partitions. Early Cretaceous (Barremian) (Fig. 5.10). (iii) Radial partitions thicken away from periphery to anastomose centrally around the aperture and form a reticulate zone in transverse section: • Calveziconus Caus and Cornella, 1982 (Type species Calveziconus lecalvezae Caus and Cornella, 1982). Thick peripheral wall with numerous long radial partitions with shorter partitions intercalated among the longer ones near the periphery. Late Cretaceous (Campanian) (Plate 5.7, fig. 11). • Iraqia Henson, 1948 (Type species: Iraqia simplex Henson, 1948). No tiers of peripheral chamberlets. With main partitions that are reticulate in the central area with transverse interconnections. Early Cretaceous (Aptian to Albian) (Plate 5.13, figs 2- 3). • Neorbitolinopsis Schroeder 1964 (Type species: Orbitolina conulus H. Douvillé, 1912). With tiers of peripheral rectangular chamberlets. Cretaceous (late Albian to early Cenomanian) (Plate 5.13, figs 5- 6). • Orbitolinella Henson, 1948 (Type species:  Orbitolinella depressa Henson, 1948). The interior of the test is subdivided by numerous radial beams, with those of successive chambers alternating in position with shorter and thin- ner secondary beams. Some beams fuse in the central zone. Cretaceous (Late Cenomanian to Turonian) (Plate 5.13, fig. 14). • Orbitolinopsis Henson, 1948 (Type species:  Orbitolinopsis kiliani Henson, 1948). Without tiers of peripheral chamberlets. Early Cretaceous (Aptian to early Albian) (Plate 5.5, fig. 4). • Valdanchella Canerot and Moullade, 1971 (Type species: Simplorbitolina (?) miliani Schroeder, 1968). The peripheral zones of chambers are subdivided into rectangular chamberlets by fine radial partitions. Long partitions alternate with smaller ones of constant thickness. Early Cretaceous (Valanginian) (Fig. 5.10). • Valserina Schroeder, 1968 (Type species:  Valserina broennimanni Schroeder and Conrad, 1968). Similar to Neorbitolinopsis except for the asymmetry of its embryon. Early Cretaceous (middle Barremian). (iv) Radial partitions become zigzag, thickened and fused centrally, giving a stellate appearance in transverse section: • Simplorbitolina Ciry and Rat, 1953 (Type species: Simplorbitolina manasi Ciry and Rat, 1953). Without tiers or peripheral chamberlets. Radial partitions are partly discontinuous and broken up into pillars. Tests are conical. Early Cretaceous (Aptian to Albian) (Plate 5.13, fig. 1). • Dictyoconella Henson, 1948 (Type species: Dictyoconella complanata Henson, 1948). The test is laterally compressed, but with a central zone having p illars and tiered peripheral chamberlets. Late Cretaceous (late Cenomanian to Maastrichtian) (Plate 5.7, figs 1- 3). (v) No pillars, but with peripheral tiered, rectangular chamberlets in two or more series. Radial partitions thicken, with triangular cross- section, away from the periphery and anastomose in the central area (Fig. 5.11). The earliest formed The Mesozoic Larger Benthic Foraminifera: The Cretaceous 303 chambers of the megalospheric generation can form a complex embryonic appa- ratus that can be divided into a protoconch, deuteroconch, a sub- embryonic zone and peri- embryonic chamberlets depending on the genera involved (see Schroeder, 1975, Simmons et al. 2000): • Conicorbitolina Schroeder, 1973 (Type species:  Orbitolites conica d’Archiac, 1837). The large proloculus is divided into a protoconch and deuteroconch with the marginal zone becoming extensively divided by vertical and horizontal par- titions. Cretaceous (late Albian to early Cenomanian) (Plate 5.12, fig. 5). Periembryonic zone exhibing irregular network of plate H I J Deuteroconch Protoconch Subembryonic zone are visible epidermis E F G Marginal zone Reculate parons Radial parons Radial chamber passage A B C D Figure 5.11. A) Montseciella arabica (Henson) Barremian from Qatar, NHM P35805, transverse section through the embryonic apparatus showing the reticulate zone; B) Palorbitolina lenticularis (Blum), sub- horizontal section, pyrite filled from the early Aptian of Venezuela, figured by Hofker (1963); C) Orbitolina sp., a random tangential section through chamberlets, radial and central zones; D, I), Mesorbitolina aperta (Erman, 1854) from the Albian of Tibet, figured by BouDagher- Fadel et al. (2017), the upper part of the deuteroconch is subdivided by several sets of alveoli of different length and breath, whereas the basal part exhibits a more or less developed irregular network of plates; E) Praeorbitolina cormyi Schroeder, from the Aptian of Tibet figured by BouDagher-F adel et al. (2017), the oldest representative of this lineage, shows a clearly developed initial spiral and a bilateral symmetrical embryonic apparatus, whose deuteroconch is still undivided; F) Palorbitolinoides hedini Cherchi and Schroeder from the Albian of Tibet, axial section through the large embryonic apparatus which is in a central position and consisting of alveolar protoconch and fused peri- embryonic chamberlets. The megalospheric embryonic apparatus in a central position G, J) Mesorbitolina texana (Roemer), from the Albian of Tibet, tangential section showing the embryonic appa- ratus with the regular partitions within the sub- embryonic zone followed by chambers triangular in shape; H) Orbitolina qatarica Henson, from early Cenomanian of Qatar, Arabia, NHM Henson coll., showing the complex embryonic apparatus with a deuteroconch about three times thicker than the sub- embryonic zone, both are subdivided by radial beams. Scale bars = 0.5mm. 304 Evolution and Geological Significance of Larger Benthic Foraminifera • Eopalorbitolina Schroeder, 1968 (Type species: Eopalorbitolina charollaisi Schroeder, 1968). The embryonic apparatus is present at one side of the apex, and consists of a globular proloculus and a lateral deuteroconch. Early Cretaceous (Barremian). • Eygalierina Foury, 1967 (Type species: Eygalierina turbinata Foury, 1968). Embryonic apparatus consisting of a flattened protoconch and a larger deu- teroconch, surrounded by a hemispherical zone of eight to ten periembryonal chambers. Early Cretaceous (Barremian). • Orbitolina d’Orbigny 1850 (Type species: Orbulites concava Lamarck, 1816). There is a large apically situated embryonic apparatus. The deuteroconch is about three times thicker than the sub- embryonic zone and both are subdi- vided by radial beams. Zigzag radial partitions occur in the central zone and there is a series of marginal and vertical plates. Cretaceous (late Albian to Cenomanian) (Figs 5.11C, 5.11H; Fig. 5.12; Plate 5.11, figs 4-6; Plate 5.12, figs 1, 4, 6; Plate 5.14, figs 5, 14- 15). • Mesorbitolina Schroeder, 1962 (Type species:  Orbitulites texanus Roemer, 1849). There is an apically situated embryonic zone in which the deuteroconch and subembryonic zone are of more or less equal thickness. Both are subdi- vided by vertical exoskeletal beams. Early Cretaceous (early Aptian to early Cenomanian) (Plate 5.12, fig.2; Plate 5.14, figs 6-12). • Naupliella Decrouez and Moullade, 1974 (Type species:  Naupliella insolita Decrouez and Moullade, 1974). There is an embryonic apparatus near the apex, consisting of an undivided protoconch and a deuteroconch divided in the upper part by vertical partitions. Early Cretaceous (late Albian) (Plate 5.16, fig. 9). • Neoiraqia Danilova, 1963 (Type species: Neoiraqia convexa Danilova, 1963). The embryonic apparatus is near the apex of a conical test. Later chambers are uniserial and discoidal with a narrow cellular subepidermal marginal zone cl ra sut sut r sl e b sf Figure  5.12. Endoskeleton in Orbitolina. The model shown was made in the years around 1955 by M. Reichel and was later published by Hottinger (2006). b: beam; cl:  chamberlet; e:  epidermis; r:  rafter; ra:  ramp; sf:  septal face; sl:  septulum; sut:  suture of the chambers. Double arrows in E and F: crosswise oblique foraminal axes. The Mesozoic Larger Benthic Foraminifera: The Cretaceous 305 formed by alternating radial beams. The central zone is wide and reticulate. Late Cretaceous (late Cenomanian) (Plate 5.2, fig. 4). • Palorbitolina Schroeder, 1963 (Type species:  Madreporites lenticularis Blumenbach, 1805). There is a large, apically sited, globular, fused protoconch and deuteroconch forming a relatively simple embryonic apparatus, which can be surrounded by a peri- embryonic ring of obliquely arranged chamberlets. Early Cretaceous (late Barremian to Aptian) (Fig. 5.28 Plate 5.14, figs 1- 4, 13, 16). • Palorbitolinoides Cherchi and Schroeder, 1980 (Type species: Palorbitolinoides hedini Cherchi and Schroeder, 1980). Similar to Palorbitolina, but the peri- embryonic ring of chambers is expanded and fused to form a sub- embryonic area with a few beams, which divide into irregular chamberlets. Early Cretaceous (late Aptian to early Albian) (Fig. 5.11 F; Fig. 5.29). • Praeorbitolina Schroeder, 1965 (Type species: Praeorbitolina cormyi Schroeder, 1965). A small embryonic apparatus is eccentrically situated and not completely surrounded by the earliest post-e mbryonic chambers. Early Cretaceous (early Aptian to late Albian) (see BouDagher-Fadel et al., 2017)) (Fig. 5.11 F; Fig. 5.29). • Rectodictyoconus Schroeder, 1964 (Type species: Rectodictyoconus giganteus Schroeder, 1964). The test is without an early coiled stage. An embryonic appa- ratus consists of a protoconch symmetrically overlying a deuteroconch, the lat- ter is surrounded by a marginal zone subdivided by vertical beams. The central zone has incomplete and anastomosing pillars. Early Cretaceous (Barremian). Superfamily COSKINOLINOIDEA Moullade, 1965 The test is conical, with an early stage trochospiral, then becoming uniserial and rectilinear. Apertures are basal, and cribrate. Late Cretaceous (Cenomanian) to Eocene (Lutetian). Family Coskinolinidae Moullade, 1965 Forms have a rectilinear part with broad, low chambers, subdivided by pillars or irreg- ular partitions. Late Cretaceous (Cenomanian) to Eocene (Lutetian). • Lituonelloides Henson, 1948 (Type species: Lituonelloides compressus Henson, 1948). The test is elongate, with an early conical stage and trochospirally enrolled cham- bers. The central regions of the chambers in the rectilinear part have irregular pillars. Late Cretaceous (Maastrichtian) (Plate 5.5, fig. 15). • Pseudolituonella Marie, 1955 (Type species: Pseudolituonella reicheli Marie, 1955). Initially there is a short, trochospiral stage, but the later stage shows broad and low uniserial chambers. The chamber interior has short tubular pillars. Late Cretaceous (Cenomanian to Campanian) and Eocene (Plate 5.6, fig. 17). Superfamily TEXTULARIOIDEA Ehrenberg, 1838 In this superfamily, the test is trochospiral, biserial or triserial in the early stages and later may be uniserial or biserial. Walls are agglutinated and canaliculated. Early Jurassic (Sinemurian) to Holocene. Family Chrysalidinidae Neagu, 1968 Tests form high trochospirals, with quinqueserial or quadriserial or triserial or biserial coiling modes, or with certain consecutive pairs of these (Figs 5.7). The aperture is central along the axis of coiling. In quadriserial or quinqueserial forms, an umbilicus is 306 Evolution and Geological Significance of Larger Benthic Foraminifera present and the aperture is covered with a broad umbilical flap, which may be penetrated by multiple accessory apertures. Internal pillars may develop between successive intra- umbilical flaps. Banner et al. (1991) divided this family into two subfamilies; the mainly Cretaceous Chrysalidininae (see below) and the mainly Jurassic Paravalvulininae (see Chapter 4). This classification is followed in this and the previous chapter. Subfamily Chrysalidininae Neagu, nom. Transl. This subfamily, as revised by Banner et al. (1991), is essentially triserial throughout its ontogeny (at least in the megalospheric generation), becoming biserial or quadrise- rial in the adult. Walls are solid but sometimes becoming canaliculate. Early Jurassic (Sinemurian) to Late Eocene. • Accordiella Farinacci, 1962 (Type species:  Accordiella conica Farinacci, 1962). Initially triserial, but later quadriserial (or multiserial) with internal pillars. Late Cretaceous (Coniacian to Santonian) (Plate 5.15, fig. 21). • Chrysalidina d’Orbigny, 1839 (Type species: Chrysalidina gradata d’Orbigny, 1839). Triserial throughout but with internal pillars. Cretaceous (Aptian to Cenomanian) (Plate 5.15, figs 6- 8, 16). • Dukhania Henson, 1948 (Type species: Dukhania conica Henson, 1948). Initially tri- serial but biserial in the adult stage, with internal pillars and convex septa. Cretaceous (Hauterivian to earliest Albian) (Plate 5.15, figs 10, 19- 20). • Praechrysalidina Luperto Sinni, 1979 (Type species: Praechrysalidina infracretacea Luperto Sinni, 1979). Triserial throughout, no internal pillars. Early Cretaceous (Hauterivian to Albian) (Plate 5.15, figs 12-1 4). Superfamily CYCLOLINOIDEA Loeblich and Tappan, 1964 Members of this superfamily have a test that is composed of cyclical chambers fil- led with numerous radial pillars. Late Jurassic (Kimmeridgian) to Late Cretaceous (Santonian). Family Cyclolinidae Loeblich and Tappan, 1964 The test shows annular or cyclical adult chambers that may be subdivided by pillars or partitions. Cretaceous (Valanginian to Campanian). Subfamily Cyclopsinellinae Loeblich and Tappan, 1984 The tests have cyclical chambers, with a median region with numerous radial pillars, but there are no subdivisions present in the external subepidermal area. A radial stolon system is present. Late Jurassic (Kimmeridgian) to Late Cretaceous (Santonian). • Cyclopsinella Galloway, 1933 (Type species: Cyclopsina steinmanni Munier-C halmas, 1887). The chamber interiors are undivided in the early stage, but later they have pil- lars which may bifurcate and fuse to form anastomosing pillars. Apertures form a double row of pores. Late Cretaceous (Cenomanian to Santonian). • Mangashtia Henson, 1948 (Type species: Mangashtia viennoti Henson, 1948). The test is compressed, discoidal with numerous cyclical or annular chambers. Apertures are multiple, aligned in one row in the middle of the apertural face. The axes of the stolons are radial. Numerous subcylindrical or beam shaped pillars that are The Mesozoic Larger Benthic Foraminifera: The Cretaceous 307 perpendicular to the septa are present in the central zone of the chambers. The mar- ginal zone of the chamber is not internally subdivided. Originally listed by Loeblich and Tappan in their “genera of uncertain status”. They noted that the genus is “unrec- ognisable” because many of the essential characters were not described by Henson. Fourcade et al. (1997) revised the genus based on the study of new topotype mate- rial, as well as specimens preserved in the Henson Collection housed in the Natural History Museum (London), and placed the genus in the subfamily Cyclopsinellinae. Their description is therefore regarded as the emendation. Mangashtia differs from Cyclopsinella in the nature of the internal structure (pillars in the form of beams) and in its apertural characteristics (a single row of pores rather than a double row). Late Jurassic (Kimmeridgian) to Late Cretaceous (Turonian) (Plate 5.1, fig. 14; Plate 5.9, fig. 5). Subfamily Ilerdorbinae Hottinger and Caus, 1982 The outer parts of the chambers are subdivided by secondary partitions. Apertures alternate in position from chamber to chamber, producing an oblique stolon system. Cretaceous (Valanginian to Campanian). • Dohaia Henson, 1948 (Type species: Dohaia planata Henson, 1948). Simple exoskel- etal partitions perpendicular to the wall subdivide the peripheral part of the cham- bers. Late Cretaceous (Cenomanian to Turonian) (Plate 5.16, figs 1- 3). • Eclusia Septfontaine, 1971 (Type species: Eclusia moutyi Septfontaine, 1971). Shows no alveolar hypodermis. Central pillars span flabelliform chambers with the parti- tions forming a series of chamberlets. Early Cretaceous (Valanginian to Barremian) (Plate 5.16, fig. 4). ORDER ROTALIIDA Delage and Hérouard, 1896 Members of this order have tests that are multilocular with a calcareous wall, made of perforate hyaline lamellar calcite. They exhibit apertures that are either simple or have an internal tooth plate. Triassic to Holocene. Superfamily PLANORBULINOIDEA Schwager, 1877 Tests are trochospiral in the early stages, but later stages may be uncoiled and rectilinear or biserial or with many chambers in the whorl. Apertures are intra- to extra- umbilical, but additional equatorial apertures may be present. Early Cretaceous (Berriasian) to Holocene. Family Cymbaloporidae Cushman, 1927 Chambers occur in a single layer. Late Cretaceous (Cenomanian) to Holocene. • Archaecyclus Silvestri, 1908 (Type species:  Archaecyclus cenomaniana (Seguenza)  =  Planorbulina? Cenomaniana Seguenza, 1882). The test has a large proloculus, followed by an enrolled stage of 5 chambers per whorl, later chambers occur in an annular series. Late Cretaceous (Cenomanian to Campanian) (Plate 5.5, figs 1- 7). 308 Evolution and Geological Significance of Larger Benthic Foraminifera Superfamily ORBITOIDOIDEA, Schwager, 1876 Tests are discoidal to lenticular with prominent dimorphism, in most orbitoidal species both megalospheric and microspheric generations are found. Microspheric specimens occur with a distinctly small protoconch (usually about 20 μm), but megalospheric forms have a distinctive embryonic stage, enclosed in a thicker wall. Equatorial and lateral chambers may be differentiated or indistinguishable (Figs 5.13-5.17). Late Cretaceous (Santonian) to Oligocene. Family Orbitoididae Schwager, 1876 Tests are large lenticular, non-c analiculate, and composed of a median layer, compris- ing the initial stage surrounded by concentrically arranged equatorial chambers and flanked by layers of lateral chamberlets. The embryonic chamber usually consists of two parts (a globular first chamber or protoconch and a reniform second chamber or deuteroconch) surrounded by a relatively thick wall that is in turn surrounded imme- diately by equatorial chambers, called peri- embryonic chambers (Fig. 5.13). The third chamber is called an auxiliary chamber, and in advanced species, there may be two auxiliary chambers. Stolons (apertures) are present and connect the equatorial cham- bers. Most genera of Cretaceous orbitoidal foraminifera are ornamented with rounded Embryo wall II adauxiliary chamber PAC PAC I Auxiliary chambers Figure 5.13. Embryonic and periembryonic chambers in Orbitoides spp. according to van Hinte’s 1966 con- cept. I: protoconch; II: deuteroconch; PAC: principal auxiliary chamber. The Mesozoic Larger Benthic Foraminifera: The Cretaceous 309 pustules. Late Cretaceous (Santonian to Maastrichtian). For a comprehensive study of this group see van Gorsel (1978). Subfamily Orbitoidinae Schwager, 1876 Lateral chambers may be present and differentiated from the equatorial layer in the post embryonal stage. Late Cretaceous (late Santonian to Maastrichtian). • Monolepidorbis Astre, 1928 (Type species:  Monolepidorbis sanctae- pelagiae Astre, 1927). Monolepidorbis differs from Orbitoides by the absence of lateral chamberlets. However, long incipient lateral chamberlets may be present. Several authors (van Gorsel, 1978; Loeblich and Tappan, 1988) have argued in favour of the inclusion of these forms in Orbitoides. Late Cretaceous (late Santonian to Maastrichtian) (Plate 5.17, figs 1- 4; Fig. 5.14). • Orbitoides d’Orbigny, 1848 (Type species: Orbitoides media (d’Archiac) = Orbitolites media d’Archiac, 1889). Lenticular to discoidal in shape. The test is made up of a single, roughly flat layer of large chambers (the median layer) with a mound of much smaller chambers on either sides (the lateral chambers). In plan view, the test is slightly asymmetric, usually with a central knob on one side and radial ridges on the test. In the central part of the median layer, there is a group of two to four chambers with thin and straight mutual walls, the whole enclosed within a thick and conspicuously porous outer wall. Epi-a uxilliary chambers (chambers originating from stolons in the embryonal wall enveloping the protoconch, deuteroconch and auxiliary chambers) are present. This is followed immediately by concentric rings of simultaneously- formed smaller chambers. Stolons are big enough to be observed in vertical sections. Late Cretaceous (late Santonian to Maastrichtian) (Plate 5.17, figs 5- 9; Plate 5.18, figs 1-6 ; Figs 5.15, 5.16). • Pseudomphalocyclus Meriç, 1980 (Type species:  Pseudomphalocyclus blumenthali Meriç, 1980). The embryo is followed by a single row of arcuate equatorial cham- bers, growing to two layers and three layers near the periphery. Three to four layers or lateral chamberlets are present at the centre of the test, decreasing to one or two layers near the periphery. Late Cretaceous (late Maastrichtian) (Fig. 5.14). • Simplorbites de Gregorio, 1882 (Type species: Simplorbites cupulimis de Gregorio, 1882 = Nummulites papyraceus Boubeé, 1832). This form shows a median layer, arched and added to by an annular series. Lateral chamberlets occur with pillars. Equatorial chamberlets are curved and surrounded on both sides by small curved thickened chamberlets. Simplorbites differs from Orbitoides in having a larger nucle- oconch, which is subdivided by numerous arcuate partitions into a multilocular form instead of a quadrilocular or a bilocular one, as in Orbitoides. Late Cretaceous (Maastrichtian) (Plate 5.19, figs 1-6 ; Fig. 5.14). Subfamily Omphalocyclinae Vaughan, 1928 The lateral chambers are not differentiated from the equatorial layer in the post embry- onal stage. Late Cretaceous (Maastrichtian). • Omphalocyclus Bronn, 1853 (Type species:  Omphalocyclus macroporus Bronn, 1853). The test is discoidal and biconcave. The central part of the test may consist of one or two layers of chambers. Additional layers of arched chambers are added in alternate stages, 310 Evolution and Geological Significance of Larger Benthic Foraminifera Orbitoididae Hel. Lepidorbitoididae Pseudorbitoididae Period, Epoch, Stage Small rotaliines Archaecyclus (Cen.-Camp.) Figure 5.14. The different lineages of the orbitoids. The shaded boxes highlight forms found only in the Caribbean province. 1 4 5 . 5 9 9 . 6 6 5 . 5 A g e ( M a ) C r e t a c e o u s L a t e E a r l y B e r . V a l . H a u t . B a r r . A p t . A l b . C e n . T u r . C o n . S a n t . C a m p . M a a s t . M o n o l e p i d o r b i s O r b i t o i d e s O m p h a l o c y c l u s T o r r e i n a S i m p l o r b i t e s P s e u d o m p h a l o c y c l u s H e l l e n o c y c l i n a S i r t i n a P s e u d o s i d e r o l i t e s H e l i c o r b i t o i d e s L e p i d o r b i t o i d e s A r n a u d i e l l a V a n d e r b e e k i a S u l c o p e r c u l i n a C l y p e o r b i s H i s t o r b i t o i d e s S u l c o r b i t o i d e s C o n o r b i t o i d e s C t e n o r b i t o i d e s V a u g h a n i n a P s e u d o r b i t o i d e s A k t i n o r b i t o i d e s O r b i t o c y c l i n a A s t e r o r b i s The Mesozoic Larger Benthic Foraminifera: The Cretaceous 311 instead of an orbitoid spiral (Fig. 5.15). The embryonic stage has two to four chambers. Equatorial chambers communicate through large marginal stolons. Late Cretaceous (Maastrichtian) (Figs 5.15A, B; Fig. 5.16C; Plate 5.20, figs 1- 5; Plate 5.21, figs 2- 4, 6- 8). • Torreina Palmer, 1934 (Type species: Torreina torrei Palmer, 1934). The test is glob- ular with a centrally located embryon. The embryonic stage has four to five cham- bers, surrounded by a thick wall. From the embryon, lateral chambers are low and arcuate, and added uniformly in all directions. This growth pattern is characteristic of Cretaceous orbitoids, but is also found in the Tertiary genus Sphaerogypsina (see Chapter 6). Late Cretaceous (Maastrichtian) (Fig. 5.14). Family Hellenocyclinidae Freudenthal, 1969 Shows an orbitoid test, but without lateral chamberlets. Late Cretaceous (Maastrichtian). • Hellenocyclina Reichel, 1949 (Type species:  Hellenocyclina beotica Reichel, 1949). The test is small and slightly conical. The embryonic stage consists of five to six chambers, cruciformly arranged around the protoconch. The embryon is surrounded by arcuate equatorial chambers, arranged in a regular concentric growth pattern. The lateral walls are traversed by fine pores. It differs from Lepidorbitoides (below) by the absence of lateral chambers, the small size of the test and initial chambers. Late Cretaceous (Maastrichtian) (Fig. 5.14). A B C D Figure 5.15. A, B) development of chamberlet cycles, enlargement of the arched chambers of Omphalocyclus sp. They are added in alternate stages, instead of an orbitoid spiral. The embryonic stage has two to four chambers, and equatorial chambers communicate through large marginal stolons; C- D) Orbitoides sp., South Africa, with concentric rings of simultaneously formed smaller chambers, and with stolons that are big enough to be observed in vertical sections. 312 Evolution and Geological Significance of Larger Benthic Foraminifera Annular chamberlets A Lateral regularly arranged Stolons chamberlets absent Long incipient lateral chamberlets B C Figure 5.16. A) Equatorial section of Monolepidorbis; B) Vertical section of Monolepidorbis; C) Vertical section of Omphalocyclus. Family Lepidorbitoididae Vaughan, 1933 The Lepidorbitoididae differ from the Orbitoididae by having a distinct structure to their embryonic apparatus, and by the form of their median chambers. An embryonic stage with two chambers is followed by hexagonal or arcuate equatorial chamberlets and by differentiated lateral chambers (Fig.  5.17). Late Cretaceous (Santonian) to Middle Eocene. • Arnaudiella Douvillé, 1907 (Type species: Siderina douvillei Abrard, 1926=Arnaudiella grossouvrei H. Douvillé. 1907). The test is large, flat, lenticular with a wall much thickened into a broad flattened flange at the periphery, pierced by radial canals that open into coarse pores at the periphery, or into the chamber lumen of the succeed- ing whorl. Numerous pustules cover the ventral area. The embryonic apparatus is composed of a spherical protoconch and a slightly larger deuteroconch, followed The Mesozoic Larger Benthic Foraminifera: The Cretaceous 313 Orbitoides Lepidorbitoides 1- Nucleoconch 2- Plan view of segment of median layer 3- section across median layer Figure 5.17. The main important features which differentiate Orbitoides from Lepidorbitoides. by two to five whorls of spirally arranged primary chambers that increase rapidly from about eight chambers in the earliest whorl to up to forty five in the latest whorl. No secondary equatorial chambers are found. At the peripheral side of the whorls of the primary chambers, radial canals and rods are formed. On the lateral sides, orbitoidal chamberlets are added between the planispiral- involute spiral whorls. Late Cretaceous (late Campanian) (Fig. 5.14). • Clypeorbis Douvillé, 1915 (Type species:  Orbitoides mammillatus Schlumberger, 1902). An asymmetrical test, with a thin equatorial layer of curved chambers, with numerous hexagonal lateral chamberlets on both sides. Late Cretaceous (late Maastrichtian) (Fig. 5.14). 314 Evolution and Geological Significance of Larger Benthic Foraminifera • Helicorbitoides Macgillavry, 1963 (Type species: Pseudorbitoides longispiralis Papp and Küpper, 1953). Species of this genus have lateral chambers and a long primary spiral, between the whorls of which secondary equatorial chambers are formed. Pseudosiderolites- like radial canals and plates are conspicuous in the primitive forms. Late Cretaceous (late Campanian to Maastrichtian) (Fig. 5.14). • Lepidorbitoides Silvestri, 1907 (Type species:  Orbitoides socialls Leymerie, 1851). Forms are flat, large (up to 10 mm in diameter, and rarely up to 25 mm), lenticu- lar in shape, and made up of a single roughly flat layer of moderately large cham- bers (median layer), with a mound of much smaller chambers on either side (lateral chambers), ornamented with rounded pustules. In plan view the central part of the median layer usually consists of two moderately large chambers, the protoconch is the smaller of the two. It is partially embraced by the second chamber, the reni- form deuteroconch. Together they make up the nucleoconch. They are followed by concentric rows of closely-s paced, simultaneously formed small spatulate to arcu- ate equatorial chambers arranged in alternating positions around the nucleoconch. Nepionic arrangements vary from reduced uniserial to quadriserial, with up to 15 adauxilliary chambers (small chambers originating from solons in the deuterocon- chal wall). The equatorial chambers communicate with each other via diagonal and median stolons, while the lateral chambers do so via pores. In European material it is quite easy to distinguish Lepidorbitoides, with its regular, porous and relatively thin lateral walls, from Orbitoides with it more solid walls that have a distinct black inner lining. Late Cretaceous (late Campanian to Maastrichtian) (Fig. 5.17; Plate 5.2, figs 12- 14). • Pseudosiderolites Smout, 1955 (Type species:  Siderolites vidali Douvillé, 1907). The test is large and bilaterally symmetric. Walls are thick and lamellar. The umbilical region has spiral canals between the pillars. Apertures are single in the protoconch and deuteroconch, but multiple in later chambers with a single row of openings at the base of the apertural face. Late Cretaceous (Campanian) (Fig. 5.14). • Sirtina Brönnimann and Wirz, 1962 (Type species: Sirtina orbitoidiformis Brönnimann and Wirz, 1962). Tests are lenticular with a two- chambered embryo of approximately equal size protoconch and deuteroconch, followed by a trochospiral early stage, but later they become nearly planispiral and involute. Thin vertical canals form between well developed umbilical pillars. Several layers of orbitoidal lateral chambers are present on the dorsal side in addition to the pillars. Late Cretaceous (Santonian to early Maastrichtian) (Fig. 5.14). • Sulcoperculina Thalmann, 1939 (Type species: Camerina? dickersoni Palmer, 1914). A  rotaliine without lateral chambers but with a peripheral sulcus in which small radial plates or rods can be observed. Late Cretaceous (Campanian to Maastrichtian) (Figs 5.14; 5.18). • Vanderbeekia Brönnimann and Wirz, 1962 (Type species:  Vanderbeekia trochoi- dea Brönnimann and Wirz, 1962). This form differs from Sirtina by having a thin layer of equatorial chambers between the lateral chambers of the dorsal side and by exhibiting ventral spiral chambers. Late Cretaceous (early Maastrichtian) (Fig. 5.14). The Mesozoic Larger Benthic Foraminifera: The Cretaceous 315 radial plates Vaughanina Orbitocyclina Pseudorbitoides Campanian - Campanian - l. Campanian - Maastrichtian Maastrichtian Maastrichtian Ctenorbitoides l. Campanian - e. Maastrichtian Conorbitoides alternating "radial plates" = marginal canal system with l. Campanian -e. Maastrichtian radial elements Later chambers Sulcorbitoides Campanian laminae Sulcoperculina (L. Cretaceous) Figure 5.18. The three lineages that evolved from Sulcoperculina in the Late Cretaceous. The “radial plates” of Brönnimann (1956) and other US workers are in fact a marginal canal system with radial elements that is, however, not a marginal cord, thus unrelated to the Tertiary nummulitids. 316 Evolution and Geological Significance of Larger Benthic Foraminifera Superfamily ROTALIOIDEA Ehrenberg, 1839 Tests are involute to evolute, initially trochospiral or planispiral, commonly with many chambers in numerous whorls. As new chambers are added, septal flaps attach to previous apertural faces and enclose radial canals, fissures, umbilical cavities, and intraseptal and subsutural canals. The wall is perforate, hyaline calcite, and generally optically radial in structure. Primary apertures are single or multiple. Small opening into the canal system may occur along the sutures. Late Cretaceous (Coniacian) to Holocene. Family Pseudorbitoididae Rutten, 1935 The test is lenticular, canaliculated with a two- chambered embryon followed by a spire of nepionic chambers. This group is characterized by the presence of a structure of radially arranged calcareous elements in the equatorial layer (Figs 5.14, 5.18). Late Cretaceous (Campanian to Maastrichtian). Subfamily Pseudorbitoidinae Rutten, 1935 The equatorial (median) layer is subdivided vertically by variously arranged “radial plates or rods”. The “radial plates” of Brönnimann (1956) and other US workers are in fact a marginal canal system with radial elements that are, however, not marginal cords, thus are unrelated to the Tertiary nummulitids. Late Cretaceous (Campanian to Maastrichtian). • Conorbitoides Brönnimann 1958 (Type species:  Conorbitoides cristalensis Brönnimann 1958). Characterised by a conical test with a pointed apex and a neanic stage similar to Sulcoperculina. Late Cretaceous (late Campanian to early Maastrichtian) (Fig. 5.18). • Historbitoides Brönnimann 1956 (Type species: Conorbitoides kozaryi Brönnimann 1958). Similar to Pseudorbitoides, but the single set of radial plates are irregularly interconnected laterally, and with incipient radii and interradii. Late Cretaceous (late Campanian to early Maastrichtian). • Pseudorbitoides Douvillé, 1922 (Type species:  Pseudorbitoides trechmanni Douvillé, 1922). An orbitoidal foraminifera with a circular outline and median and lateral layers, but with the median layer being double or triple towards the periphery as in Omphalocyclus. The radial (equatorial) layer lacks annular walls. The lateral chambers are irregular in form and relatively thin walled, and connected by numerous pores in the chamber walls. The neanic stage has a single set of radial plates, that are not inter- connected radially. Late Cretaceous (late Campanian to Maastrichtian) (Fig. 5.18). • Sulcorbitoides Brönnimann, 1955 (Type species: Sulcorbitoides pardoi Brönnimann, 1955). The equatorial chambers are trochospirally arranged in three whorls. It dif- fers from Pseudorbitoides in having two sets of alternating systems of vertical radial plates, which Brönnimann referred to as “radial rods”, separated by a median gap and originating from the peripheral sulcus. This form differs from Sulcoperculina (see above) in having lateral chambers and lengthened radial plates or rods. Late Cretaceous (Campanian) (Fig. 5.18). The Mesozoic Larger Benthic Foraminifera: The Cretaceous 317 Subfamily Vaughaninae MacGillavry, 1963. Members of this family have a juvenile stage similar to Sulcoperculina developing lat- eral chambers. The equatorial layers are similar to those of the pseudorbitoids. Late Cretaceous (Campanian to Maastrichtian). • Aktinorbitoides Brönnimann, 1958 (Type species:  Aktinorbitoides browni Brönnimann, 1958). The test is lenticular, with a stellate outline. The juvenarium consists of two to three whorls of trochospirally arranged chambers, followed an equatorial layer comprising seven to ten rays (radii), consisting of annular chambers traversed by two alternating systems of radial plates, separated by a medium gap as in Vaughanina. Late Cretaceous (Campanian to Maastrichtian) (Fig. 5.14). • Ctenorbitoides Brönnimann, 1958 (Type species:  Ctenorbitoides cardwelli Brönnimann, 1958). The test is conical with a comb-l ike apex and a neanic stage similar to Vaughanina. Late Cretaceous (late Campanian to early Maastrichtian) (Fig. 5.14). • Vaughanina Palmer, 1934 (Type species:  Vaughanina cubensis Palmer, 1934). The test is lenticular and circular with a conspicuous flange on which protrud- ing radial plates are usually visible. Vaughanina differs from Pseudorbitoides in having its annular equatorial chambers crossed by two systems of alternat- ing vertical radial plates, extending from the roof and floor into the equatorial layer. Late Cretaceous (late Campanian to Maastrichtian) (Plate 5.20, fig.  7; Fig. 5.18). Subfamily Pseudorbitellinae Hanzawa, 1962 The subfamily has a pseudorbitoidal structure, but lacks plates. Late Cretaceous (Campanian to Maastrichtian). • Asterorbis Vaughan and Cole, 1932 (Type species: Asterorbis rooki Vaughan and Cole, 1932). The test is stellate with four to eight rays. The surface has large pustules on the umbo, the bilocular embryo which is surrounded by a thick wall and two large auxiliary chambers. Equatorial chambers are diamond shaped to ogival (see Fig.  5.19) in plan, and increasing in height from the centre to the test periphery. The layers decrease in number from the umbo towards the periphery, where the lateral layer is left exposed without lateral chambers, but with well developed pillars. Late Cretaceous (late Campanian to Maastrichtian) (Fig. 5.14). • Orbitocyclina Vaughan, 1929. The test is circular with a bilocular embryon, enclosed by a thick wall and followed by spiral chambers. The equatorial layer is composed of arcuate to diamond-s haped chambers, interconnected by stolons and surrounded by lateral layers of chambers on both sides. Late Cretaceous (Campanian to Maastrichtian) (Figs 5.18). 318 Evolution and Geological Significance of Larger Benthic Foraminifera Primitive "four stolon" system Arcuate chambers with basal stolons Advanced "six stolon system" with diagonal and annular stolons Ogival chambers Spatulate chambers Figure 5.19. Schematic drawing showing the shape and connections of the equatorial chambers in orbitoi- dal foraminifera, modified after van Gorsel (1975, 1978). Family Rotaliidae Ehrenberg, 1839 Tests are trochospiral, with umbilical plugs and radial canals throughout, or with fis- sures and intraseptal and subsutural canals. Apertures are umbilical, basal, single to multiple. Late Cretaceous (Maastrichtian) to Miocene. • Kathina Smout, 1954 (Type species:  Kathina delseata Smout, 1954). The test is lenticular, with a single spire of small simple chambers, without supplementary chamberlets or umbilical extensions. The umbilicus has a central plug with numer- ous strong vertical canals. This genus is distinguished from later Paleogene forms Dictyoconoides and Dictyokathina by having a simple spire, and from Lockhartia and Sakesaria by the lack of umbilical cavities (see Chapter 6). Late Cretaceous to Paleocene. • Pararotalia Le Calvez, 1949 (Type species: Rotalina inermis Terquem, 1882). Biconvex, low trochospiral, smooth with a simple plug that fills the umbilicus. Loeblich and Tappan (1988), in their new description of the genus Pararotalia, reported the pres- ence of a septal flap partially doubling each of the septa. They describe the aperture as interiomarginal, extending obliquely into the apertural face, and the intercameral foramen as areal, due to the attachment of an imperforate tooth plate that extends to the distal margin of the aperture. Hottinger et al. (1991) emended the descrip- tion of the Pararotaliinae to include the canal system. It is also distinguished from The Mesozoic Larger Benthic Foraminifera: The Cretaceous 319 Neorotalia (see Chapter 6) by the lack of the closed interlocular spaces (“envelop- ing canal system”, Hottinger et al., 1991). Late Cretaceous (Coniacian) to middle Eocene. • Fissoelphidium Smout, 1955 (Type species:  Fissoelphidium operculiferum Smout. 1955). The test is bilaterally symmetric, planispiral with numerous chambers. Deeply fissured dendritic patterns occur along the sutures and finely divided umbilical boss. Apertures are basal and multiple. Late Cretaceous (Maastrichtian) (Plate 5.2, figs 10-1 1). • Laffitteina Marie, 1946 (Type species: Laffitteina bibensis Marie, 1946). The test is lenticular, asymmetric with bifurcate interseptal canals that open as two alternat- ing rows of openings along the septal sutures on the dorsal side. Late Cretaceous (Maastrichtian) to Paleocene (Danian) (Plate 5.20, fig. 6). • Orbitokathina Hottinger, 1966 (Type species:  Orbitokathina vonderschmitti Hottinger, 1966). The test is conical, but only the microspheric generation of this genus shows orbitoidal character. Pillars are only found on the concave ventral side. The spiral stage is followed by an orbitoidal stage consisting of arcuate chambers, arranged in irregular concentric rows. Lateral chambers are absent. Late Cretaceous (Coniacian). • Rotalia Lamarck, 1804 (Type species: Rotalites trochidiformis Lamarck, 1804). The test is trochospiral and biconvex. Secondary deposits build thick, short lamellar pil- lars with marked fissures that fill the umbilical area. Apertures are umbilical, basal with umbilical extensions to the umbilical canal. Late Cretaceous (Coniacian) to Eocene (Plate 5.1, fig. 13). Family Calcarinidae Schwager, 1876 Tests are enrolled, with large inflated spines. Late Cretaceous (Maastrichtian) to Holocene. • Siderolites Lamarck, 1801 (Type species:  Siderolites calcitrapoïdes Lamarck, 1801). The test is large with a globular proloculus followed by an involute coil of about four whorls. In the final whorl two to seven large coarse spines are present. Spiral canals occur in the umbilical region on both sides of the test. Multiple foramina are present. Pillars appear in the umbilical region as solid pustules. Late Cretaceous (Maastrichtian) (Fig.  5.20; Plate 5.20, figs 8-1 1; Plate 5.21, figs 1, 6). ORDER MILIOLIDA Delage and Hérouard, 1896 The miliolides have tests that are porcelaneous, imperforate and made of high Mg- calcite with fine randomly oriented crystals. They range from the Carboniferous to the Holocene. Superfamily ALVEOLINOIDEA Ehrenberg, 1839 Tests are enrolled along an elongate axis, being initially planispiral or streptospiral, or miliolide with chambers added in varying planes. Cretaceous to Holocene. 320 Evolution and Geological Significance of Larger Benthic Foraminifera globular proloculus large coarse spines septal flap Chambers canaliculate spines A C Mul ple foramina B are present Solid pustules in the Spiral canals occur in umbilical region the umbilical region on both sides of the test F D E Figure  5.20. Siderolites showing canaliferous enveloping, marginal and pseudospinose structures:  A- B) Horizontal section; C) Oblique horizontal thin sections; D- E) Sketch of vertical and a stereodiagram drawn by Hofker (1927); F) Axial section with broken spines. Family Fabulariidae Ehrenberg, Munier- Chalmas, 1882, emend. Hottinger et al., 1989 The test is large, dimorphic, multi- chambered with a miliolide coiling, tending to become reduced in subsequent growth stages, either to bilocular or to monolocular chamber cycles with a trematophore as aperture. According to Hottinger (2006) “A trematophore, or a sieve constituting the face of many porcelaneous larger foramin- ifera, is in miliolides produced by the coalescence of teeth, covering a large pre- septal space. May be supported by residual pillars”. Chambers have a thickened basal layer, subdivided by pillars or secondary partitions. Foraminifera in the adult growth stages have fixed apertural axis. For the terminology and the orientation of sections needed for detailed structural analysis see Drobne (1974; 1988) and Hottinger et al. (1989). Late Cretaceous to Early Oligocene. • Adrahentina Bilotte, 1978 (Type species: Adrahentina iberica Bilotte, 1978). The test is large and subspherical, with a megalospheric stage with a large proloculus fol- lowed by enveloping chambers. Radial pillars arise from the chamber floor but never reach the chamber roof. Pillars are alternating on adjacent ribs. Late Cretaceous (Maastrichtian). • Lacazina Munier- Chammas, 1882 (Type species:  Alveolina compressa d’Orbigny, 1850). The test is large, discoid to ovoid. The early miliolide coiling is followed by chambers that completely overlap the early ones. Chambers are subdivided by longitudinal partitions. Apertures are cribrate at one extremity of the test. Late The Mesozoic Larger Benthic Foraminifera: The Cretaceous 321 Cretaceous (Coniacian to Maastrichtian) of Europe, but in the Early Oligocene of Indonesia (Plate 5.22, fig. 6; Fig. 5.21). • Pseudochubbina De Castro, 1990 (Type species:  Pseudedomia globularis Smout, 1963). The test is subspherical to flaring with parallel anastomising passages in the basal layer, and an outer layer with regular chamberlets. Late Cretaceous (Campanian). • Pseudolacazina Caus, 1979 (Type species: Pseudolacazina hottingeri Caus, 1979). The test is globular to ovate with a quinqueloculine early stage. Microspheric chambers have pluriloculine chamber cycles to monoloculine cycles in the adult. The megalospheric stage is biloculine, with completely embracing chambers. Chambers are subdivided by longitudinal partitions or by pillars supporting the chamber roof. Late Cretaceous (Santonian to Maastrichtian) and Middle to Late Eocene. Family Rhapydioninidae Keijzer, 1945 Tests are planispiral to streptospiral in the early stage, but become uncoiled com- pressed, peneropliform- flaring or cylindrical in later stages. The embryonal apparatus is simple or miliolide. The central zone shows canaliculated thickening except for a pre- septal space with buttresses (residual pillars). Apertures are multiple on the final chamber face. Late Cretaceous (Cenomanian to Maastrichtian). • Chubbina Robinson, 1968 (Type species:  Chubbina jamaicensis Robinson, 1968). Tests are large globular and streptospiral, later flaring planispiral and penero- pline. Interiors of chambers are subdivided by numerous septula. Late Cretaceous (Campanian to Maastrichtian) (Plate 5.3, fig. 12; Plate 5.6, figs 10- 11; Plate 5.9, figs 8- 10: Fig. 5.21). • Murciella Fourcade, 1966 (Type species: Murciella cuvillieri Fourcade, 1966). Tests are planispiral and involute, later uncoiling and becoming rectilinear. Horizontal and vertical septa subdivide the chambers into secondary chamberlets. Late Cretaceous (Campanian) (Fig. 5.21; Plate 5.4, fig. 11). • Pseudedomia Henson, 1948 (Type species: Pseudedomia multistriata Henson, 1948). Tests are globular to lenticular with the last chambers becoming progressively longer and strongly overlapping the preceding chambers, until they become cyclical with compressed arched chambers in its final whorls. Late Cretaceous (Cenomanian to Maastrichtian) (Plate 5.1, figs 1- 6; Plate 5.22, fig. 12; Fig. 5.21). • Raadshoovenia van den Bold, 1946 (Type species: Raadshoovenia guatemalensis van den Bold, 1946). The early stage is streptospiral, then planispiral and becoming rec- tilinear. The centre of the test is pillared with pillars fusing laterally to form cham- berlets opening into the preseptal space. Late Cretaceous (Campanian). • Rhapydionina Stache, 1913 (Type species:  Peneroplis liburnica Stache, 1889). The test is involute, lenticular, with pronounced dimorphism. Cylindrical megalospheric forms have an enrolled flexostyle followed by three to four chambers in a single plani- spiral coil, with a rectilinear adult stage. Planispiral to uncoiled microspheric forms have a small proloculus followed by uniserial, flabelliform and flattened adult stage with low arched chambers. The peripheral region is subdivided by vertical septula. Late Cretaceous (late Cenomanian to Maastrichtian) (Fig. 5.21). 322 Evolution and Geological Significance of Larger Benthic Foraminifera Period, Epoch, Stage Quinqueloculina gp. Jurassic Figure 5.21. The evolution of the Alveolinoidea in the Cretaceous. 1 4 5 . 0 1 0 0 . 5 6 6 A g e ( M a ) C r e t a c e o u s L a t e E a r l y T i t h . B e r . V a l . H a u t . B a r r . A p t . A l b . C e n . T u r . C o n . S a n t . C a m p . M a a s t . A r c h a e a l v e o l i n a M u r c i e l l a C h u b b i n a R h a p y d i o n i n a P s e u d e d o m i a O v a l v e o l i n a P r a e a l v e o l i n a C i s a l v e o l i n a M u l t i s p i r i n a S i m p l a l v e o l i n a S e n a l v e o l i n a L a c a z i n a The Mesozoic Larger Benthic Foraminifera: The Cretaceous 323 Family Alveolinidae Ehrenberg, 1839 Tests are free, large, planispiral to fusiform, cylindrical or globular, and coiled about elongate axis. Chambers are divided into chamberlets by secondary septa (septula) per- pendicular to the main septa, and connected along the chamber by passages. Apertures are numerous, circular, and aligned in horizontal rows. Early Cretaceous (Aptian) to Holocene. • Archaealveolina Fourcade, 1980 (Type species: Ovalveolina reicheli De Castro, 1966). The test is subglobular, small, and streptospiral in early coiling, followed by a pla- nispiral stage. Septa are oblique, with pre- septal passages present. Apertures form a single row of round openings at the base of the apertural face. Early Cretaceous (Aptian) (Plate 5.22, Fig. 2; Fig. 5.21). • Cisalveolina Reichel, 1941 (Type species: Cisalveolina fallar Reichel, 1941). The test is globular and streptospiral in the early stage, but later planispiral, The megalo- spheric generation is planispiral throughout ontogeny. Later stages have alternat- ing chamberlets. There are wide post- septal passages and a long, low single, slit- like aperture. Late Cretaceous (Cenomanian) (Fig. 5.21; Plate 5.22, figs 3- 4; 13- 14). • Helenalveolina Hottinger, Drobne and Caus, 1989 (Type species:  Helenalveolina tappanae Hottinger, et al.,1989). The test is spherical, with streptospiral coiling that may have a planispiral- involute adult stage. Early chambers are undivided in both generations, later chambers are subdivided by longitudinal partitions (septula) into alternating chamberlets as in Cisalveolina. The pre-s eptal passage is wide, with a ribbed floor, and a short post-s eptal passage. The distal margin of the apertural slit is notched. Helenalveolina is distinguished from the Fabulariidae by the absence of pillars subdividing the chambers, by its streptospiral coiling, and by the lack of a trematophore in its apertural face. Late Cretaceous (late Coniacian to early Santonian). • Multispirina Reichel, 1947 (Type species: Multispirina iranensis Reichel, 1947). The test is spherical, large with numerous chambers divided into chamberlets by the sep- tula, which is continuous from chamber to chamber. There are intercalated whorls and sutural apertures. Pre-s eptal passages are large. Apertures form a row of pores in the apertural face. Late Cretaceous (Cenomanian) (Fig. 5.21; Plate 5.1, fig. 9). • Ovalveolina Reichel, 1936 (Type species: Alveolina ovum d’Orbigny, 1850). The test is globular or ovoid with planispiral coiling. Chambers are short with widely spaced septula. Pre-s eptal passages are large. There is one row of apertures in the apertural face. Cretaceous (Albian to Cenomanian) (Fig. 5.21; Plate 5.22, figs 5, 9A, 13). • Praealveolina Reichel, 1933 (Type species: Praealveolina tenuis Reichel, 1933). Test ranges in shape from subglobular to fusiform. They consist of a number of cham- bers, planispirally coiled around an elongate axis, each chamber being broken into tubular chamberlets by secondary septa (septula) that run in the direction of coiling. It differs from Alveolina (see Chapter  6) in having its septula and tubular cham- berlets in continuous alignment from chamber to chamber. Differences in shape will ordinarily distinguish Praealveolina from Ovalveolina, but where Praealveolina is represented by subglobular species, the most practical means of differentiation is by comparison of chamberlet proportions: Ovalveolina has chamberlets that are much higher and broader in proportion to their length than those of Praealveolina. 324 Evolution and Geological Significance of Larger Benthic Foraminifera A pre- septal passage is present. Apertures form a single row of pores on the apertural face near the equator but increases pole- ward to many rows of openings. Cretaceous (late Albian to Turonian) (Plate 5.21, fig. 9, Plate 5.22, figs 9-1 1). • Senalveolina Fleury, 1984 (Type species: Senalveolina aubouini Fleury, 1984). The test is globular. Early coiling is streptospiral, but later it becomes planispiral with numerous long low chambers per whorl. Narrow pre- septal passages are present. Walls and septula are relatively thick. Apertures form a row of pores in the apertural face. Late Cretaceous (early Campanian) (Fig. 5.21). • Simplalveolina Reichel, 1964 (Type species:  Praealveolina simplex Reichel, 1936). The test is small and ovoid. Numerous septula are aligned from chamber to chamber to form a single layer of chamberlets of oval section. It lacks the secondary cham- berlets of Praealveolina. The pre- septal canal has a circular section. The aperture is a single row of pores in the apertural face. Late Cretaceous (Cenomanian) (Plate 5.21, fig. 5; Plate 5.22, fig. 8). • Streptalveolina Fourcade, Tardy and Villa, 1975 (Type species: Streptalveolina mexi- cana Fourcade, et al., 1975). The test is globular and streptospirally coiled. Septa are oblique. Chambers are divided by partial septula. A pre- septal passage is pres- ent. The aperture is a row of openings in the apertural face. Late Cretaceous (early Cenomanian). • Subalveolina Reichel, 1936 (Type species: Subalveolina dordonica Reichel, 1936). The test is spherical to fusiform. Septula have no definite arrangement. The secondary cham- berlets lie below the plane of the primary chamberlets in an irregular pattern. A large pre- septal canal is present, occupying the total height of the chamber. Above the row of primary apertures in the apertural face is a row of more numerous small apertures con- necting with alveoli. Late Cretaceous (late Santonian to Campanian) (Plate 5.22, fig. 7). Superfamily SORITOIDEA Ehrenberg, 1839 Chambers are planispiral, uncoiling, flabelliform or cyclical, and may be subdivided by partitions or pillars. Late Permian to Holocene. Family Meandropsinidae Henson, 1948 The test is planispiral to annular discoid, laterally compressed, involute with partially developed partitions. Numerous septula divide the interior of the chambers. Apertures are multiple. Late Cretaceous (Campanian to Maastrichtian). • Ayalaina Seiglie, 1961 (Type species: Meandropsina? rutteni Palmer, 1934). The test is flaring with numerous chambers increasing rapidly in height. Late Cretaceous (Campanian to Maastrichtian). • Broeckina Munier- Chalmas, 1882 (Type species:  Cyclolina dufrenoyi d’Archiac and Haime, in d’Archiac, 1854). The test is discoidal, the early stage is planispi- ral, and later chambers are annular and divided by incomplete partitions. The multiple apertures form two separate rows. Late Cretaceous (Cenomanian to Maastrichtian). The Mesozoic Larger Benthic Foraminifera: The Cretaceous 325 • Fallotia Douvillé, 1902. (Type species:  Fallotia jacquoti Douvillé, 1902). The test is involute, planispiral with chambers subdivided by numerous partitions. Late Cretaceous (Santonian) (Fig. 5.22). • Larrazetia Ciry, 1964 (Type species:  Meandropsina larrazeti Munier- Chalmas, in Schlumberger, 1898). There is an annular test, laterally flanked by a series of polar capsules, slightly biconvex with annular chambers subdivided throughout with par- tial partitions. Late Cretaceous (Campanian to Maastrichtian). • Meandropsina Munier-C halmas, 1898 (Type species:  Meandropsina vidali Schlumberger, 1898). The test is large, compressed, with the initial part of the test lenticular and consisting of a planispiral coil of chambers that are completely invo- lute and very thinly and evenly spread out over the surface of the preceding whorl. Late Cretaceous (Coniacian to Maastrichtian) (Fig. 5.22). • Nummofallotia Barrier and Neumann, 1959 (Type species:  Nonionina cretacea Schlumberger, 1900). The test is lenticular, planispiral with an umbonal plug, with early whorls evolute, later becoming involute. Late Cretaceous (Coniacian to Maastrichtian). • Pastrikella Cherchi, Radoičić and Schroeder, 1976 (Type species:  Broeckina (Pastrikella) balcanica Cherchi et al., 1976). The test is large, flat, becoming annular. The interior is subdivided by regularly arranged septula. Late Cretaceous (middle to late Cenomanian). • Perouvianella Bizon, Bizon, Fourcade and Vachard, 1975 (Type species: Orbiculina peruviana Steinmann, 1930). The test is lenticular to discoidal, planispiral with chambers never completely annular. Interiors of chambers are subdivided by radial pillars of triangular sections. Late Cretaceous (Santonian). • Pseudobroeckinella Deloffre and Hamaoui, 1969 (Type species: Pseudobroeckinella soumoulouensis Deloffre and Hamaoui, 1969). The test is large, planispiral with cyclic chambers in the final stage. Chambers are subdivided by three types of sep- tula, primary and secondary septula that are vertical to the septa, and transverse septula that are parallel to the septa. Late Cretaceous (Santonian). • Spirapertolina Ciry, 1964 (Type species:  Spirapertolina almelai Ciry, 1964). The test is discoidal, planispiral, involute becoming peneropliform, then reniform. The cyclic chambers comprise most of the test, subdivided by numerous parti- tions, those of successive chambers alternating in position. Late Cretaceous (Santonian). Family Soritidae, Ehrenberg, 1839 Tests are involute, planispiral to uncoiled evolute, flaring, annular discoids with par- tial or complete partitions. Apertures are multiple. Late Cretaceous (Cenomanian) to Holocene. • Cycledomia Hamaoui, 1964 (Type species: Edomia iranica Henson, 1948). Exhibits a discoidal, flattened and biconcave test. The early part is involute, but later is evolute and cyclic. Chambers are divided by short septula. Late Cretaceous (late Cenomanian to early Turonian). 326 Evolution and Geological Significance of Larger Benthic Foraminifera 1- Fallotia stage 2- Effect on chamber shape in Meandropsina of decrease in chamber width septulum septum 4- Fragment of meandriform stage of Meandrospina 3- Derivation of cyclical stage by increase in length of chamber Figure 5.22. The stages of developments from the involute, planispiral form of 1- 2) Fallotia to a complete meandriform cyclical form of 3- 4) Meandrospina. • Edomia Henson, 1948 (Type species:  Edomia reicheli Henson, 1948). Tests are large discoidal, planispiral and involute, later becoming evolute with cyclic cham- bers. Interiors have irregularly distributed pillars. Late Cretaceous (Cenomanian to Turonian) (Plate 5.16, figs 5- 8). • Lamarmorella Cherchi and Schroeder, 1975 (Type species:  Lamarmorella sarda Cherchi and Schroeder, 1975). The test is discoidal, planispiral to peneropline, finally becoming annular. The interior is subdivided by partial partitions. Late Cretaceous (Coniacian to Santonian). The Mesozoic Larger Benthic Foraminifera: The Cretaceous 327 • Murgella Luperto Sinni, 1965 (Type species: Murgella lata Luperto Sinni, 1965). Tests are large, planispirally coiled, later uncoiling. Chambers are divided by numerous thick radially arranged septula. Late Cretaceous (Coniacian to Maastrichtian). • Praetaberina Consorti, Caus, Frijia and Yazdi- Moghadam, 2015 (Type spe- cies:  Taberina bingistani Henson, 1948). According to Consorti et  al. (2015), Praetaberina differs from the Paleogene genus Taberina (see Chapter 6) in having a more complex internal structure with two orders of marginal partitions and aper- tures alternating with the main septula. Late Cretaceous (late Cenomanian) (Plate 5.18, figs 7- 8). • Pseudorhapydionina De Castro, 1971 (Type species:  Rhapydionina laurinensis De Castro, 1965). Tests are elongate, planispiral, later uncoiling and rectilinear. Chambers have subdivided interiors with thin radial partitions. Late Cretaceous (late Cenomanian). • Pseudorhipidionina De Castro, 1971 (Type species: Rhipidionina casertana De Castro, 1965). The test is compressed, planispiral, becoming peneropliform, then uncoiling. The interiors of chambers are subdivided by numerous short vertical septula. Late Cretaceous (Cenomanian). • Scandonea De Castro, 1971 (Type species: Scandonea samnitica De Castro, 1971). The test is enrolled, streptospiral to planispiral with uncoiled chambers, subdivided by very short radial vertical partitions. Late Cretaceous (Campanian) to Early Paleocene. • Tarburina Schlagintweit, Rashidi and Barani, 2016 (Type species: Tarburina zagro- siana Schlagintweit et al., 2016). The cylindro-c onical test is formed by a large glob- ular proloculus followed by few, planispirally coiled chambers. The chambers are partly subdivided by vertical (radial) partitions, often continuous from one cham- ber to the next. Shorter secondary partitions may be intercalated. Cretaceous (late Maastrichtian). • Zekritia Henson, 1948 (Type species: Zekritia langhami Henson, 1948). The test has lamelliform buttresses. Late Cretaceous (Turonian) (Plate 5.22, fig. 1). Family Keramosphaeridae Brady, 1884 Tests are globular with concentric chambers connected by stolons in the same series as well as those of successive series. Cretaceous (Berriasian to Maastrichtian), Miocene to Holocene • Keramosphaerina Stache, 1913 (Type species:  Bradya tergestina Stache, 1889). Shows a spherical test with a miliolide early stage. Microspheric tests have a with proloculus and flexostyle, with post- embryonic chambers with thick walls and streptospiral coiling. Later stages are regularly concentric with radially aligned chambers. Stolons connect adjacent chambers. Late Cretaceous (Coniacian to Maastrichtian). • Pavlovecina Loeblich and Tappan, 1988 (Type species: Type species: Peneroplis karreri Wiesner, 1923.The test is globular with a miliolide early stage. Post embryonic chambers are numerous and extremely irregular. Late Cretaceous (Berriasian). 328 Evolution and Geological Significance of Larger Benthic Foraminifera Figure 5.23. Morphological characteristics of key species of Nautiloculina Nautiloculinidae Planispiral, involute, biumbonate, nautiliform; chambers slowly increase in height; septa bilamellar; aperture interiomarginal, equatorial, single Periphery rounded in Sutures little if at all Nautiloculina oolithica Bajocian - Aptian axial section depressed Periphery subangular in Sutures depressed, Nautiloculina circularis Bajocian - Oxfordian axial section periphery lobulate in equatorial section Sutures not much Nautiloculina cretacea Berriasian - Early depressed, periphery Aptian smooth in equatorial section; test flat, large Superfamily MILIOLOIDEA Ehrenberg, 1839 Tests are coiled commonly with two or more chambers arranged in varying planes about the longitudinal axis; later they may become involute. Advanced forms may have secondary partitions within the chambers. Late Triassic (Norian) to Holocene. Family Nautiloculinidae Loeblich and Tappan, 1985 Tests are free, lenticular, planispiral, and involute with secondary thickening in the umbilical region. Apertures are equatorial. Middle Jurassic to Late Cretaceous. • Nautiloculina Mohler, 1938 (Type species: Nautiloculina oolithica Mohler, 1938) (see Chapter 4). Jurassic (Late Bajocian) t\o Early Cretaceous (Plate 4.9, figs 6A, 7- 12; Fig. 5.23). 5.3 Biostratigraphy and Phylogenetic Evolution Cretaceous carbonate formations show sequences of foraminiferal faunas, recogniz- able in thin sections of marine limestones, which are of short range and so enable broad stratigraphic correlations to be made (see Charts 5.1 and 5.2 online). Three main orders flourished at different times: • Firstly, the agglutinated larger foraminifera, which made their first appearance in the Jurassic and particularly thrived during the Early Cretaceous, where they occur with many different short ranged groups. • Most of these forms disappeared before the Cenomanian, where they were replaced by mainly porcelaneous miliolide forms. • Many of the miliolides disappeared in turn towards the end of the Cenomanian and were replaced gradually by a thriving population of hyaline calcareous rotaliines and orbitoids. The Mesozoic Larger Benthic Foraminifera: The Cretaceous 329 These three orders are discussed in detail below. 5.3.1 The Textulariides of the Cretaceous In the Cretaceous, agglutinated foraminifera retained some of the features of their Jurassic forebears but also acquired new features, which became important in the later Cretaceous. Four main major superfamilies are of most biostratigraphic value (see Fig. 5.24): • the Textularioidea, • the Lituoloidea, • the Ataxophragmioidea, and • the Orbitolinoidea. Some of these forms exhibit complex labyrinthic features, while others are parti- tioned; many were quite short ranged and so are valuable as index fossils. They have been closely studied by many workers, and detailed evolutionary lineages have been established by, amongst others, Banner (1970), Banner et al. (1990), and Simmons et al. (2000). Cretaceous Textularioidea seem to have evolved independently from those of the Jurassic (see Chapter  4). They are represented by the Chrysalinidae (see Fig.  5.25), Period Epoch and ORBITOLINOIDEA TEXTULARIOIDEA LITUOLOIDEA ATAXOPHRA. Stage Jurassic Figure  5.24. The biostratigraphic range and diversity of the main agglutinated superfamilies found in Tethys during the Cretaceous. 145.0 100.5 66 Age (Ma) Cretaceous Early Late Tith. Ber. Valang. Haut. Barr. Apt. Alb. Cen. Tur. Con. Sant. Camp. Maast. 330 Evolution and Geological Significance of Larger Benthic Foraminifera Period - CHRYSALIDINIDAE Stage Chrysalidina Dukhania Praechrysalidina Jurassic Figure 5.25. The evolution of the Chrysalinidae in the Cretaceous. which evolved directly from a simple Verneulinoides in the early Cretaceous by the appearance of the triserial, non-p illared Praechrysalidina. The development of pillars in the Aptian and Albian was gradual, and it was not before the Cenomanian that com- pletely pillared forms were developed. These included the rapidly tapering, terminally biserial, Dukhania and the slowly tapering, terminally triserial, Chrysalidina. These forms had no descendants and are confined to central Tethys (Banner et al., 1991). On the other hand, their possible ancestor Praechrysalidina spread throughout western and central Tethys throughout the Early Cretaceous and might have been ancestral to Accordiella in the Coniacian to Santonian (see Fig. 5.25, and Charts 5.1 and 5.2). Cretaceous lituolids (see Fig. 5.26) vary in shape from planispiral (e.g. Buccicrenata, Choffatella, Fig. 5.26 and Chapter 4) to low trochospiral (e.g. Sornayina, Fig. 5.26), becoming flabelliform (e.g. Pseudochoffatella, Plate 5.3, fig.  13), becoming spheri- cal (e.g. Reticulinella, Plate 5.4, fig.3) to fusiform (e.g. Loftusia, Plate 5.8) in the Late Cretaceous. However, the main morphological trend in this superfamily is the develop- ment of labyrinthic structures along a number of separate lineages, and the develop- ment of central pillars in genera such as Spirocyclina (Plate 5.3, fig.  4). Cretaceous lituolids have their ancestors in the Jurassic. In the Sinemurian, a descendant of a species of Ammobaculites, with a single areal aperture, developed an alveolar wall with the septa remaining solid, to give rise to Everticyclammina. The early Everticyclammina, with a coarsely labyrinthic hypodermis (e.g. E. praevirguliana, Plate 4.16, fig. 3), evolved in the Early Cretaceous into individuals with a smooth, regularly structured hypoder- mis with much finer and more regularly formed alveolae (e.g. E. kelleri, Plate 4.16, 145.0 101 Age (Ma) Cretaceous Early Tith. Ber. Valang. Haut. Barr. Apt. Alb. Cen. Tur. The Mesozoic Larger Benthic Foraminifera: The Cretaceous 331 fig.  4). Throughout the Cretaceous, the alveoles of Everticyclammina become regu- larly and tightly bifurcating. In the Barremian- Aptian and before its disappearance, Everticyclammina developed forms (E. greigi, Plate 4.16, fig. 5) in which the lower parts of the adult septa become oblique, then tangential to the spiral suture, and thicken and coalesce to form imperforate “basal layers” to the chambers. The reduction of the solid septa produced the Albian- Cenomanian Hemicyclammina (Fig.  5.26), in which the upper parts of the septa are thinned and meet the test periphery (almost at right angles), so that the chambers become almost quadrangular in equatorial sec- tion (e.g. H. whitei, Plate 5.6, figs 14-1 5). Quite independently from Hemicyclammina and in the Barremian-A ptian, Everticyclammina evolved into evolute and streptospiral Hottingerita (see Fig. 5.26; Plate 5.2, fig.2). Another lituolid lineage thrived in the Cretaceous parallel to the Everticyclammina lin- eage, namely that of Buccicrenata. Unlike Everticyclammina, where the septa remained solid and angularly differentiated from the walls, Buccicrenata had septa with alveolar extensions of the lateral chamber wall, and seems to have directly descended from the Lituola group camerata in the Kimmeridgian, quite independently (see BouDagher- Fadel, 2000, 2001). In the Cretaceous, in Buccicrenata the sequence of alveolae became finer, the walls themselves became thinner but all the microspheric forms are able to become rectilinear after the first two whorls for the final three chambers. The single aperture is areal in all species, but became more compressed and slit-l ike as the forms became stratigraphically younger and the microspheric chambers became terminal. Pseudocyclammina had similarly evolved from Lituola in the late Sinemurian by developing an areal cribrate aperture and a coarsely alveolar wall. These forms passed through the Jurassic-C retaceous boundary and persisted to the Aptian. The Cretaceous forms developed alveolar layers more strongly, with several independent evolutionary modifications occurring; these include trends towards increased com- plexity in the hypodermis (see Banner, 1970), a trend which was independently fol- lowed and reached acmes in Martiguesia, where delicate pillars bridge from septum to septum; Choffatella, which developed tighter coiling and septa as complex and thick as the hypodermis; and Torinosuella with increased test compression and coil- ing rate. Parallel to these lines of evolution many forms evolved in the middle to Late Cretaceous, with flabelliform forms and even more coarsely agglutinated epidermis, such as Pseudochoffatella. In the Cenomanian Choffatella gave rise to Spirocyclina (see Fig. 5.26), with a flat initial spire, becoming peneropliform, with a coarse reticulate hypodermis and a few scattered central pillars. In turn Spirocyclina may have given rise to Sornayina in the Coniacian, by increased trochospirality and the development of more regularly spaced and stronger transverse sub-h ypodermal septulae (Banner, 1970). Anchispirocyclina (Plate 4.15, figs 8- 9) may have evolved from Pseudocyclammina in the Oxfordian by developing a central zone filled by a complex reticulum of densely spaced pillars. Anchispirocyclina did not become extinct before the Campanian. In the Late Cretaceous, lituolids became large and internally very complicated. Alveoliniform, globose spherical genera, such as Reticulinella, with interseptal hori- zontal lamellae, appeared in the Cenomanian, while Loftusia with interseptal pillars and the strong hypodermal development of Martiguesia appeared in the Maastrichtian and attained immense sizes, while acquiring an increased tightness of coiling leading 332 Evolution and Geological Significance of Larger Benthic Foraminifera Period, Epoch, Stage Jurassic Figure 5.26. The evolution of the Cretaceous lituolids. to fusiform tests. However, as with so many other forms, the end of the Cretaceous saw the extinction of most of the lituolids. The Ataxophragmioidea were common in the Cretaceous. They were trochospiral, tightly or loosely coiled with high spires, or conical to sub- flabelliform (see Fig. 5.27). The main morphological trend in this superfamily is the subdivision of the periphery of the chambers by radial or curved partitions and buttresses, forming occasionally small chamberlets. 145 100.5 66 Age (Ma) Cretaceous Early Late Tith. Berr. Valang. Haut. Barr. Apt. Alb. Cen. Tur. Con. Sant. Camp. Maast. Loftusia Martiguesia Sornayina Spirocyclina Reticulinella Pseudochoffatella Choffatella Torinosuella Pseudocyclammina Buccicrenata Lituola Everticyclammina Hemicyclammina Hottingerita Anchispirocyclina The Mesozoic Larger Benthic Foraminifera: The Cretaceous 333 Figure 5.27. The evolution of the Ataxophragmioidea in the Cretaceous. In the Campanian, the Cretaceous pfenderids lost their continuous composite cen- tral columella and evolved into the ataxophragmids. The tightly coiled Ataxophragmium acquired peripheral partitions, (e.g. Opertum) and buttresses (e.g. O.  orbignyna). Quite independently in the Albian, the trochospirally enrolled simple Arenobulimina acquired arched partitions and buttresses that meet and form irregular chamberlets as in Voloshinoides (see Fig. 5.8). Period, Epoch, Stage pfenderinid ancestor Jurassic 1 4 5 1 0 0 . 5 6 6 A g e ( M a ) C r e t a c e o u s L a t e E a r l y T i t h . B e r r . V a l . H a u t . B a r r . A p t . A l b . C e n . T u r . C o n . S a n t . C a m p . M a a s t . V e r c o r s e l l a C u n e o l i n a S a b a u d i a P s e u d o t e x t u l a r i e l l a D i c y c l i n a A t a x o p h r a g m i u m O p e r t u m O r b i g n y n a A r e n o b u l i m i n a V o l o s h i n o i d e s 334 Evolution and Geological Significance of Larger Benthic Foraminifera Another parallel evolution from a pfenderid ancestor saw the development of the cuneolids in the Early Cretaceous (see Fig. 5.27). In the Early Cretaceous, the conical pfenderids lost their central columella and the chambers became divided by vertical partitions, or rarely by horizontal partitions (as in Vercorsella), or by both vertical and horizontal partitions as in Cuneolina. In the Cenomanian, ataxophragmids with a tro- chospiral initial stage subdivided their chambers by vertical and horizontal partitions (e.g. Pseudotextulariella). Annular growth was achieved in the Ataxophragmioidea, and Dicyclina has two layers with radial partitions. Another important agglutinated group from Early Cretaceous in the Tethyan realm are the orbitolinids (see Figs 5.10– 5.12). The orbitolinids are characterized by conical tests that are usually a few millimeters in height and diameter (although they can attain diameters of 5 cm or more). They evolved from forms such as Urgonina, with a trocho- spiral early stage but simple uniserial later stage with chambers lacking partitions, to forms that are partially subdivided by radial partitions or with pillars, Coskinolinoides. These forms evolved into forms with peripheral tiered rectangular chamberlets, Paleodictyoconus, which eventually lost its initial coil and filled its centre with thick radial partitions, as in Orbitolina (see Fig. 5.10). The earliest formed chambers of the Period, Epoch, Stage Jurassic Figure  5.28. The evolution of the orbitolinids. Drawing of embryonic apparatus after Simmons et  al. (2000). 145.0 100.5 66.0 Age (Ma) Cretaceous Early Late Tith. Ber. Val. Haut. Barr. Apt. Alb. Cen. Tur. Con. Sant. Camp. Maast. Palorbitolina Praeorbitolina Mesorbitolina Orbitolina Conicorbitolina The Mesozoic Larger Benthic Foraminifera: The Cretaceous 335 megalospheric generation can form a complex embryonic apparatus, which can be divided into protoconch, deuteroconch, sub- embryonic zone and peri-e mbryonic cham- berlets, which is the most important feature for their taxonomic division (see Fig. 5.28). The orbitolinid test is defined by the shape of the embryonic apparatus, and by the size and shape of the chamber passages that can be seen in tangential sections. The chamber passages are formed in the radial part of the central zone of each chamber layer (Figs 5.11 and 5.12), where each chamber passage is subdivided by vertical main partitions, which are prolongations of the vertical main partitions of the marginal zone. In cross section, these can be triangular, rectangular or oval, or can show a gradation between shapes (Schroeder, 1975). In the radial zone of Orbitolina, the stolons are arranged in radial rows alternating from one chamber to the next one (see Figs 5.11 and 5.12). Their alternating position obliges the protoplasm to flow in oblique direction (Hottinger, 1978). In axial section, the embryo is located at the apex of the cone, followed by a series of discoidal chamber layers. In transverse section, the chambers are seen divided into a marginal zone, with subepidermal partitions and a central zone with radial partitions. The radial partitions in Orbitolina thicken away from the periphery and anastomose in the central area, producing an irregular network. The earliest formed chambers of the megalospheric generation can form a complex embryonic apparatus that can be divided into a protoconch, deuteroconch, a sub- embryonic zone and peri- embryonic chamber- lets depending on the genera involved. It evolved from a simple apparatus, consisting of a large globular fused protoconch and deuteroconch, followed by peri- embryonic chambers as in Palorbitolina, to an embryonic apparatus divided into a protoconch and deuteroconch but not completely divided sub-e mbryonic zone, as in Praeorbitolina. This latter evolved in turn into forms in which the deuteroconch and sub- embryonic zone are more or less of equal thickness, as in Mesorbitolina. In Conicorbitolina the marginal zone became extensively divided by vertical and horizontal partitions, while in Orbitolina the deuteroconch is highly subdivided and of much greater thickness than sub- embryonic zone (see Fig. 5.28; Schroeder, 1975; Hottinger, 1978; Simmons et al., 2000; BouDagher- Fadel, 2008; Schroeder et al., 2010; BouDagher-F adel et al., 2017). The orbitolinids are very useful biostratigraphic markers in mid-C retaceous Tethyan carbonate platforms (Henson, 1948; Schroeder, 1975; BouDagher- Fadel et al., 2017). They have short ranges and are easily identified in thin sections (e.g. see Figs 5.28, 5.29 and 5.30). Together with the miliolides (see below), they enrich Cretaceous biota, but unlike the miliolides they show provincialism. 5.3.2 The Miliolides of the Cretaceous For the first time in their history, with the appearance of the Alveolinoidea, the mili- olides exhibited a fusiform morphology. In an example of convergent evolution, the external appearance the alveolinids closely resemble the fusulinides of the Permian, but they can be seen to be quite distinct when studied in axial and equatorial (median) sec- tions. The alveolinids differ fundamentally from the fusulinides in that they have an 336 Evolution and Geological Significance of Larger Benthic Foraminifera Praeorbitolina - Mesorbitolina Palorbitolina - Orbitolinopsis - Conicorbitolina Palorbitolinoides Paleodictyoconus 3 2 1 h 4 g 3 f 2 e 1 d 4b c 4a 3 b 2 a 1 Figure 5.29. Phylogenetic evolution of Tibetan orbitolinids (after BouDagher-F adel et al., 2017). Planktonic Zones after BouDagher- Fadel (2015). 1 2 2 . 9 1 0 0 . 5 A g e ( M a ) 1 1 3 . 0 M i d - C r e t a c e o u s P e r i o d A p t . A l b . C e n . S t a g e P l a n k t o n i c z o n a t i o n T L K 1 b i o z o n a t i o n s P r a e o r b i t o l i n a c o r m y i P r a e o r b i t o l i n a w i e n a n d s i P r a e o r b i t o l i n a c f . w i e n a n d s i M e s o r b i t o l i n a l o t z e i M e s o b i t o l i n a p a r v a M e s o r b i t o l i n a t e x a n a M e s o r b i t o l i n a s u b c o n c a v a M e s o r b i t o l i n a b i r m a n i c a M e s o r b i t o l i n a a p e r t a P a l o r b i t o l i n a l e n  c u l a r i s P a l o r b i t o l i n o i d e s o r b i c u l a t a P a l o r b i t o l i n o i d e s h e d i n i O r b i t o l i n o p s i s s p . A P a l e o d i c t y o c o n u s s p . A C o n i c o r b i t o l i n a c f . c u v i l l i e r i C o n i c o r b i t o l i n a s p . A The Mesozoic Larger Benthic Foraminifera: The Cretaceous 337 Figure 5.30. Evolutionary trends in the orbitolinids. imperforate, porcelaneous wall structure, which consists of an external lamina, usually light in colour and an internal darker basal layer, and spiral septula. The basic mor- phology of the alveolinids is shown in Fig. 5.21. Unlike the fusulinides, septal folding does not occur and the spiral septula reach the floor across the chamber. Deposition of secondary calcite (flosculinisation) occurs more evenly across the floor of the chamber, rather than being almost entirely concentrated in the axial zone as in the fusulinides. The embryonic apparatus consists of a spherical proloculus followed by a spiral tube (flexostyle), succeeded in some genera by a miliolide nepion (pre-a dult stage). In the adult, there are numerous chambers planispirally coiled around an elongate axis, each chamber being broken into tubular chamberlets by secondary septa (septula) that run in the direction of coiling. Communication through the chambers is maintained through the vertical septula (canals/p assages). These may be situated in front of the septum (post- septal) or behind it (pre-s eptal). All Cretaceous genera have pre-s eptal canals, a few genera, such as Cisalveolina (Plate 5.22, figs 3- 4), have post- septal passages. Archaealveolina, a subglobular alveolinid, with early streptospiral whorls and a single row of apertures, makes its first appearance in the Aptian. Ovalveolina and Praealveolina General trends of orbitolinid development Marginal radial Megalospheric Dominant shape (MR) and reculate Sub-epidermal cells embryonal apparutus and size zones (RZ) MR and RZ very 4-6 per annular More or less narrow; RZ complex interval. Irregularly discoidal with Large, apical with arranged concave base; over complex "roseŒe" and with large 5mm diameter. embryonal area Large, apical with simple "roseŒe"; supra-embryonal area very small or absent Low conical and depressed +/_ 5mm Small, excentric diameter;forms with simple with concave base "roseŒe"; no supra- dominant embryonal area Very small, excentric MR and RZ 2-4 per annular Conical, less than without "roseŒe"; relavely well interval, regularly 5mm diameter; secondary chambers developed; RZ arranged forms with concave possibly in spiral form. relavely simple base rare H a u t e r i v i a n - B a r r e m i a n A p  a n - A l b i a n C e n o m a n i a n 338 Evolution and Geological Significance of Larger Benthic Foraminifera appear in the Albian, while most of the alveolinids appear in the Cenomanian, many with short ranges. Very few appear in the Late Cretaceous, such as Subalveolina in late Santonian- Campanian and Senalveolina in the early Campanian. Helenalveolina and Subalveolina, alveolinids with a septulate endoskeleton lacking pillars and with streptospiral early stage and planispiral- involute adult stage, may be linked phylogenetically to Pseudonummuloculina (a simple miliolide foraminifera with streptospiral-i nvolute coiling) of mid-C retaceous age (Hottinger et al., 1989). Subalveolina is interpreted by Hottinger et al. (1989) as an endemic offspring of American immigrants. Early chamber arrangement in alveolinids was shown to be linked with proloculus size by Pêcheux (1984, 1988) for Cenomanian streptalveolinids. The rhapydioninids may have evolved from Ovalveolina via Pseudedomia in the Cenomanian (Hamaoui and Fourcade, 1973). Pseudedomia resembles Ovalveolina in its globular initial part, but develops compressed curved chambers in its final whorls. Hamaoui and Fourcade (1973) also suggest that Chubbina arose from intermediate forms of Ovalveolina. The trend towards uncoiling, and becoming cylindrical, con- tinued through Murciella and Rhapydionina in the Late Cretaceous (see Fig.  5.21). The latter was described by Stache (1913) from the Late Cretaceous of the Adriatic Platform as spiroliniform shells with a planispiral- involute nepiont followed by a uniserial- cylindrical adult stage. However, he also used the name Rhipidionina to descibe peneropliform shells with flaring adult chambers producing a flattened fan reaching a semicircular outline. However, Stache’s original specimens have identical structural patterns (Reichel, 1984) and are one dimorphic species, representing respec- tively a megalospheric and a microspheric generation (Hottinger, 2007). The Fabulariidae, with round trematophore, were common in the Late Cretaceous and arose directly from the simple Quinqueloculina group (Haynes, 1981) or a true miliolide ori- gin exemplified by Idalina (see Chapter 6). Idalina, with a quinqueloculine to biloculine test, a round trematophore but with a bar- like tooth, acquired incomplete partitions in Periloculina (see Hottinger et al., 1989) and the overlapping throughout in Lacazina (Plate 5.22, fig. 6). The Soritoidea of the Cretaceous are separated by a considerable stratigraphic gap from those of their younger descendants in the Eocene and they seem to have origi- nated from different ancestors. Haynes (1981) suggested that they may have arisen in the Senonian from a simple foraminifera Praepeneroplis, with a close- coiled and tri- angular chambers. The main morphological trend in the resulting superfamily is the uncoiling of the planispiral test to become flabelliform and cyclical. The development of Meandropsina is an example of the trend of evolution followed by the Soritoidea in the Late Cretaceous. Each chamber is divided into numerous chamberlets by transverse partitions that are in continuous alignment from chamber to chamber. The resulting composite lines spiral outward from the centre of the test and maintain approximately the same spacing by periodic intercalation of new lines of septa. At this stage Meandropsina is exactly the same as Fallotia (see Fig. 5.22, stage 1). As growth continues, however, the volume of the individual chambers does not increase at a sufficiently rapid rate to maintain the same shape when stretched across the growing bulk of the test, the radial septa necessary fall closer and closer together and, as they approach parallelism, tend more and more to assume a shape like that of the outward spiraling lines of secondary septa (Fig. 5.22, stage 2). The chambers are now long and thin, and they appear at this stage to have reached a stable minimum width that is maintained throughout the remainder of the individual’s development. Since the width of the chambers is now stable, any further increase in size must take The Mesozoic Larger Benthic Foraminifera: The Cretaceous 339 the form of an increase in length. This eventually results in a cyclical form of growth in which the chambers form concentric rings that are no longer embracing to the cen- tre of the test (Fig. 5.22, stage 3). In the final stage, the chambers no longer lie astride the periphery, but they withdraw completely to one side and wander haphazardly over the surface of the previously formed test in wavy meandriform line (Fig. 5.22, stage 4). In conclusion, the radial partitions subdividing the long narrow chambers into cham- berlets and the tendency toward uncoiling, becoming flabelliform and annular in the final stage of the test, are two important trends followed by the Soritoidea in the Cretaceous. 5.3.3 The Rotaliides of the Cretaceous The rotaliides of the Cretaceous developed important evolutionary trends, which reappeared independently again in the Paleogene and Neogene. Large rotaliides were Siderolites (Maastrichtian) Fissoelphidium (Maastrichtian) Rotalia (Coniacian - Eocene) Bipertorbis (Turonian - Eocene) Figure 5.31. The possible evolution of the calcarinid Siderolites from a simple rotaliid test, by developing spines and multiple foramina. 340 Evolution and Geological Significance of Larger Benthic Foraminifera derived from different families of smaller rotaliides by opening up of the umbilical area and its infilling by pillars. Rotalia appeared in the Late Cretaceous possibly from a discorbinellid, Bipertorbis (see Fig.  5.31), which appeared in the Turonian with a small plug filling the umbilical area. In Rotalia the plug is enlarged and composed of thick lamellar pillars with septal fissures. Rotalia in turn evolved rapidly given rise to the different families of the Rotalioidea. An example (Fig. 5.31) of one possible line of evolution from Rotalia to a calcarinid test, is the development of involute, bilater- ally symmetrical forms, Fissoelphidium in the Maastrichtian, which may have evolved peripheral spines and multiple foramina in Siderolites. Foraminifera with an orbitoidal growth first appeared in the Late Cretaceous. They have been studied extensively by many authors, and one of the most comprehensive studies was that of van Gorsel (1978). All orbitoidoids were derived from small ben- thic foraminifera with a simple, spiral chamber arrangement. A typical orbitoidal test (Fig. 5.32) varies from globular to lenticular, and is composed of a median layer, con- sisting of an initial stage surrounded by concentrically arranged equatorial chambers, flanked on either side by layers of lateral chambers. The central part of the equatorial layer is composed of the embryonic chambers that are surrounded in turn by the peri- embryonic chambers or the neanic phase. The arrangement of the embryonic chambers is one of the most important characters needed for the identification of an individual species. The embryonic chambers usually comprise two chambers surrounded by a relatively thick wall, a protoconch and a ren- iform second chamber or deuteroconch (see Fig. 5.33). In many species of Orbitoides there are four chambers within the embryonic wall (Fig. 5.34). The peri- embryonic chambers or nepionic chambers immediately surround Equatorial chambers in horizontal section Lateral chambers Equatorial (median) chambers Embryonal in vertical section in vertical section. chamber Figure 5.32. Schematic sketch of an orbitoidal foraminifera (modified after Carpenter et al., 1862). The Mesozoic Larger Benthic Foraminifera: The Cretaceous 341 D D ax ax P P D ax P ax e d e d D ax ax P ax P ax Figure 5.33. Schematic drawing showing chamberlet cycles, each cycle corresponding to one chamber and the change from spiral to concentric growth in orbitoidal foraminifera, as a consequence of the introduction of retrovert apertures, modified after van Gorsel (1978). P = protoconch, D = deuteroconch, ax = auxiliary chambers, d = adauxilliary chambers, e = epiauxilliary chambers. the embryon, the third chamber is the auxiliary chamber. In advanced forms there may be two auxiliary chambers. In Lepidorbitoides, adauxiliary chambers (small chambers originating from stolons in the deuteroconchal wall) are present, while in Orbitoides, the chambers originating from the stolons in the thick embryonal wall are usually called epi- auxiliary. In younger species there is an increase in the number of these chambers. In the spiral growth, chambers with one aperture, “uni-a pertural chambers” are one chamber- forming apertures, and their number is often designated by the parameter Y, while chambers with two apertures give rise to two chambers, one at each aperture. In the equatorial layer the shape of the chambers varies from open arcuate to ogival or spatulate (see Fig. 5.34). The main features which differentiate Orbitoides from Lepidorbitoides can be sum- marized as follows (see Figs 5.17, and 5.34): 1. The nucleoconch of Orbitoides is larger than that of Lepidorbitoides, and con- sists of a porous, thick-w alled spheroidal body subdivided internally into two or more chambers that are separated from each other by thin, straight walls. The nucleoconch of Lepidorbitoides is smaller than that of Orbitoides and consists of two chambers, of 342 Evolution and Geological Significance of Larger Benthic Foraminifera Pseudorbitoides Maastrichtian Lepidorbitoides Maastrichtian Orbitoides Late Santonian - Maastrictian Figure 5.34. Sketches showing important features of different orbitoidal genera. which the second (deuteroconch) is the larger and partially embraces the first chamber (protoconch). 2. Individual chambers in the rings of simultaneously- formed chambers in the median layer are widely- spaced in Orbitoides and crowded together in Lepidorbitoides. They tend to be wider than high in Orbitoides and higher than wide in Lepidorbitoides. 3. Cross- sections through the layers of lateral chambers in Orbitoides show them as a series of low arches, the arches of one layer overlapping the arches of the next layer. The lateral chamberlets in Lepidorbitoides are rectangular in section and are lined up in vertical tiers. The origin of the orbitoids has been the subject of many studies and there are many different interpretations. Unlike previous workers, such as Küpper (1954), who regarded the initial chambers of the non- canaliculate Orbitoididae to be biserial, van Gorsel (1978) interpreted them as spiral in the microspheric generation, becoming irregularly alternate. He noted that the equatorial chambers of early Orbitoides spe- cies were acervuline, thus suggesting an attached mode of life. He therefore supported the idea that non- canaliculate orbitoids were derived from an unknown “Cibicides”. Here I suggest that a planorbulinid ancestor, such as Archaecyclus, may have been the The Mesozoic Larger Benthic Foraminifera: The Cretaceous 343 ancestor of the orbitoids. The evolution from Monolepidorbis seems to have happened gradually, by the subdivision of the solid walls of Monolepidorbis by lateral cham- berlets (see Fig. 5.16). However, in Torreina, the embryon is similar to Orbitoides and Omphalocyclus, but no median layer can be distinguished; this form is only found in the Caribbean. The Helicorbitoides- Lepidorbitoides lineage seems to have evolved from the helical Pseudosiderolites, which in turn evolved from a small rotaliide (van Gorsel, 1978) by developing incipient lateral chambers. Secondary equatorial chambers were added between the primary spires in Helicorbitoides before acquiring well developed lateral chamberlets in Lepidorbitoides. This evolutionary lineage provides one of the best examples of nepionic acceleration, where the ancestral spiral is gradually reduced and a concentric growth pattern is reached progressively. The Lepidorbitoides lineage shows other evolutionary trends, such as a systematic increase with time of adauxiliary cham- bers and the protoconch- deuteroconch diameter ratio (Caus et al., 1988). The Caribbean forms (see Fig.  5.18) evolved from the cosmopolitan genus Sulcoperculina (van Gorsel, 1978) by the acquisition of lateral chambers and the fur- ther development of the system of incipient radial rods, as in Sulcorbitoides. This latter acquired protruding radial plates in Vaughanina (Plate 5.35), stellate growth pattern in Aktinorbitoides and arcuate secondary equatorial chambers without the radial plates in Pseudorbitoides. Sulcoperculina evolved into Orbitocyclina by suppressing the system flange with protruding radial plates Figure 5.35. The main features exhibited by a solid specimen of Vaughanina sp., Late Campanian, Cuba, UCL coll. Scale bar = 0.3mm. 344 Evolution and Geological Significance of Larger Benthic Foraminifera of plates, and by acquiring lateral and curved equatorial chambers, which in turn it evolved to a stellate growth pattern in Asterorbis. Several more unrelated and rapidly evolving lineages of rotaliines thrived in the Late Cretaceous, but most of them became extinct at the end of the Maastrichtian. Due to their rapid evolution and short ranges, the orbitoids are excellent biostratigraphic markers for the Cretaceous. 5.4 Palaeoecology of the Cretaceous Larger Foraminifera The cooling trend of the Late Jurassic continued into the Early Cretaceous, however the climate warmed up fairly rapidly after the middle Berriasian, triggering an increase in the level of oxygen in the atmosphere, which is inferred (see Ward, 2006) to have steadily increased throughout this period (see Fig.  3.10). A  tropical climate existed throughout the whole subsequent Cretaceous period over the Tethyan realm (see Fig. 5.36), which was the main breeding ground of the larger benthic foraminifera and rudist reefs. While coral reefs were predominant in the Jurassic, just as they are today, during the early Cretaceous, a superheated, hypersaline ocean- climate zone favoured the proliferation of rudists over scleractinian corals (Kauffman and Johnson, 1988; Johnson et al., 1996). The larger benthic foraminifera which survived the minor Jurassic-C retaceous boundary crisis were mainly robust, shallow, clear- water forms. They consisted mainly of agglutinated foraminifera with large intramural alveoles, such as Everticyclammina, which could tolerate water rich in argillaceous suspensions. A  very few forms with Figure 5.36. Climate zones in Early Cretaceous (140 Ma). The Mesozoic Larger Benthic Foraminifera: The Cretaceous 345 narrow alveoles and a regularly labyrinthic hypodermis (e.g. Choffatella, see Chapter 4 and Plate 5.5, figs 8- 9) were also present and occupied a deeper neritic palaeoenviron- ment, with reduced illumination, in which codiacean and dasyclad algae were rare or absent (Banner and Whittaker, 1991). By the Valanginian-H auterivian, agglutinated foraminifera, largely adapted in their evolution to the increase in oxygen level and the superheated oceans, increased the sizes of their tests and the complexity of their wall structures. The Chrysalidinidae developed canaculi which had the same function as that of the alveoles in the Cyclamminidae. They facilitated ionic exchange between intrathalamous cytoplasm and external sea- water, allowing the foraminifera to survive and proliferate under different conditions. By the Barremian, new forms and a major new lineage appeared in Tethys, the agglutinated orbitolinids. These robust forms had the ability to survive in many shal- low carbonate environments (Arnaud-V anneau, 1980), however, they were most com- mon in the outer platform (Vilas et al., 1995). They strengthened their conical test by subdividing it into many small chamberlets, which housed within their walls symbiotic algae. By studying the size and shape of their test, Masse (1976) deduced that they had a free, epifaunal mode of life. They lived lying on the substrate, on the flat base of their conical test, by their apertural face. However, the small primitive forms of the orbitolinids may have been epiphytic (Arnaud- Vanneau, 1975). Using associated algae, Banner and Simmons (1994) noted that Palorbitolina lenticularis was most com- mon in sediments thought to be deposited at depths of between 10 and 50m. Simmons et al. (1997) noted that muddy, orbitolinid- rich beds, with large, flat, orbitolinids seem to be characteristic of transgressive deposits, while more conical forms thrive in the shallowest water. The relationship between orbitolinid shape and palaeobathymetry seems to mimic that observed for Holocene larger foraminifera (e.g. Operculina) by Reiss and Hottinger (1984) (see Chapter 7). In association with the orbitolinids and Choffatella, mid- Cretaceous reefs were typically composed of rudists, a group of aber- rant bivalve mollusks (Lehmann et al., 1999). Corals may also be associated with rudist reefs, but possibly only those reefs far from the equator. It has been argued at some length that equatorial surface waters during the Aptian approached 30°C, and were lethal to corals. Thus, this first generation of rudist reefs may have obtained its start because the seas were simply too hot for corals to thrive. Rudist diversity began to escalate in the Aptian- Albian time, but collapsed abruptly end of the Maastrichtian, and the entire taxon became extinct at the end of the Mesozoic. The increase in the number of foraminiferal forms having large alveoles in the early to late Aptian may have been as a consequence of adaptation to the adverse conditions during the anoxic events that accompanied this interval. The dramatic increase of car- bon dioxide in the atmosphere (Gradstein et al., 2004), which may have been triggered by the Ontong-J ava volcanism, must have led to an increase in temperature, with pos- sible subsequent release of marine clathrates (methane hydrates) leading a greenhouse period and water anoxia (e.g. Kerr, 2006). Within the Cenomanian, another important group made its first appearance, namely the alveolinids. In comparing the occurrences of similar Holocene forms, we can safely deduce that these fusiform types, with large alveoles (Fig. 5.37A), were found in sedi- ments deposited in warm, shallow water. In fact, all alveolinids are regarded as neritic (inner shelf) by Reichel (1964), restricted to littoral, tropical, protected shelf and reef 346 Evolution and Geological Significance of Larger Benthic Foraminifera A B Figure 5.37. Thin sections of a A) limestone containing Lacazina and a few Monolepidorbis, Santonian of Spain; B) sandy limestone containing Orbitoides and a few Siderolites, Maastrichtian of Spain. Scale bars = 1mm. shoals (Dilley, 1973; Hottinger, 1973). Their appearance might have been caused by one of the most widespread oceanic anoxic events (OAE) at the Cenomanian– Turonian boundary (93 Ma), which may in turn have been triggered by the Madagascar and Caribbean provinces basaltic events (see discussion below). In the Late Cretaceous, the global climate was warm and moist but cooler than that in Jurassic (210 to 140 Ma). The arrangement of the continents and oceans encouraged warm waters at the equator to circulate freely. During this period larger foraminifera became cosmopolitan and colonized deeper waters. They occupied a wide tropical- subtropical zone, as cool water was only found near the polar regions. Coral reefs grew 5 to 15 degrees closer to the poles than they do today. The niches favoured by the larger foraminifera, which became extinct towards the end of the Cenomanian (see discus- sion below), were gradually filled by new, highly evolved rotaliides, having an orbitoi- dal test (Fig. 38B). The introduction of this orbitoidal character, and the subsequent The Mesozoic Larger Benthic Foraminifera: The Cretaceous 347 development of equatorial and lateral chamberlets, may have given these foramin- ifera a selective advantage in adverse environments over the smaller ones. Orbitoids now achieved large sizes, which might have reflected the high growth rates that can be sustained by symbiotic species, with a tendency to increase the size of their proto- conchs. Large- sized protoconchs are considered to be advantageous in speeding up test growth (van Gorsel, 1975) and could be related to high light intensity in the photic zone (Fig. 5.37B; Drooger and Raju, 1973). The tendency by the orbitoids towards radial symmetry may be closely related to the more sedentary life possible for symbiotic forms (Chaproniere, 1975), as it is generally accepted that orbitoids thrived in shallow tropical and sub- tropical seas, in areas with little or no clastic influx (van Gorsel, 1975). However, van Gorsel (1975) noted that not all of the genera had the same ecological requirements. For most genera, a water depth of 50m is the limit of habitation, but in the Caribbean, Pseudorbitoides appeared to have lived in deeper fore- reef environments, while Orbitoides and Vaughanina (often associated in many localities with Asterorbis) are found in sediments deposited in back-r eef environments. In Europe, Omphalocyclus probably lived in shallower waters than Lepidorbitoides and Clypeorbis, but their depth ranges overlap (van Gorsel, 1975, 1978). Some rotaliides developed spines, as in the calcarinids, to spread the weight of their shell, as in the case of Siderolites in the Maastrichtian. The larger benthic foraminifera in the Cretaceous occupied various facies, close to wave base, in lagoonal settings or in sediments overlying or underlying massive reefs (Noujaim Clark and BouDagher- Fadel, 2001). Different associations of foraminifera belong to different biofacies and different depositional environment; the Nezzazatidae and the Cuneolinidae are found mainly in lagoonal facies together with Miliolidae and Verneulinidae. However, their absence points to a tidal environment Orbitoididae Orbitolinidae Chofatellidae Alveolinidae Everticyclamminidae Ataxophragmiidae Nezzazatidae Reef Cuneolinidae Small rotaliines Rudist bank Small miliolines Base of wave action Base of wave action Forereef shelf Backreef shelf Sea bed Lagoon Figure  5.38. The special distribution of the Cretaceous larger benthic foraminifera along the shelf of Tethys. 348 Evolution and Geological Significance of Larger Benthic Foraminifera (Amodiao, 2006). Fig. 5.38 summarizes the palaeoecological range and significance of most Cretaceous forms. 5.5 Palaeogeographic Distribution of the Cretaceous Larger Foraminifera The breakup of the super- continent Pangaea, which began during the Jurassic, con- tinued during the Early Cretaceous. Many of the land masses were covered by shallow continental oceans and inland seas. Europe, Asia, Africa and North America were all a series of islands. The Cretaceous saw the lengthening and widening of the proto- Atlantic Ocean, which began to spread further south, while the Alps began to form in Europe. India broke free of Gondwana and became an island continent. Africa and South America split apart, Africa moving north and closing the gap that was once the Tethys Sea. The continents began to take on their modern forms (see Fig. 5.39). Figure 5.39. Palaeogeographic and tectonic reconstruction of the Early Cretaceous (by R. Blakey, http:// jan.ucc.nau.edu/ ~rcb7/p aleogeographic.html). The Mesozoic Larger Benthic Foraminifera: The Cretaceous 349 Cretaceous Speciation 25 20 15 10 5 Total Tethyan 0 Cosmopolitan Caribbean Figure 5.40. The total number of new genera found in each Cretaceous stage. Cretaceous Genera 60 50 40 30 20 10 Total Tethyan 0 Cosmopolitan Caribbean Figure 5.41. The number of genera occurring in each Cretaceous stage. Berriasian Berriasian Valanginian Valanginian Hauterivian Hauterivian Barremian Barremian Aptian Aptian Albian Albian Cenomanian Cenomanian Turonian Turonian Coniacian Coniacian Santonian Santonian Campanian Campanian Maastrichtian Maastrichtian 350 Evolution and Geological Significance of Larger Benthic Foraminifera In the Berriasian, most of the characteristic larger benthic foraminifera of the Cretaceous had yet to evolve. With very few new appearances in the Berriasian (5%), nearly all of the larger benthic foraminifera were Jurassic survivors, and 55% of the Berriasian larger foraminifera were restricted to the Tethyan realm (Figs 5.40 and 5.41). They colonised, together with “aragonitic” scleractinian corals, calcareous green algae (e.g. Actinoporella podolica) and rudists bivalves, all early Cretaceous reefs. Most of the forms continued through to the Valanginian. Few foraminifera were to be found in the Caribbean (see Fig 5.41). The Pacific Plate was growing, but still quite small, and the entire western margin of North and South America was fringed with volcanic island arcs (Zharkov et al. 1998). Towards the late Berriasian only 5% of the larger foraminifera had disappeared from Tethys where the cooling trend of the late Jurassic was at an end. The Valanginian was a period of transition between the relatively cold time at the end of the Jurassic to the “greenhouse” world that continued for the rest of the Cretaceous. During this period, new forms evolved (14% of the whole Valanginian component, see Figs 5.40 and 5.41), increasing the dominance of the agglutinated fora- minifera in the Tethyan realm. Only a few of new forms (20%) were cosmopolitan (see Fig. 5.40 and Chart 5.1). The Hautervian saw the continuation of forms from the Valanginian, with very few new species appearing (18%), and again mainly in Tethys. Although the end of the Valanginian and early Hauterivian, saw considerable volcanic activity, such as the devel- opment of the Paraná traps in South America (Fig. 5.42), together with their smaller sev- ered counterpart of the Etendeka traps in Namibia and Angola (Courtillot and Renne, 2003), very few species of larger foraminifera became extinct (only ~7%), and those that did were mainly in Tethys (see Fig. 5.43). The extinct forms were mainly the strong, elon- gated forms with solid cores that had survived the Jurassic boundary. This is in contrast to the statement made by Courtillot and Renne (2003) that the end of the Valanginian is Figure 5.42. The end Barremian-e arly Aptian world, showing the position of the Ontong–J ava LIP and other middle Cretaceous events. The Mesozoic Larger Benthic Foraminifera: The Cretaceous 351 Cretaceous Extinctions 45 40 35 30 25 20 15 10 5 Total Tethyan 0 Cosmopolitan Caribbean Figure 5.43. The number of genera that went extinct at the top of each Cretaceous stage boundary. a prominent extinction level (although there may have been a major crisis for bryozoan faunas). This stage boundary may have also coincided with a cooling event, as noted by Walter (1989) that was followed by a warming event around (133– 132 Ma). At the beginning of the Barremian, foraminifera flourished and many new forms appeared (32%, see Fig. 5.40), however, they were short lived and did not have time to become established, as most of them became extinct towards the end of the Barremian. The end of the Barremian, a time of a major anoxic event and also of global sea- level rise (Courtillot and Renne, 2003), saw a modest increase in the number of extinctions (22% of the Barremian assemblages, see Fig. 5.43). The end Barremian-e arly Aptian extinctions could have been the results of, or trig- gered by, the Ontong–J ava volcanic event (see Fig. 5.42), which gave rise to the largest of all oceanic basaltic plateaus, that in places reaches a thickness of 40 km (Courtillot and Renne, 2003). Hallam and Wignall (1997) point out that most early Cretaceous biota were largely unaffected by this and other major volcanic episodes, and indeed as seen in Fig. 5.43 and Chart 5.1, there are a relatively modest number mass extinction of larger benthic foraminifera at that time. This may be because these eruptions were submarine and, as explained by Courtillot and Renne, the enormous mass of ocean water is expected to act as a strong buffer; SO2 is not expected to reach the atmosphere but CO2 would still lead to water anoxia and a greenhouse period (Kerr, 2006). For this reason, there are good reasons to correlate the Ontong–J ava volcanism and the Berriasian Valanginian Hauterivian Barremian Aptian Albian Cenomanian Turonian Coniacian Santonian Campanian Maastrichtian 352 Evolution and Geological Significance of Larger Benthic Foraminifera end- Barremian Ocean Anoxic Event (OAE), but perhaps because of their reef forming niche this led to the extinction of only a few larger foraminiferal genera. The Aptian was an eventful, long stage, with a rapid rate of oceanic spreading in the Atlantic. The Atlantic Ocean opened wide enough to allow significant mixing of waters across the equator that seemed to have been associated with a series of OAEs, which continued periodically for ~35 Ma (into the Santonian). Black shales, signalling deep ocean anoxia, have been found in both the Central Tethys (Europe) and the western Pacific provinces, while the Tethys Ocean was closing and the Alps began to form. Ecological evidence points towards an unusually high rate of volcanic activity, especially at mid-o ceanic ridges. The Kerguelen plateau (Fig. 5.42), the sec- ond largest oceanic plateau after the Ontong– Java plateau, erupted in the southern Indian Ocean in the early Aptian, around 118 Ma, which is also the time of eruption of the Rajmahal traps in eastern India. The Kerguelen plateau, Rajmahal traps, and eruptions in western Australia were located close to each other at the time of erup- tion, and were roughly coeval with the breakup of eastern Gondwana (Courtillot and Renne, 2003). During these eruptions the sea floors rose, forcing the sea water to rise worldwide, flooding up to 40 % of the continents. Sea levels were over 200m higher than present day levels. Such a high rate of volcanic activity released mas- sive amounts of carbon dioxide into the ocean and ultimately the atmosphere. This quantity of carbon dioxide would have made the oceans relatively short of oxygen. The evidence for this is displayed in the geological record. The abundance of black shale and petroleum-r ich formations, suggest they formed in an oxygen poor envi- ronment. Dramatic rises in temperature was recorded in the early Aptian and mid- Aptian (Jenkyns, 2003; Jenkyns and Wilson, 1999), and there is a significant turnover of larger benthic foraminifera throughout the Aptian in Tethys (see Fig. 5.43 and Chart 5.1). Eleven new agglutinated genera appeared in the Aptian, ten of them restricted to Tethys, while the other was cosmopolitan (Fig.  5.40 and Chart 5.1). Large miliolides made their first appearance in the Aptian and went on to colonise hot lagoonal environments in the Albian and Cenomanian of Tethys. Large conical, agglutinated, internally complicated orbitolinids invaded the shallow warm reefal environments of Tethys. Rudists were also very common during the Aptian replacing corals in many niches. The late Aptian saw an increase in extinctions, but some of these were replaced by new genera at the Aptian- Albian boundary. These extinctions coincided with a rapid sea-l evel fall at the Aptian-A lbian boundary and the collapse of reef ecosystems (Walliser, 1996). In the Albian, and for the first time in the Cretaceous, a small percentage of larger foraminifera become restricted to the Caribbean province (2%). With 50% belonging only to Tethys, the larger foraminifera set a provincialism trend that extends into the Late Cretaceous. Towards the end of the Albian, the Caribbean foraminifera became extinct and many of the Early Cretaceous Tethyan forms disappeared. This event may have been correlated to the Hess Rise volcanic event in the North Pacific Ocean (Eldholm and Coffin, 2000). Following the late Albian event, the early Cenomanian interval is marked by a short term turnover in larger foraminifera, 25 new genera appeared in Tethys, out of which 68% were alveolinids. It is the highest faunal turnover in the Cretaceous (see Figs 5.40, 5.41 and 5.43). The main breeding grounds of the larger foraminifera were The Mesozoic Larger Benthic Foraminifera: The Cretaceous 353 the tropical to subtropical lagoonal and shallow shelf waters of Tethys, with only one new genus appearing in the Caribbean. However, by the end of the Cenomanian many of them became extinct and indeed many of the early Cretaceous forms and Jurassic survivors failed to survive the end of the Cenomanian. These extinctions were initi- ated by the collapse of palaeotropical reef ecosystems, which happened according to Walliser (1996) near the middle Cenomanian. The extinction affected most aggluti- nated and porcelaneous foraminifera, leaving globally many empty niches. As with the epifaunal bivalves, which were virtually unaffected by these events (Harries and Little, 1999), some of the surviving taxa of larger foraminifera were also epifaunal and thus pre-a dapted to low-o xygen conditions, such as Ataxophragmium. Bambach (2006) considered this peak of extinction to be connected with an oceanic event, while Walliser (1996) associated these extinctions with a near-p eak Mesozoic eustatic sea- level high- stand (sea levels rose dramatically to over 200m higher than present levels in the Cenomanian and 300m above present stand in the early Turonian). These events might also have been affected by another major sub- marine event (the Wallaby erup- tion) in the Indian Ocean (Eldholm and Coffin, 2000), which again would have been associated with the high emission of CO2 contributing to the global warming peak and greenhouse climates during that period. In the Turonian, very few new larger foraminifera appeared. However, towards the end of the Turonian many of the foraminifera that survived the Albian-C enomanian crisis become extinct (see Fig. 5.43). Although none of the larger foraminifera were restricted to the Caribbean in the Turonian, a major extinction among Caribbean rudist occurred at this time (Walliser, 1996), and the restriction of reef ecosystems might have contributed to the disappearance of the global foraminifera. Johnson et al. (1996) proposed that the collapse of mid-C retaceous dominated reef ecosystems, may be attributed to the collapse of the Tethyan ocean- climate system. Courtillot and Renne (2003) stated that the volcanism, which occurred between 91 and 88 Ma in a number of short, discrete events, may be the cause of the oceanic events. The Caribbean- Colombian Cretaceous Igneous province (CCIP), the Madagascar event (see Fig. 5.44) and Phase 2 of the Ontong- Java event all occurred within this short interval, and would have certainly contributed to the extinction phase at the end of the Turonian. Kerr (2006) has proposed volcanism-r elated CO2 lead to an enhanced greenhouse climate and global warming, leading eventually to the extinction of the most vulnerable reefal communities of larger foraminifera and rudists. There is an oxygen isotope anomalies at 91.3 Ma (Bornemann et al., 2008), that has been inter- preted as being indicative of a glaciation event in the middle of the Turonian, one of the warmest periods of the Mesozoic. If this event occurred, it may have been triggered by short term atmospheric changes associated by flood basalt events in the Caribbean and Madagascar regions. At the Coniacian boundary, a short-t erm turnover of larger benthic foramin- ifera occurred. About a quarter of the Coniacian foraminifera were new, and only a few genera became extinct towards the end of this stage. The foraminifera turnover in the Coniacian takes place according to Walliser (1996) during a global flood- ing interval and the termination of the second regional oxygen depletion event, which is recognized by many places by organic- rich dark shales. The Santonian saw the appearance of a significant number of new genera, however, by the end of the 354 Evolution and Geological Significance of Larger Benthic Foraminifera Figure 5.44. The end Maastrichtian world showing the position of the Chicxulub impact and other major Late Cretaceous events. Santonian a high percentage of foraminifera again became extinct, mainly in the Tethyan realm. This may again have been triggered by renewal of the Kerguelen, Broken Ridge volcanic event that generated over 9  million km3 eruptive material (Eldholm and Coffin, 2000). During the mid to early Late Cretaceous agglutinated and porcelaneous fora- minifera were dominant in the shallow water of the Tethyan realm, however, in the Late Cretaceous, a major shift occurred when new groups developed and progressed to fill the empty niches left by the two successive extinction events at the end of the Cenomanian and end of the Turonian. The calcareous hyaline orbitoids increased in diversity and abundance in the Campanian and Maastrichtian, showing for the first time provincialism, until they outnumbered the agglutinated and porcelaneous groups. By this stage, larger benthic foraminifera were either cosmopolitan or restricted to two different provinces. This might be the result of the Late Cretaceous developing isolated land masses (see Fig. 5.45) because of the dramatic rise in sea level, which produced major inland seas. As a result, the Campanian saw a rapid development of new species, the second highest in the Cretaceous (see Fig. 5.40), many of which colonized deeper water than their ancestors, the orbitoids. About 50% of the Campanian genera were new. Of these new appearances, 20% belonged exclusively to the Caribbean and only 20% were cosmopolitan, while the rest thrived in Tethys. The new appearances in the Caribbean might have been the result of the filling of reefal niches previously occupied by rudists before they were destroyed in the mid- Cretaceous. Towards the end of the Campanian, 15% of the foraminifera became extinct. This might have been related by two volcanic events in the Atlantic Ocean, the Sierra Leone Rise and the Maud Rise The Mesozoic Larger Benthic Foraminifera: The Cretaceous 355 Figure 5.45. Palaeogeographic and tectonic reconstruction of the Late Cretaceous (by R. Blakey, http://j an. ucc.nau.edu/~ rcb7/ paleogeographic.html). (Eldholm and Coffin, 2000). This was accompanied by a regional oxygen reduction event (Walliser, 1996) occurring during the Campanian. The beginning of the Maastrichtian was a period of high turnover, and about a quarter of the foraminifera had their first appearance at the Campanian- Maastrichtian boundary. The larger foraminifera in the Maastrichtian were highly developed and complicated internally, in order to compensate for their large sizes. Tethys was again the main breeding ground for new forms. The larger benthic foraminifera of the Cretaceous were brought to an end by one of the greatest mass extinctions of all time, the Cretaceous - Paleogene (K- P) event, or terminal Mesozoic extinction (Macleod et al., 1997). About 83% of the Maastrichtian foraminifera died out, including all orbitoids and alveolinids. The survivors were the smallest and toughest, 8% of the Tethyan forms and the rest being cosmopolitan. The K- P extinction has been the subject of many studies. Some have argued that it really began before the end of the Cretaceous, between 67.5 and 68 Ma, with the abrupt 356 Evolution and Geological Significance of Larger Benthic Foraminifera extinction of rudist bivalve dominated reef ecosystems (Johnson and Kauffman, 1990), with new radiometric dates putting this extinction around 68-6 8.5 (Walliser, 2003). Others blame the K- P extinction on a single or multiple impacts at ~66 Ma (Keller et al., 2003, Stüben et al., 2005, Schulte et al., 2006, Alegret et al., 2005, Kuroda et al., 2007), or on an associated massive volcanism, which in turn triggered significant climatic changes, inducing biotic crises and oceanic anoxia. There is certainly evidence support- ing multiple impact events at the end Maastrichtian. An impact crater of 24km diam- eter occurred in Ukraine around 65.17 Ma, the Boltysh impact (Kelley and Gurov, 2002; see Fig 5.45). On the other hand, the largest impact, that is now widely accepted to have been the cause of the mass extinctions, occurred at Chicxulub, Yucatan in Mexico (Glikson, 2005; Macleod, 2013; see Fig 5.44). However, at about the same time as these impact events there was also a major volcanic event that formed the Deccan Traps (see Fig 5.45). Courtillot and Renne (2003) pointed to the Ir bearing layer related to the Chixculub crater impact within the traps, which indicate that Deccan volcan- ism began prior to this impact and straddled it in time. There was also a sea level fall ~100kyr before the Ir- defined K- P boundary. The low point occurred 10kyr before the K- P event according to Hallam and Wignall, who attribute the subsequent rise and warming to Deccan eruptions and associated CO2 release. The K-P boundary impact coincided with the sea level regression and the Chicxulub event contributing to the demise of all tropical and subtropical larger foraminifera, to say nothing of the dinosaurs and most other life forms! 1 2 3 4 5 6 7 8 9 10 13 11 12 14 Plate 5.1 Scale bars: Figs 1, 3, 10-1 4 = 1mm; Fig. 2, 4- 8 = 0.5mm; Fig. 9 = 0.15mm. Figs 1- 4. Pseudedomia complanata Eames and Smout, Campanian, Um Gudair, Kuwait, 1) holotype, NHM P42909; 2- 4) paratypes, NHM P42916-9 . Figs 5- 6. Pseudedomia globularis Smout, Campanian, Iraq, paratypes, NHM P42646. Fig. 7 Meandropsina vidali Schlumberger, Santonian, Spain, B-F orm, NHM P35882. Fig. 8. Meandrospina vidali Schlumberger biconcava Henson, oblique section, Cenomanian, Wadi Meliha, Israel, paratype NHM P35886. Fig.  9. Multispirina sp., Cenomanian, Qatar, NHM M2425. Figs 10-1 1. Nezzazatinella picardi (Henson), paratype, Santonian, Egypt, 10) NHM P39125; 11) NHM P39100,. Fig. 12. Nezzazata conica (Smout), paratype, Cenomanian, Iraq, NHM M8311. Fig.  13. Rotalia sp., Maastrichtian, Qatar, NHM M3687. Fig. 14. Mangashtia viennoti Henson, syntype, Cenomanian toTuronian, Tang- i- Kurd, Iran, NHM P35881. Plate 5.2 Scale bars: Figs 1- 7 = 0.5mm; Figs 8- 11 = 0.25mm; Figs 12- 14 = 1mm. Fig. 1. Montsechiana aff. montsechiensis Aubert et  al., Valanginian, Murban-2 , Thamama V, Abu Dhabi, NHM coll. Fig.  2. Hottingerita complanata (Hottinger), figured by Hottinger (1967), Barremian, Switzerland. Fig.  3. Paracoskinolina sp., Albian, Libya, UCL coll. Fig.  4. Neoiraqia convexa Danilova, figured by Loeblich and Tappan (1988), Cenomanian to Turonian, Yugoslavia. Fig.  5. Gyroconulina columellifera Schroeder and Darmoian, Maastrichtian, Saudi Arabia, NHM M7434. Figs 6-7 . Coxites zubairensis Smout, para- type, Cenomanian, Iraq, 6) NHM P42954; 7) NHM P42952. Figs 8-9 . Elphidiella multiscissurata Smout, Maastrichtian, Qatar, 8) NHM P42175; 9) NHM P4217577. Figs 10- 11. Fissoelphidium operculiferum Smout, paratypes, Maastrichtian, Qatar, 10)  NHM P42167; 11)  NHM 42168. Figs 12- 13. Lepidorbitoides minor (Schlumberger), topotypes, Maastrichtian, Netherlands, MHM M3163- 4. Fig.  14. Lepidorbitoides sp., Maastrichtian, Kuh - i Bizadan, near Jahrun, Iran, NHM P32949. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 1 2 3 4 5 6 7 8 B 10 9 11 12 13 14 15 16 Plate 5.3 Scale bars: Figs 1- 4, 8-1 6 = 0.5mm; Figs 5- 7 = 0.16mm Figs 1- 2. Banatia aninensis Schlagintweit and Bucur, types figured by Schlagintweit and Bucur (2017) from the upper Barremian of Reşita, Moldova Nouă Zone, SW Romania, (Courtesy of Dr Schlagintweit). Fig. 3. Tarburina zagrosiana Schlagintweit et al., holotype figured by Schlagintweit et al. (2016) from the late Maastrichtian of the Tarbur Fm., Zagros Zone, SW Iran (Courtesy of Schlagintweit). Fig. 4. Spirocyclina sp. Middle Cretaceous, Shilaif (Khatiyah) Fm., Qatar, NHM coll. Fig.  5. Demirina meridionalis Özcan, Cenomanian, southeastern Turkey (courtesy of Dr Özcan). Fig. 6. Lupertosinnia pallinii Farinacci, type specimen from Farinacci (1996), upper Campanian limestones, southern Italy. Fig. 7. Takkeina anatoliensis Farinacci and Yeniay, type specimen from Farinacci and Yeniay (1994), western Taurus, Turkey. Fig. 8. Feurtillia frequens Maync, topotype from Switzerland figured by Hottinger (1967). The thick septa and strong apertural neck distinguish Feurtillia even in small wholly coiled specimens. Fig. 9. Zagrosella rigaudii Schlagintweit and Rashidi, type figured by Schlagintweit and Rashidi (2017) from upper Maastrichtian, Tarbur Fm. of the Naghan section, Zagros Zone, SW Iran. Fig. 10. Everticyclammina eccentrica (Redmond), Valanginian, Mutriba-2 , core, 11,690 ft, MINAGISH D, Yamama, Arabian Peninsula, NHM coll. Fig. 11. Daxia cenomana Cuvillier and Szakall, figured by Maync (1972) from late Cenomanian, France. Fig.  12. Chubbina cardenasensis (Barker and Grimsdale), figured by Robinson (1968) as a cotype megalospheric specimen of Chubbina jamaicensis Robinson, Campanian, Jamaica, BMNH P48053. Fig. 13. Pseudochoffatella cuvillieri Deloffre, figures by Gušić (1975) from the lat- est Aptian- earliest Albian of North Croatia. Fig. 14. Spiroconulus perconigi Allemann and Schroeder, holo- type figured by Allemann and Schroeder (1972) the Bathonian of Cadiz, South Spain. Fig. 15. Textulariida, miliolids and Dasyclad sp., Aptian, Tibet, UCL coll. Fig. 16. Bacinella irregularis Radoičič, Aptian, West Africa, UCL coll. Plate 5.4 Scale bars: Figs 1 – 4, 6- 11 = 0.5mm; Fig. 5 = 0.25mm. Figs 1- 2. Praereticulinella cuvilleri Deloffre and Hamaoui, figured type by Deloffre and Hamaoui (1970), Barremian, Spain, 1)  paratype, axial sec- tion; 2)  holotype, equatorial section. Fig.  3. Reticulinella reicheli (Cuvillier, et  al.), metatypic topotypes, Cenomanian to Turonian, Libya, UCL coll. C2.6412 Fig. 4. Martiguesia cyclamminiformis Maync, topotype, Santonian, Les Martigues, equatorial section showing the progressive infilling of the chambers with exten- sions of the alveolar hypodermal, signalling the evolution of Loftusia (see 26), NHM coll. Figs 5-6 . Qataria dukhani Henson, paratypes, 5) NHM P35978; 6) NHM P35984. Fig. 7. Daxia sp., This form has thicker test and higher chambers, than D. cenomana (Cenomanian) and D. minima (Aptian), Aptian, Zakum -1 . Upper Kharaib (upper Thamama II), Arabian Peninsula, NHM coll. Fig. 9. Gendrotella rugoretis Gendrot, topo- type figured by Banner (1966), Santonian, Les Martigues, NHM coll. Fig. 10. Stomatostoecha plummerae Maync, paratype figured by Maync (1972), Albian, Texas. Fig. 11. Murciella cuvillieri Fourcade, figured by Fourcade (1966), Campanian, Spain. 1 2 3 4 5 6 7 8 9 10 11 1 2 3 4 6 5 7 8 9 10 11 12 13 14 15 16 Plate 5.5 Scale bars: Figs 1-7 , 12-1 4 = 0.25mm; Figs 8-1 1, 15, 16 = 0.5mm Figs 1-7 . Archaecyclus midorien- talis Eames and Smout, Campanian, Um Gudair, Kuwait, paratypes, NHM P42924-8 . Figs 8-9 . Choffatella decipiens Schlumberger, Early Cretaceous, Dukhan no.1 well, Qatar, NHM P35780. Fig. 10. Pseudochoffatella cuvillieri Deloffre, A- form, figured by Cherchi and Schroeder (1982), Aptian, Spain. Fig. 11. Sornayina fois- sacensis Maries, figured by Loeblich and Tappan (1988), Coniacian, France. Fig.  12. Charentia cuvillieri Neumann, figured by Hofker (1965), Aptian to Albian, Spain. Fig. 13. Debarina hahounerensis Fourcade, paratype, figured by Fourcade et al. (1972), Aptian, Algeria. Fig. 14. Lituola sp. aff. nautiloides (Lamarck), Late Cretaceous, Iraq, NHM P35873. Fig. 15. Lituonelloides compressus Henson, paratype, Maastrichtian, Dukhan no. 1, Qatar, NHM P35876. Fig. 16. Pseudocyclammina bukowiensis (Cushman and Glaz.), Upper Zangura Limestone, Valanginian, Iraq, NHM P52301. Plate 5.6 Scale bars: Figs 1- 5, 7, 10- 11 = 1mm; Figs 6, 8-9 , 12- 19 = 0.5mm. Fig. 1. Pseudocyclammina bukowiensis (Cushman and Glaz.), the thick- walled, coarsely agglutinated, inner hypodermal alveolae dis- tinguish this species from P. lituus, Upper Zangura Limestone, Valanginian, Iraq, NHM P52301. Figs 2- 5. Pseudocyclammina lituus (Yokoyama), Berriasian to Hauterivian, Iran, NHM P52300; 3- 5) Valanginian, Libya, UCL coll. Fig. 6. Pseudocyclammina cylindrica Redmond, Berriasian to Valanginian, United Arab Emirates, NHM P52304. Fig.  7. Pseudocyclammina rugosa (d’Orbigny), topotype, Cenomanian, Île de Madame, Charente Inférieur, France, NHM coll. Fig.  8. Cylammina sp., equatorial section, Cretaceous, UCL coll. Fig. 9. Lituola obscura Barnard and Banner, paratype, Late Cretaceous, Labiatus Zone, Norfolk, England, 4507, UCL coll. Figs 10-1 1. Chubbina jamaicensis Robinson, lectotypes, figured by Robinson (1968), Campanian, Jamaica. Figs 12-1 3. Hemicyclammina sigali Maync, Albian, Rumaila 1, Asara Formation, Iraq, NHM M8348. Figs 14 - 15. Hemicyclammina whitei (Henson), Aptian toAlbian, Dukhan no. 3, Qatar, NHM P35797-9 9, 14) holotype; 15) paratype. Fig. 16. Buccicrenata hedbergi (Maync), paratype figured by Maync (1953), middle Albian, Venezuela. Fig. 17. Pseudolituonella reicheli Marie, figured by Sartorio and Venturini (1988), late Cenomanian, Iran. Fig. 18. Ammobaculites gr. edgelli Gollestaneh, late Valanginian, Fuwairat-1 , core, 4150-5 5 ft, Yamama, Arabian Peninsula. Fig. 19. Spirocyclina choffati Munier- Chalmas, near-t opotype, showing thick early epidermis, but rapid development of subepidermal (hypodermal) alveolae and sub- hypodermal transverse septulae in second whorl, figured by Banner (1966), from the Santonian, Carrière de Martigues, Bouches- du- Rhône, France. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 2 1 3 5 4 6 7 8 9 10 11 12 13 14 Plate 5.7 Scale bars: Figs 1-8 , 11- 14 = 0.5mm; Figs 9- 10 = 0.15mm. Figs 1- 3. Dictyoconella minima Henson, paratypes, Cretaceous, Dukhan no.3 well, Qatar, 1)  NHM P35836; 2)  NHM P35835; 3)  solid specimen, NHM P35838. Figs 4- 6. Dictyoconella complanata Henson, Maastrichtian, Dukhan no. 1 well, Qatar, 4) hol- otype, NHM P36832; 5-6 ) paratypes, 5) NHM P35831; 6) NHM P35830. Fig. 7. Cribellopsis neoelongata (Cherchi and Schroeder), holotype, figured by Loeblich and Tappan (1988), early Aptian, France. Fig. 8, 14. Pseudorbitolina marthae Douvillé, Maastrichtian, Dukhan no.1 well, Qatar, 8) NHM P35964; 14) NHM P35965. Fig. 9. Abrardia mosae (Hofker), figured by Loeblich and Tappan (1988), Maastrichtian, France. Fig. 10. Campanellula capuensis De Castro, holotype figured by Loeblich and Tappan (1988), Valanginian to Barremian, Italy. Fig. 11. Calveziconus lecalvezae Caus and Carnella, figured by Loeblich and Tappan (1988), Campanian, Spain. Fig.  12. Balkhania balkhanica Mamontova, type figures from Mamontova (1966), Barremian, Great Balkhan Mountain, Turkmenistan. Fig. 13. Broeckinella arabica Henson, holo- type, Maastrichtian, Dukhan no. 1 well, Qatar, NHM P35778. Plate 5.8 Scale bars:  Figs 1- 8  =  1mm. Fig.  1. Loftusia coxi Henson, syntype, Maastrichtian, Dukhan no.1cwell, Qatar, NHM P35378. Fig. 2. Loftusia persica Brady, Maastrichtian, Iran, NHM P34679. Fig. 3. Loftusia harrison Cox, Maastrichtian, Adiyaman, Turkey, UCL coll. Figs 4-8 . Loftusia sp., Maastrichtian, 4, 6- 8) UCL coll.; 5) Iraq, NHM 7467. 1 2 3 4 6 5 7 8 3 1 2 4 5 A 6 7 8 9 10 Plate 5.9 Scale bars: Figs 1- 2, 4- 6, 8- 10 = 0.5mm; Fig. 3 = 1mm.; Fig. 7 = 0.25mm. Fig. 1. Chrysalidina sp., Cenomanian, Iran, NHM coll. Fig. 2- 4. Montseciella arabica (Henson), Barremian, Dukhan no.2 well, Qatar, 2)  NHM P35815; 3)  NHM 35753; 4)  NHM 35760. Fig.  5. Mangashtia viennoti Henson, syntype, Cenomanian to Turonian, Tang- i- Kurd, Iran, NHM P35881. Figs 6- 7. Loftusia persica Brady, Maastrictian/ ?Eocene, Iran, NHM coll., 6) enlargement of centre of test; 7) enlargement of one chamber to show A) Turborotalia sp. (an Eocene planktonic foraminifera) trapped in the test. Figs 8-1 0. Chubbina cardenasensis (Barker & Grimsdale), Late Cretaceous, Cardenas, San Luis Potosi, Mexico, NHM P33042- 4. 1 2 3 4 6 75 8 9 10 11 12 13 14 A B 15 17 16 18 19 20 Plate 5.10 Scale bars: Figs 1, 3, 5, 6, 8, 9- 11, 20 = 0.5mm; Figs 2, 4, 7, 12- 19 = 0.25. Figs 1- 2. Opertum sp., Aptian, Zakum- 1, 7116ft, Thamama II, Abu Dhabi, NHM coll. Fig. 3. Voloshinoides sp., Albian, England, NHM coll. Fig. 4. Voloshinoides sp., Albian, Iran, NHM coll. Figs 5-8 . Cuneolina parva Henson, paratypes, Santonian, Sudr Heitan, Egypt, 5) solid specimen, NHM P39116; 6) vertical section, NHM P39119; 7) hor- izontal section, NHM P39121; 8) vertical section, NHM P39120. Fig. 9. Cuneolina cylindrica Henson, holo- type, Maastrichtian, North Iraq, NHM P39122. Fig. 10. Cuneolina sp., Late Cretaceous, Spain, NHM coll. Fig. 11. Cuneolina hensoni Dalbiez, syntypes, vertical and horizontal sections, figured by Dabiez (1958), from the Valanginian, France. Fig. 12. Vercorsella camposaurii Sartoni and Crescenti, holotype figured by Sartoni and Crescenti (1962), Valanginian to Aptian, Central Apennines, Italy. Figs 13, 15. Thin section photomicro- graphs from Cenomanian, Audignon, near Saint- Sever, France, UCL collection, 13) horizontal section of Cuneolina pavonia d’Orbigny; 15) A) Pseudonummuloculina sp., B) fragment of Cuneolina pavonia d’Orbigny. Fig. 14. Vercorsella arenata Arnaud- Vanneau, Aptian, Tibet, UCL coll. Fig. 16. Pseudotextulariella cour- tionensis Brönnimann, holotype figured in Brönnimann (1966) from the Barremian of Switzerland. Fig. 17. Pseudotextulariella cretosa (Cushman), near-t opotype, figured by Brönnimann (1966), Cenomanian, England. Figs 18-1 9. Sabaudia capitata Arnaud-V anneau, figured by Chiocchini et al. (1984) from Aptian, Italy. Fig. 20. A packstone of Pseudocyclammina lituus (Yokoyama), Early Cretaceous, Makhul no.1 well Iraq, NHM P35970. Plate 5.11 Scale bars:  Figs 1-5   =  1mm; Fig.  6  =  0.25mm. Figs 1-3 . Dicyclina qatarensis Henson, syn- types, Cenomanian, Dukhan well, Qatar, NHM P39113- 5. Figs 4- 6. Orbitolina concava (Lamarck), 4- 5) Cenomanian, Spain, UCL coll; 6) late Albian, Upper Greensand, Devon, Elliot coll., NHM P43429. 1 3 2 4 6 5 1 2 3 4 5 6 7 8 10 11 9 Plate 5.12 Scale bars: Figs 1-1 1 = 0.5mm. Fig. 1. Orbitolina sefini Henson, Cenomanian, NE of Erbil, Iraq, NHM P35903. Fig. 2. Mesorbitolina texana (Roemer), Albian, south of Husainabad, Iran, NHM M/ 2050. Fig. 3. Rabanitina basraensis Smout, Cenomanian, Zubair 1, 8600- 50ft, Iraq, NHM P42961. Fig. 4. Orbitolina cf. duranddelgai Schroeder, Cenomanian, Iran, NHM M2058. Fig.  5. Conicorbitolina cobarica (Schroeder), Cenomanian, Iran, NHM M2074. Fig. 6. Orbitolina sp., SEM photograph of a solid specimen, UCL coll. Fig. 7. Orbitolina qatarica Henson, solid specimen, early Cenomanian, Dukhan no.1 well, Qatar, NHM P35916. Figs 8-9 . Dicyclina qatarensis Henson, syntypes, early Cenomanian, Dukhan well, Qatar, NHM P39112, M2512. Fig.  10. Coskinolinoides texanus Keijzer, type species from Loeblich and Tappan (1988), Albian, USA. Fig.  11. Paracoskinolina sunnilandensis (Maync), paratype, from Maync (1955), Aptian- Albian, USA. Plate 5.13 Scale bars: Figs 1-1 4 = 0.25mm; Figs 15-1 6 = 0.5mm. Fig. 1 Simplorbitolina manasi Ciry and Rat, type figure from Loeblich and Tappan (1988), Early Cretaceous, Spain. Figs 2-3 . Iraqia simplex (Henson), syntype, Aptian, Iraq, NHM P35871. Fig.  4. Orbitolinopsis kiliani (Prever), type figure, Aptian, France. Figs 5- 6. Neorbitolinopsis conulus (Douvillé), type species figured from Albian to early Cenomanian, Spain, 5) vertical section; 6) transverse thin section showing the second-o rder peripheral rectangular chamberlets. Figs 7- 10. Montseciella arabica (Henson), Barremian, Dukhan no.2 well, Qatar, 7- 8) NHM P35810; 9- 10) NHM P35805. Figs 11- 12. Dictyoconus walnutensis (Carsey), middle to early Albian, Walnut Clay, U.S.A., NHM coll. Fig.  13. Paleodictyoconus cuvillieri Foury, latest Aptian- earliest Albian, top Shuaiba, Saudi Arabia, NHM coll. Fig. 14. Orbitolinella depressa Henson, Cenomanian- Turonian, syntype, Dukhan no.2 well, Qatar, NHM coll. Figs 15- 16. Pseudocyclammina lituus (Yokoyama), Early Cretaceous, Makhul no.1 well, Iraq, NHM P35970. 2 1 3 4 5 6 9 7 8 10 11 12 13 14 15 16 Plate 5.14 Scale bars: Fig. 1 = 1mm; Figs 2- 16 = 0.5mm. Figs 1-4 . Palorbitolina lenticularis (Blumenbach, 1805), identified as Palorbitolina discoidea Gras, Barremian to Aptian, Thamama Fm., Oman, NHM coll., 1)  transverse section; 2) enlargement of the Fig. 1; 3) axial section; 4) oblique transverse section. Fig. 5. Orbitolina trochus (Fritsch), late Albian, Pilatus Kulu, Lucerne, Switzerland, NHM P38097. Figs 6- 9. Mesorbitolina delicata Henson, Hadhramaut, 6-7 ) ideotype, early Aptian, NHM P35918; 8) Qatar, NHM P35919; 9) Qatar, NHM coll. Figs 10-1 1. Mesorbitolina libanica Henson (= Mesorbitolina texana (Roemer)), 10) syntype, Aptian, Mdeireidj, Lebanon, NHM P35930; 11) Aptian, Medjel Chems, Syria, NHM M/ 2160. Fig. 12. Mesorbitolina kurdica Henson, syntype, late Aptian, Aoraman, Iraq, NHM P35935. Figs 13, 16. Palorbitolina cf. Lenticularis (Blum), Aptian, east of Beirut, Lebanon, NHM P25946; 16) Aptian, North Lattakia, Syria, NHM P35944. Figs 14- 15. Orbitolina qatarica Henson, early Cenomanian, Dukhan no.1 well, Qatar, NHM Henson coll. Plate 5.15 Scale bars:  Figs 1 - 5, 21  =  0.25mm. Figs 6- 20  =  0.5mm. Fig.  1. Neotrocholina lenticularis (Henson), paratype, Cenomanian, Dukhan no. 1 well, Qatar, NHM P38489. Figs 2- 4. Neotrocholina minima (Henson), paratype, late Cenomanian, Dukhan no. 1 well, Qatar, 2) NHM M4674; 3- 4) solid specimen, para- type, NHM P38550. Fig. 5. Trocholina altispira Henson, paratype, Cenomanian, Dukan no. 3 well, Qatar, NHM P38486. Figs 6- 8, 16. Chrysalidina gradata d’Orbigny, 6) Cenomanian, Deh Luran, Iran, NHM P39127; 7) Cenomanian, Jebel Madamar, Natih Fm., Oman, NHM P52608; 8) Cenomanian, Auvignon, France, UCL coll; 16) Cenomanian, Île de Madame, Charente Inférieur, France, NHM P5203. Fig. 9. Paravalvulina arab- ica (Henson), Valanginian, Dukhan no. 2 well, Qatar, NHM P52645. Fig. 10. Morphologically intermediate between Praechrysalidina infracretacea Luperto Sinni and Dukhania conica Henson, figured by Banner et al. (1991), Albian, Northern Iraq, NHM P52591. Fig. 11. Pseudomarssonella cf. plicata Redmond, Bajocian or Bathonian, Abu Dhabi, Um Shaif, NHM P52629. Figs 12-1 4. Praechrysalidina infracretacea Luperto Sinni, 12)  Aptian, Um Shaif, Upper Shuaiba Fm., United Arab Emirates, NHM P52584; 13-1 4) Hauterivian, Dukhan no. 2 well Qatar, NHM P52580. Fig. 15. Paravalvulina arabica (Henson), Valanginian, Zakum Fm., Well-J umayla- 1, United Arab Emirates, NHM P52641. Figs 17-1 8. Morphologically intermediate between Praechrysalidina infracretacea Luperto Sinni and Chrysalidina gradata d’Orbigny, 17) Cenomanian, WM 99, Natih Fm. Oman, NHM coll; 18) middle Cretaceous, Iran, NHM P52595. Fig. 19- 20. Dukhania conica Henson, paratypes, Cenomanian, Dukhan no. 3 well, Qatar, 19) NHM P52601; 20) NHM P52597. Fig. 21. Accordiella conica Farinacci, figured by Sartorio and Venturini (1988), Coniacian to Santonian, Italy. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 A B 17 18 19 20 21 1 2 2 3 4 6 5 9 8 7 10 11 Plate 5.16 Scale bars:  Figs 1- 11  =  0.5mm. Figs 1- 3. Dohaia planata Henson, paratypes, Cretaceous, Dukhan no.1, well Qatar, NHM P35843. Fig. 4. Eclusia moutyi Septfontaine, figured by Azema et al. (1977), Valanginian, France. Figs 5-8 . Edomia reicheli Henson, Cenomanian, Wadi Meliha, Israel, 5)  paratype, NHM P35850; 6) holotype; 7) paratype, megalospheric form (A- form), NHM P35851; 8) paratype, NHM P35853. Fig.  9. Naupliella insolita Decrouez and Moullade, figured by Loeblich and Tappan (1988), late Albian, Greece. Figs 10- 11. Orbitoides sp., Sabkha, Arab Fm., Abu Dhabi, UAE. 1 2 5 4 3 7 6 8 9 Plate 5.17 Scale bars: Figs 1- 4, 6- 7, 9 = 0.5mm; Figs 5, 8 = 1mm. Figs 1- 4. Monolepidorbis douvillei (Silvestri), Late Cretaceous, 1- 3) Jawan no.2, Iraq, NHM P40685- 6; 4) Qatar, NHM M8239. Figs 5, 6. Orbitoides sp., 5) Sabkha, Arab Fm. Abu Dhabi, UAE; 8) solid specimen, Gensa, France. Figs 7-8 . Orbitoides cf. tissoti Schlumberger, shales below quartz sandstone, Punjab, NHM coll. Fig.  9. Orbitoides faujasii (Defrance), Maastrichtian, Iraq, NHM M568. 1 2 3 4 6 5 7 8 9 Plate. 5.18 Scale bars: Figs 1-7 , 9 = 1mm; Fig. 8 = 0.5mm. Fig. 1. Orbitoides medius (d’Archiac), Turkey, UCL coll. Figs 2- 3. Orbitoides browni (Ellis), Late Cretaceous, Havana Province, Cuba, NHM P33354. Fig.  4. Orbitoides faujasii (Defrance), Maastrichtian, NHM coll. Fig.  5. Orbitoides faujasii (Defrance), Maastrichtian, Iraq, NHM M7305. Fig.  6. Orbitoides vredenburgi Douvillé, Maastrichtian, Iraq, NHM, M7302. Figs 7-8 . Praetaberina bingistani (Henson), syntypes, registered as Taberina bingistani Henson, mid- dle Cretaceous, Kuh-i - Bingistan, Iran, NHM P35786. Fig. 9. Sirelella safranboluensis Özgen- Erdem, holo- type figured by Özgen-E rdem (2002) from Turkey. 1 2 4 3 5 6 Plate 5.19 Scale bars: Figs 1, 3, 5-7  = 1mm; Fig. 2, 8-1 3 = 0.5mm; Fig. 4 = 0.15mm. Figs 1-6 . Simplorbites gensacicus (Leymerie), Late Cretaceous, Dukhan no. 51 well, Qatar, NHM coll. 1 2 12 3 4 5 6 7 8 9 10 11 13 Plate 5.20 Figs 1- 4. Omphalocyclus macroporus (Lamarck), Maastrichtian, Iran, NHM P3467-8 . Fig. 5. Omphalocyclus sp., solid specimen, Maastrichtian, France, UCL coll. Fig. 6. Laffitteina vanbelleni Grimsdale, Musharah no. 1 well, Iraq, NHM P40693. Fig. 7. Vaughanina sp., Campanian, Cuba, UCL coll. Figs 8- 10. Siderolites calcitrapoides Lamarck, Maastrichtian, Belgium, 8) NHM P54874; 9) P41580; 10) P41575. Fig.  11- 12. Siderolites sp., Maastrichtian, Holland, UCL coll. Fig.  13. Siderolites spengleri (Grimsdale), Maastrichtian, St Pietersberg Maastricht, Belgium, NHM P278. 1 2 3 4 5 7 6 8 9 Plate 5.21 Scale bars: Figs 1- 6, 9 = 1mm; Fig. 7 = 0.5mm; Fig. 8 = 0.25 mm. Fig. 1. Siderolites calci- trapoides Lamarck, Maastrichtian, Maastricht Valley, Holland, UCL coll. Figs 2-4 , 7- 8. Omphalocyclus sp., Maastrichtian, Libya, UCL coll. Fig. 5. Simplalveolina sp., Cenomanian, France, UCL coll. Fig. 6. Thin section of Orbitoides sp. and Siderolites sp, Maastrichtian, Spain, UCL coll. Fig.  9. Praealveolina spp., Cenomanian, France, UCL coll. 1 2 5 3 4 7 8 6 B 10 A 9 11 12 13 14 Plate 5.22 Scale bars: Figs 1-2 , 6, 11 = 0.5mm; Figs 3- 5, 7-1 0, 12- 14 = 1mm. Fig. 1. Zekritia langhami Henson, holotype, Turonian, Dukhan no. 3 well, Qatar, NHM P36030. Fig. 2. Archaealveolina reicheli (de Castro), figured by Loeblich and Tappan (1988), Aptian, Italy. Figs 3-4 . Cisalveolina sp., Cenomanian, India, UCL coll. Figs 5, 13. Ovalveolina sp., Cenomanian, France, 5)  solid specimens; 13)  thin section. Fig.  6. Lacazina sp., Coniacian, India, UCL coll. Fig. 7. Subalveolina sp., Santonian, France, UCL coll. Fig. 8. Simplalveolina sp., Cenomanian, France, UCL coll. Fig. 9. A) Ovalveolina sp., B) Praealveolina sp., oblique sections, early Cenomanian, Mexico, UCL coll. Figs 10- 11. Praealveolina tenuis Reichel, Cenomanian, Qatar, NHM M4804; 11)  enlargement of Fig.  12. Pseudedomia (= Sellialveolina sp.) registered at the NHM as Praealveolina cretacea (d’Archiac), Cenomanian, Qatar, NHM M7092. Figs 14. Cisalveolina lehneri Reichel, middle Cretaceous, Shilaif (Khatiyah) Fm, Qatar, UCL coll. 387 Chapter 6 The Cenozoic Larger Benthic Foraminifera: The Paleogene 6.1 Introduction As noted in the previous chapter, the Cretaceous- Paleogene crisis wiped out over 80% of the Maastrichtian larger benthic foraminifera. The Early Paleocene was a recovery period for larger foraminifera. As was the case during the recovery stage after previous extinctions, larger foraminifera were morphologically small and rare, and even the newly evolved foraminifera exhibited a morphological manifestation of post- crisis ecological stress, (i.e. the “Lilliput effect”, which characterises a temporary, within-l ineage size decrease after extinction events (Twitchett, 2006)). Their evolution and development occurred at different rates and followed different lines in different parts of the globe. It is now possible to recognise larger benthic foraminiferal bioprovinces in Tethys, the Americas, the Far East, in and Southern Africa (see for example BouDagher- Fadel and Price, 2010a,b,c, 2013, 2014, 2017). In the Tethys, for example, morphologically larger miliolides and rotaliides (espe- cially nummulitids and orthophragminids) did not appear before the Late Paleocene. Here, the miliolides included large fusiform alveolinids, which show examples of con- vergence with the extinct Cretaceous alveolinids, and the discoid soritids that become prominent throughout the Eocene. Some rotaliides developed a complex system of marginal cords, characteristic of Nummulites. These latter exhibited high rates evo- lutionary diversification and became very abundant in the Eocene, thriving (together with the three layered orthophragminid, Discocyclina and its descendants) in the forer- eef and shallow marine open platforms of Tethys. Parallel to these lines of evolution, the agglutinated textulariides developed into forms that imitated their Cretaceous ancestors by developing, in the case of the textulariides, internal pillars, and with the orbitolinids, complicated partitions. In the American province, however, the recovery period was longer than in Tethys, and it was not before the Middle Eocene that the rotaliides developed a new evolu- tionary lineage; they evolved into the three layered Lepidocyclina and Eulepidina. These American forms migrated into Tethys in the Oligocene, however a reverse migration of alveolinids and discocylinids from Tethys to the American province never occurred. The nummulitids of the American province never reached the giant sizes of their analogues in the Tethyan realm (see BouDagher-F adel and Price, 2010a, 2013, 2014, 2017). Paleogene larger foraminifera from different localities around the world have been studied in detail by many workers. Among the most recent studies are those of Hottinger and Drobne (1980), Adams (1987), Racey (1995), Banerjee et  al (2000), Hottinger (2001, 2013), Özgen-E rdem (2002), Meinhold, and BouDagher- Fadel et al. (2010a,b,c), Benedetti and Pignatti (2012), Politt et al. (2012), Hu et al. (2013), BouDagher-F adel 388 Evolution and Geological Significance of Larger Benthic Foraminifera and Price (2013, 2014, 2017), Scotchman et al. (2014). In this chapter, the taxonomy of the major larger Paleogene benthic foraminifera is summarized, and then their bio- stratigraphic, paleoenvironmental and paleogeographic significance are discussed. 6.2 Morphology and Taxonomy of Paleogene Larger Benthic Foraminifera The three orders of larger benthic foraminifera that dominated the Paleogene were the: • Textulariida • Miliolida • Rotaliida The relationship and evolution of the superfamilies of these orders is shown in Fig. 6.1. ORDER TEXTULARIIDA Delage and Hérouard, 1896 The tests of these agglutinated foraminifera are made of foreign particles bound by organic cement. They range from Early Cambrian to Holocene. Age Epoch Ma Paleogene larger benthic foraminifera 23.0 TEXTULARIIDA 33.9 MILIOLIDA 56.0 66.0 ROTALIIDA Figure  6.1. The evolution of the Paleogene orders (thick lines) and superfamilies (thin lines) of larger foraminifera. Late Cretaceous Paleocene Eocene Oligocene Miocene TEXTULARIOIDEA SPIROPLECTAMMINOIDEA COSCINOPHRAGMATOIDEA LITUOLOIDEA ATAXOPHRAGMIOIDEA ORBITOLINOIDEA COSKINOLINOIDEA ASTERIGERINOIDEA NUMMULITOIDEA ROTALIOIDEA ACERVULINOIDEA PLANORBULINOIDEA ORBITOIDOIDEA SORITOIDEA MILIOLOIDEA ALVEOLINOIDEA The Cenozoic Larger Benthic Foraminifera: The Paleogene 389 Superfamily ATAXOPHRAGMIOIDEA Schwager, 1877 Members of this superfamily have a multilocular, trochospiral test that becomes bise- rial or uniserial in later stages. Middle Triassic to Holocene. Family Globotextulariidae Cushman, 1927 Tests have a highly trochospiral stage followed by a quadriserial, triserial or biserial stage. The interior of the chambers may be subdivided by internal partitions. The wall is non-c analiculated. Late Cretaceous (Campanian) to Holocene. • Cubanina Palmer, 1936 (Type species: Cubanina alavensis Palmer, 1936). The test is triserial to uniserial. Late Oligocene to Holocene (Plate 6.1, fig. 1). • Liebusella Cushman, 1933 (Type species:  Lituola nautiloidea Lamarck var. soldanii Jones and Parker, 1860). The test is trochospiral to uniserial. Chambers are slightly overlapping and subdivided by vertical partitions projecting inward from the outer wall and extending from chamber floor to roof. Late Eocene to Holocene (Plate 6.1, fig. 3). • Matanzia Palmer, 1936 (Type species: Matanzia bermudezi Palmer, 1936). The test is trochospiral to biserial. Chambers are subdivided by narrow vertical partitions that radiate inward from the outer wall. Late Cretaceous to Miocene (Plate 6.1, fig. 4). Superfamily PAVONITINOIDEA Loeblich and Tappan, 1961 Tests are coiled in the early stages, and triserial to biserial or uniserial. The interiors of the chambers are partially divided by numerous vertical partitions (beams) or septula, that project downwards from the roof, and rarely may have a few connecting horizontal partitions (rafters). Late Cretaceous and Oligocene to Pliocene. Family Pavonitinidae Loeblich and Tappan, 1961 Tests are palmate, triangular in section, with an early stage this has triserial to biserial or uniserial coiling. Interiors of chambers are partially divided by numerous vertical partitions (beams) or septula that project downwards from the roof and rarely may have a few connecting horizontal partitions (rafters). Apertures are terminal, and sin- gle or multiple. Oligocene to Pliocene. Subfamily Spiropsammiinae Seiglie and Baker, 1984 The test is strongly compressed, but later stages may be uncoiled. Interiors of cham- bers are subdivided by numerous elongate septula. Apertures are single, and terminal. Oligocene to Early Pliocene. • Spiropsammia Seiglie and Baker, 1984 (Type species:  Spiropsammia uhligi (Schubert) = Cyclammina uhligi Schubert, 1901). The early stage is evolute, planispi- rally enrolled, but later uncoiled and rectilinear. The aperture is single, and terminal. Oligocene to Early Pliocene (Plate 6.1, figs 5- 7). Subfamily Pavonitininae Loeblich and Tappan, 1961 The test is triserial, biserial or uniserial. Chambers are undivided by secondary septula. Oligocene to Miocene. 390 Evolution and Geological Significance of Larger Benthic Foraminifera • Pavonitina Schubert, 1914 (Type species: Pavonitina styriaca Scubert, 1914). The test is broad and flattened. A large proloculus is followed by a biserial stage, composed of low and curved chambers, followed by broad and centrally curved uniserial recti- linear chambers. Oligocene to Miocene (Plate 6.1, fig, 8). • Pavopsammia Seiglie and Baker, 1984 (Pavopsammia flabellum Seiglie and Baker, 1984). The test is palmate, flattened, with a proloculus followed by a triserial to bise- rial and finally uniserial stage. Oligocene (Plate 6.1, fig. 9). • Zotheculifida Loeblich and Tappan, 1957 (Type species: Textularia lirata Cushman and Jarvis, 1929). The test is compressed, and biserial throughout. Late Oligocene to Early Miocene (Plate 6.1, fig. 10). Superfamily COSCINOPHRAGMATOIDEA Thalmann, 1951 The test is free or attached, may be coiled in the early stages, but is later uncoiled or branched. The walls are finely agglutinated, traversed by pores, or with a coarsely per- forate or canaliculate inner layer and an outer imperforate layer. Triassic to Holocene. Family Haddoniidae Saidova, 1981 The test is attached. The aperture is terminal, simple to complex. Paleocene to Holocene. • Haddonia Cushman, 1898 (Type species:  Haddonia torresiensis Chapman, 1898). The test is large. Chambers are broad and irregular in shape. The walls are tra- versed by numerous large pores. The aperture is a slit. Early Paleocene to Holocene (Plate 6.1, fig. 12). Superfamily TEXTULARIOIDEA Ehrenberg, 1838 The test is trochospiral, biserial or triserial in early stages and later may be uniserial or biserial. The walls are agglutinated, and canaliculated. Early Jurassic (Sinemurian) to Holocene. Family Chrysalidinidae Neagu, 1968 The test is high trochospiral, with quinqueserial or quadriserial or triserial or biserial coiling modes, or with certain consecutive pairs of these. The aperture is central along the axis of coiling. Subfamily Chrysalidininae Neagu, 1968 This subfamily, as revised by Banner et al. (1991), is essentially triserial throughout its ontogeny (at least in the megalospheric generation), becoming biserial or quadriserial in the adult. The walls are solid but sometimes becoming canaliculate. Early Jurassic (Sinemurian) to Late Eocene. • Pfendericonus Hottinger and Drobne, 1980 (Type species: Lituonella makarskae van Soest, 1942). The early trochospiral stage is multichambered and occupies a third of the test, becoming uniserial in the final stage. Chambers are internally undivided by The Cenozoic Larger Benthic Foraminifera: The Paleogene 391 partitions and connected by thin vertical pillars. The apertural face is convex. Late Paleocene to Eocene (Fig. 6.2). • Pseudochrysalidina Cole, 1941 (Pseudochrysalidina floridana Cole, 1941). The test is initially triserial, but biserial in adult with internal pillars and oblique septa. Eocene (Fig. 6.2). • Vacuovalvulina Hofker, 1966 (Type species: Marssonella keijzeri van Bellen, 1946). It differs from Pseudochrysalidina in having plano-c oncave septa. Paleocene (Fig. 6.2). Superfamily LITUOLOIDEA de Blainville, 1825 Members of this superfamily have a multilocular, rectilinear and uniserial test. The early stage has plani- (strepto-) or trochospiral coiling. The peripheries of the cham- bers have radial partitions, but centrally there are either no or scattered, separated pillars. The aperture is simple or multiple cribrate. Early Jurassic (Sinemurian) to Holocene. Family Cyclamminidae Marie, 1941 The test is involute with alveolar walls. The aperture is near the septal face. Jurassic to Holocene. • Cyclammina Brady, 1879 (Type species:  Cyclammina cancellata Brady, 1879). The test is planispiral, and flattened. The walls are thick, with an alveolar subepider- mal meshwork of a thickness exceeding that of the chamber lumen. Paleocene to Holocene (Plate 6.1, fig. 11). Family Lituolidae de Blainville, 1827 The early stage is enrolled, but later it may be rectilinear. There are few chambers (less than 10) per whorl. Carboniferous to Holocene. Family Spirocyclinidae Munier- Chalmas, 1887 The test is planispiral, becoming peneropliform to annular in later stages. Chambers are partially subdivided by septula. Jurassic to Eocene • Saudia Henson, 1948 (Type species: Saudia discoidea Henson, 1948). Microspheric forms have a uniserial flabelliform last stage. The megalospheric stage has a large proloculus and evolute, annular last stage. Chambers possess a superficial subepider- mal cellular layer (“pigeon- hole” structure). Paleocene to Middle Eocene (Plate 6.1, figs 13- 16). • Thomasella Sirel, 1998 (Type species:  Saudia labyrinthica Grimsdale, 1952). The megalospheric embryo is large and spherical, followed by a few undivided chambers, with successive chambers becoming cyclical. A very narrow exoskeletal subepider- mal layer is present on both sides of the test, consisting of two or more genera- tions of vertical partitions (beams), perpendicular to the septa, and two or more sets of horizontal partitions (rafters) parallel to the septa, producing a complex 392 Evolution and Geological Significance of Larger Benthic Foraminifera 66.0 56.0 33.90 23.03 Age (Ma) Paleocene Eocene Oligocene Danian Seland. Thanetian Ypresian Lutetian Bartonian Priabonian Rupelian Chattian Vacuovalvulina Pseudochrysalidina Pfendericonus Coleiconus Pseudolituonella Coskinon Coskinolina Dictyoconus Anatoliella Fallotella Karsella Daviesiconus Verseyella Cushmania Figure  6.2. The evolution of the Paleogene textulariides. Global distribution is shown by the shading of the background to genera names:  White  =  Cosmopolitan, Light Green  =  Tethys, Dark Grey = Americas. TEXTULARIOIDEA COSKINOLINOIDEA ORBITOLINOIDEA "Letter Epoch Stages" - Stage L Lower Te Td E Tc L Tb Ta3 M E Ta2 L Ta 1(b) M ? E Ta 1(a) Chysalinid Arenobulimina The Cenozoic Larger Benthic Foraminifera: The Paleogene 393 subepidermal network with numerous alveolar compartments near the surface. The central (labyrinthic) zone consists of numerous pillars aligned from one chamber to the next. Early Eocene (Plate 6.1, fig. 17). Superfamily COSKINOLINOIDEA Moullade, 1965 Members of this superfamily have a conical test with a trochospiral early stage, becom- ing uniserial and rectilinear. Late Cretaceous to Middle Eocene (Cenomanian to Lutetian). Family Coskinolinidae Moullade, 1965 Simple, or without, exoskeleton in the marginal chamber cavity. The rectilinear part has broad, low chambers, subdivided by pillars or irregular partitions. The aperture is a series of pores scattered over the apertural face. Late Cretaceous to Midlle Eocene (Cenomanian to Lutetian). • Barattolites Vecchio and Hottinger, 2007 (Type species:  Barattolites trentinaren- sis Vecchio and Hottinger, 2007). The test is highly conical with a trochospiral early part followed by a uniserial part with a slightly convex base. Chambers are subdivided by vertical partitions more or less in line from one chamber to the next. The megalospheric apparatus has simple walls and consists of two cham- bers divided by a very thin, straight septum, separating the protoconch from the deuteroconch. The following five to six chambers constitute a trochospiral nepi- ont with an exoskeleton beginning at the third chamber. Early to Middle Eocene (Plate 6.1, fig. 18). • Coleiconus Hottinger and Drobne, 1980 (Type species: Coskinolina elongata Cole, 1942). The early stage is trochospirally enrolled, as in Arenobulimina, but without any partitions; later parts have few scattered pillars. There is a simple exoskeleton (beams only) in the marginal chamber cavity. Early to Middle Eocene (Figs 6.2, 6.3, 6.4; Plate 6.1, figs 19- 21; Fig. 6.2). • Coskinolina Stache, 1875 (Type species:  Coskinolina liburnica Stache, 1875, = Lituonella roberti Schlumberger, 1905). The early stage is arenobuliminid, later uniserial with numerous chambers. The inside of the test has vertical partitions and pillars. There is no exoskeleton in the marginal chamber cavity. Paleocene to Middle Eocene (Figs 6.2, 6.3; Plate 6.1, figs 22- 23; Plate 6.2, figs 1- 2, 19; Plate 6.3, fig. 5A). • Coskinon Hottinger and Drobne, 1980 (Coskinolina (Coskinon) rajkae Hottinger and Drobne, 1980). There is a very reduced arenobuliminid early growth stage, with later parts being uniserial with scattered pillars. There is no exoskeleton in the marginal chamber cavity. Middle Paleocene to Middle Eocene (Plate 6.2, fig. 3; Figs 6.2, 6.3). Superfamily ORBITOLINOIDEA Martin, 1890 The test is conical with numerous chambers, partially subdivided by radial or trans- verse partitions or pillars. Middle Jurassic to Oligocene. 394 Evolution and Geological Significance of Larger Benthic Foraminifera Family Orbitolinidae Martin, 1890 There is an initial, low trochospire, usually very much reduced, followed by a later recti- linear, broad and conical growth, made of low uniserial chambers subdivided with pillars or vertical partitions. The aperture is cribrate. Middle Jurassic to Oligocene (Fig. 6.3). • Anatoliella Sirel, 1988 (Type species: Anatoliella ozalpiensis Sirel, 1988). The test is a low to high conical form in late ontogeny, with three chambers in each whorl. The embryonic apparatus is followed by a series of trochospiral chambers, subdivided by a network of vertical beams and horizontal partitions. The central zone is divided by numerous pillars. Anatoliella differs from all orbitolinid genera in having just a single series of shallow cuplike chambers in its triserial chamber arrangement. Paleocene (Plate 6.2, fig. 4; Fig. 6.2). • Cushmania Silvestri, 1925 (Type species: Conulites americana Cushman, 1919). The early trochospiral is very reduced, with an apical protoconch. Chambers are divided by short vertical partitions, which are intersected by at least two horizontal parti- tions. Middle Eocene (Plate 6.2, fig. 5). Genus Axial secon Horizontal basal secon No No vercal horizontal parons parons Pillars Coskinolina (= Lituonella) One cycle No of vercal horizontal parons Coleiconus parons Pillars May have 2 cycles of horizontal vercal parons parons Fallotella Pillars Several cycles Several cycles of horizontal of peripheral parons vercal parons Dictyoconus Pillars Figure 6.3. The distribution of the vertical and horizontal partitions in the coskinolinids and orbitolinids. The Cenozoic Larger Benthic Foraminifera: The Paleogene 395 • Daviesiconus Hottinger and Drobne, 1980 (Type species: Coskinolina balsilliei Davies, 1930). The initial part of the test is almost planispiral. Walls have vertical partitions in the marginal zone, which Hottinger (2007) interpreted as septula comprising a part of the endoskeleton, but lacking horizontal partitions. Early to Middle Eocene (Plate 6.2, fig. 6.6). • Dictyoconus Blanckenhorn, 1900 (Type species:  Dictyoconus egyptiensis (Chapman) = Patellina egyptiensis (Chapman)). Radial partitions thicken away from the periphery, and become broken up into pillars in the central zone. Peripheral tiered rectangular chamberlets are present. Aptian to Oligocene (Figs 6.3, 6.4; Plate 6.2, figs 7- 9, 18; Plate 6.28, figs 5- 6). Early stage trochospirally enrolled Vertical partition B A Marginal chambers Pillars Embryonic apparutus D C Figure 6.4. Features of conical foraminifera. A-B ) Coleiconus christianaensis Robinson, Middle Eocene, Upper Chapelton Formation, Jamaica, type figures, NHM P52808- 9; C-D ) Coskinolina cf. douvillei (Davies), Middle Eocene, Upper Chapelton Formation, Jamaica, UCL coll. 396 Evolution and Geological Significance of Larger Benthic Foraminifera • Fallotella Mangin, 1954 (Type species:  Fallotella alayensis Mangin, 1954). The early trochospiral coil is very reduced. The uniserial part is pillared, and has a thick marginal wall. The exoskeleton is simple (only beams) or moderately complex (beams and rafters). Middle to early Late Paleocene (Figs 6.2, 6.3; Plate 6.2, fig. 17; Fig. 6.2). • Karsella Sirel, 1997 (Type species: Karsella hottingeri Sirel, 1997). This form differs from the Late Cretaceous Calveziconus Caus and Cornella, 1982 (see Chapter 5) in having a more complex internal structure, with more than two vertical (beams) and horizontal (rafters) partitions forming an irregular network under the epidermis. It also possesses a large protoconch and periembryonic chambers. It differs from Orbitolina d’Orbigny, 1850 (see Chapter  5) in having a more complex exoskeletal structure and from Cushmania Silvestri 1925 in having a shorter trochospiral early stage. Paleocene (Thanetian) (Plate 6.7, fig. 3; Fig. 6.2). • Verseyella Robinson, 1977 (Type species:  Coskinolinoides jamaicensis Cole, 1956). The early trochospiral stage is absent. The test is biserial in the early part, becoming uniserial in later stages. The interiors of chambers are subdivided by vertical parti- tions (beams) that are aligned from chamber to chamber, forming a ring around the central shield which is supported by pillars. Early Eocene (Plate 6.2, figs 10-1 1). ORDER MILIOLIDA Delage and Hérouard, 1896 The miliolides have tests that are porcelaneous, imperforate, and made of high Mg- calcite with fine randomly oriented crystals. They range from the Carboniferous to the Holocene. Superfamily ALVEOLINOIDEA Ehrenberg, 1839 The test is enrolled along an elongate axis, and initially planispiral or streptospiral, or milioline, with chambers added in varying planes. Cretaceous to Holocene. Family Alveolinidae Ehrenberg, 1839 The test is free, large, planispiral to fusiform, subcylindrical or globular, and coiled about elongate axis. Early whorls may be irregular, streptospiral in monomorphic spe- cies, and may be restricted to microspheric forms in dimorphic species. Chambers are subdivided into chamberlets by longitudinal partitions (septula) perpendicular to the main septa, and connected by passages below the apertural face. The basal layer may include anastomosing canals (Hottinger et al., 1989), and may have thick deposition of secondary calcite (flosculinisation) (Figs 6.5, 6.6). The aperture is a slit parallel to the base of the apertural face, or a single row of circular openings, or numerous rows of such openings in horizontal rows, arranged in definite patterns matching the patterns of chamber divisions. Early Cretaceous (Aptian) to Holocene. • Alveolina d’Orbigny, 1826  =  Fasciolites Parkinson 1811 (Type species:  Oryzaria boscii Defrance, in Bronn, 1825). The test is large, fusiform or cylindrical with sin- gle tier chamberlets alternating in position in successive chambers. The basal layer is thick, as deposition of secondary calcite (flosculinisation) fills most of the chamber lumen. Successive chambers communicate with post- and pre- septal passages. The Cenozoic Larger Benthic Foraminifera: The Paleogene 397 Flosculin spherical alveolinids tight first whorls loose first whorls A. globosa group A. avellana group tight first whorls and strong flosculinisation tight first and last whorls A. pasticillata group A. indicatrix group tight first whorls and little flosculinisation A. minervensis group Oval alveolinids thin basal layer thick basal layer A. ellipsoidalis A. subpyrenaica Figure 6.5. Basal layers and flosculinisation in Alveolina. The final chamber has two rows of apertures alternating in position. Late Paleocene to Eocene (Figs 6.5, 6.6; Plate 6.2, figs 15-1 9; Plate 6.3, figs 1-4 , 6; Plate 6.4, figs 1-1 1; Plate 6.5, figs 1-2 , 12, 14; Plate 6.6; figs 1- 2; Plate 6.7, fig. 9; Plate 6.8, fig. 3A). • Borelis de Montfort, 1808. (Type species: Borelis melonoides de Montfort, 1808)The test is small, spherical to fusiform. It differs from Alveolina in having only a pre- septal passage and a secondary small tier of chamberlets which alternate with the larger ones producing “Y” shaped septa in axial section. The aperture is a single row of pores. Late Eocene to Holocene (Plate 6.5, figs 5A, 6-8 ). 398 Evolution and Geological Significance of Larger Benthic Foraminifera Proloculus Post-septal canal Pre-septal canal Septum C A Streptospiral nepiont B Flexostyle Single row of chamberlets D Basal layer/flosculin E Figure 6.6. Alveolina oblonga d’Orbigny. Chaussy, Val-d ’Oise, France; Figs A, B) equatorial sections; C) an enlargement of part of B; D-E ) axial sections. Scale bars: Figs A-B  = 0.5mm; C = 0.25mm; D = 1mm; E = 160μm. The Cenozoic Larger Benthic Foraminifera: The Paleogene 399 • Bullalveolina Reichel, 1936 (Type species: Alveolina bulloides d’Orbigny, 1839). The test is tiny, globular, with a streptospiral early stage which later is planispi- ral. Large pre-s eptal passages occupy about one half of the chamber floor. The final chamber has three or more rows of small apertures. Oligocene (Plate 6.5, fig. 9). • Globoreticulina Rahaghi, 1978 (Type species: Globoreticulina iranica Rahaghi, 1978). The test is globular, planispiral, and involute with an alveolar exoskeleton and only the final whorl visible on the exterior. The outer parts of the chambers are subdi- vided by parallel and transverse partitions. The aperture is cribrate. Middle Eocene (Plate 6.5, fig. 10). • Glomalveolina Hottinger, 1962 (Type species: Alveolina dachelensis Schwager,1883). The test is small, globular, consisting of a small globular proloculus followed by early streptospiral coiling, but is later planispiral. As in Alveolina, pre- septal and post- septal passages are present, however, unlike Alveolina, flosculinisation is mini- mal. The final chamber has a row of openings, intercalated with smaller ones. Middle Paleocene to Middle Eocene (Plate 6.2, fig. 18; Plate 6.9, fig. 1). • Malatyna Sirel and Acar, 1993 (Type species: Malatyna drobneae Sirel and Acar, 1993). The test shows a triloculine early stage arrangement in megalospheric gen- erations and a quinqueloculine early stage arrangement in microspheric genera- tions. In both generations the planispiral adult stage has inflated chambers with subepidermal partitions. The aperture is cribrate. Eocene (Lutetian) (Plate 6.9, fig. 2). • Praebullalveolina Sirel and Acar, 1982 (Type species: Praebullalveolina afyonica Sirel and Acar, 1982). The test is spherical with early streptospiral coiling. Later cham- bers have septula complete from floor to roof. The pre-s eptal passage is large. The apertural face has one row of primary apertures intercalated with smaller secondary apertures. Late Eocene (Plate 6.6, fig. 3). Family Fabulariidae Ehrenberg, Munier- Chalmas, 1882, emend. Hottinger et al., 1989 The test is large, dimorphic, multichambered with miliolide coiling, tending to become reduced in subsequent growth stages, either to bilocular or to monolocular cham- ber cycles. They have a trematophore aperture (see Fig.  6.7), which is defined thus (Hottinger, 2006): “A trematophore or a sieve constituting the face of many porcelaneous larger fora- minifera, in miliolines produced by the coalescence of teeth, covering a large pre-s eptal space. May be supported by residual pillars”. Chambers show a thickened basal layer, subdivided by pillars or secondary parti- tions. Foraminifera in adult growth stages have a fixed apertural axis. For further ter- minology and detailed structural analysis see Drobne (1974; 1988). • Fabularia Defrance, 1820 (Type species: Fabularia discolites Defrance,1825). The test is ovate, quinqueloculine in early stages and biloculine in later stages. Chambers have thick walls and are subdivided by subepidermal partitions, forming a series of cham- bers that are connected by pre- and post-s eptal passages. The aperture is cribrate (trematophore). Middle to Late Eocene (Plate 6.9, figs 4- 7). 400 Evolution and Geological Significance of Larger Benthic Foraminifera • Lacazina Munier-C halmas, 1882 (Type species:  Alveolina compressa d’Orbigny, 1850). The test is large, discoid to ovoid, and dimorphic. The early miliolide coil- ing is followed by chambers that partially embrace the earlier ones. Chambers are subdivided by longitudinal partitions. Longitudinal ribs are directed from one foramen to the next, so that each supports a single row of regularly spaced pil- lars, reaching the chamber roof and alternating in their position in adjacent rows. The aperture is cribrate (trematophore) at one extremity of the test. Lacazina is distinguished from Pseudolacazina (below) by the monocular-c oncentric cham- ber arrangement and by the alternating positions of its pillars. It is distinguished from Fabularia by its monolocular growth and by the presence of radial parti- tions. Late Cretaceous (Coniacian) to Early Oligocene (Plate 6.9, fig. 9; Plate 6.27, fig. 9). • Lacazinella Crespin, 1962 (Type species: Lacazina wichmanni Schlumberger, 1894). The test is ovoid, with concentric chambers completely embracing and divided by low longitudinal partitions that do not reach the chamber floor and support the trematophore. Late Paleocene to Eocene (Plate 6.9, figs 10-1 1). • Neolacazopsis Matsumaru, 1990 (Type species: Neolacazopsis osozawai Matsumaru, 1990). The test is ovate in outline. A large proloculus is followed by chambers in a biloculine arrangement in megalospheric forms, while microspheric forms are quin- queloculine to triloculine and finally become biloculine adults. The chambers are subdivided into arcuate to turtle- neck bottle- like form chamberlets. Walls have a finely to coarsely alveolar inner layer, and outer layers with perforations. The aper- ture is cribrate (trematophore). Middle to Late Eocene. • Pseudofabularia Robinson, 1974 (Type species: Borelis matleyi Vaughan, 1929). The test is almost globular, and bilocular. The interiors of chambers are subdivided by longitudinal partitions spiralling around the test and dividing the chambers into elongate chamberlets. Middle Eocene (Plate 6.9, fig. 12). • Pseudolacazina Caus, 1979 (Type species: Pseudolacazina hottingeri Caus, 1979). The test is globular to ovate, and dimorphic. Megalospheric forms are biloculine (with completely embracing chambers) throughout their ontogeny. Microspheric forms are quinqueloculine in the early stage, but later they show biloculine to monoloculine cycles as adult. The chambers are subdivided by longitudinal partitions or septula or by ribs with pillared extensions supporting the chamber roof. Late Cretaceous (Santonian) to Eocene (Fig. 6.7). Superfamily MILIOLOIDEA Ehrenberg, 1839 The test is coiled in varying planes with two chambers per whorl, with the axis of coil- ing normal to the apertural axis and rotated, so that several angles exist between the median planes of consecutive chambers, such as 72° (quinqueloculine), 120° (trilocu- line) or 180° (spiroloculine or biloculine). The test may become uncoiled, cylindrical or compressed with partial partitions. The proloculus is followed by a spiral passage. The aperture is single, and may be accompanied by additional teeth that project from the opposite margins of the aperture, from the chamber roof or from the lateral wall, or with a sieve (trematophore), that in fact is present in many porcelaneous larger foraminifera. The Cenozoic Larger Benthic Foraminifera: The Paleogene 401 Idalina Periloculina A simple Late Cretaceous form A Late Cretaceous with a quinqueloculine to biloculine stage adult completely overlapping, with last chamber overlapping. vertical partitions incomplete. Lacazina Late Cretaceous - Early Oligocene form eqs overlapping throughout. axs eqs Miliola tr an Eocene form with a quinqueloculine test. axs Fabularia A Middle to Late Eocene uinqueloculine to involute biloculine test. apax Pseudolacazina pi Santonian, Middle to Late s Eocene quinquelocune to biloculine and a Lacazinella monoloculine. Late Paleocene - Eocene completely overlapping chambers throughout. Figure 6.7. Sketches of Paleogene Fabulariidae, modified from Hottinger (2006), Drobne (1989) and Loeblich and Tappan, 1988. 402 Evolution and Geological Significance of Larger Benthic Foraminifera Family Austrotrillinidae Loeblich and Tappan, 1986 This family has a miliolide test with coarse alveolar walls. Middle Eocene to Middle Miocene. • Austrotrillina Parr, 1942 (Type species: Trillina howchini Schlumberger, 1893). The test is triloculine with fine to coarse blind alveoles that can bifurcate. The aperture is a simple tooth in the early chambers, branching to form smaller openings in the adult. Middle Oligocene to Middle Miocene (Langhian) (Plate 6.3, figs 7- 9; Plate 6.9, figs 15- 17; Plate 6.10, figs 1-2 ). Family Hauerinidae Schwager, 1876 The early part of the test has a globular proloculus followed by two chambers per whorl. The chambers may be added in a quinqueloculine arrangement, but later may be uncoiled. The aperture may range from a simple to a bifid tooth, or may be a trema- tophore. Jurassic to Holocene. • Heterillina Munier-C halmas, 1905 (Type species:  Heterillina guespellensis Schlumberger, 1905). The test is rounded, early chambers have a quinqueloculine arrangement, while later it is planispiral and evolute. The aperture is a trematophore. Middle Eocene to Oligocene (Plate 6.10, fig. 9). • Kayseriella Sirel, 1999 (Type species: Kayseriella decastroi Sirel, 1999). The len- ticular test has a milioline early stage (with a quinqueloculine embryont in the microspheric forms, and a triloculine arrangement in the megalospheric forms), followed by undivided planispiral chambers. The final stage is rectilinear. The aperture is simple, with a single basal opening with teeth and a single aperture with thick ribs in the uniserial chambers of the adult stage. This form is similar to the Late Cretaceous Scandonea De Castro 1971 (see Chapter 5), but the latter has thin subepidermal partitions that appear in the axial, oblique and horizon- tal sections of the uniserial chambers (De Castro 1971, Sirel, 1999). Kayseriella differs from Heterillina in having a simple, single aperture with teeth in the plani- spiral stage and uniserial chambers with a final ribbed aperture. Early Paleocene (Danian). Family Rivieroinidae Saidova, 1981 The test is planispiral, and ovate in outline with chambers that are subdivided by oblique sutures. Middle Eocene to Holocene. • Pseudohauerina Ponder, 1972 (Type species: Hauerina occidentalis Cushman, 1946). The test is quinqueloculine in the early stage with chambers one half-c oil in length; later the chambers are planispiral with more than two chambers in each adult whorl. The interior of the test is subdivided by numerous incomplete radial subepidermal partitions that project inward from the walls for about one- third of the breadth of the chamber. The adult test has a complex trematophore with many openings. Oligocene to Holocene. The Cenozoic Larger Benthic Foraminifera: The Paleogene 403 Family Spiroloculinidae Wiesner, 1920 A planispiral test consisting of a cornuspirine flexostyle, followed by a biserial part. The aperture is a simple, single basal opening with a bifid tooth. • Elazigella Sire1, 1999 (Type species: Elazigella altineri Sirel, 1999). A lenticular test has umbonal thickening on both sides and a triangular apertural opening with a slender tooth. Late Paleocene (Thanetian) (Plate 6.10, Fig. 7). Superfamily SORITOIDEA Ehrenberg, 1839 The chambers are planispiral, uncoiling, flabelliform or cyclical, and may be subdi- vided by partitions or pillars. Late Permian to Holocene. Family Peneroplidae Schultze, 1854 The test is closely coiled in the early stage, becoming uncoiled in the later stage. Chambers have a simple interior. The aperture is single, rounded, slit-l ike or multiple. Late Cretaceous to Holocene. • Archiacina Munier-C halmas, 1878 (Type species: Cyclolina armorica d’Archiac, in Tournouër 1868). The test is a large discoid, planispiral, and semi- involute, later chambers are of peneropline shape and then cyclical. The aperture is multiple. Oligocene (Plate 6.10, fig. 4). • Dendritina d’Orbigny, 1826 (Type species:  Dendritina arbuscula d’Orbigny, 1826). The test is planispiral, and involute. The surface has numerous striae. A single areal aperture is modified by a heavily folded peristome. Middle Eocene to Holocene (Plate 6.10, fig. 5). • Haymanella Sirel, 1998 (Type species: Haymanella paleocenica Sirel, 1999). The test is porcelaneous but with superficial, coarse agglutinated grains. The cham- bers are partially subdivided by irregularly disposed, radial, short partitions. The terminal stellate aperture has a protruding peristome. Paleocene (Plate 6.10, fig. 6). • Hottingerina Drobne, 1975 (Type species: Hottingerina lukasi Drobne, 1975). A len- ticular, dimorphic peneroplid, planispiral and involute test in the early stage, with later chambers arranged in an uniserial pattern in both megalospheric and micro- spheric generations. The interior of the chambers is subdivided by thin, short subepi- dermal partitions. The aperture is a simple, basal slit in the planispiral stage. Middle to Late Paleocene (Plate 6.10, fig. 10). • Penarchaias Hottinger, 2007 (Type species:  Peneroplis glynnjonesi Henson, 1950). The test is lenticular, composed of numerous planispiral- involute chambers with alar prolongations. The chambers are undivided and arranged in tight coils throughout ontogeny. In the alar prolongations of the chambers there is a single, interiomarginal row of apertures alternating with low endoskeletal ridges of the basal layer, that are perpendicular to the septal wall. The apertural face has multiple apertures. Late- Middle Eocene. • Peneroplis de Montfort, 1808 (Type Species Nautilus planatus Fichtel and Moll, 1798). The test is planispiral in the early part, tightly coiled and semi-i nvolute, 404 Evolution and Geological Significance of Larger Benthic Foraminifera lenticular and compressed, becoming uncoiled and flaring, with an umbilical depres- sion. However, only the mineralized shell has this involute tendency, the chamber lumina are evolute and thus have no alar prolongations (Hottinger, 2007). The aperture is a single row in the median line. These apertures may be modified and/o r subdivided by folded, partially fused peristomes (Hottinger et al., 1993). Eocene to Holocene (Fig. 6.8). • Puteolina Hofker, 1952 (Type species: Peneroplis proteus d’Orbigny, 1839). The test is strongly flaring with a simple interior. It is considered by Loeblich and Tappan (1988) to be a synonym to the Miocene form Laevipeneroplis Šulc, 1936. Oligocene to Holocene (Fig. 6.8). INDO-PACIFIC TROPICAL-SUBTROPICAL SHELF SEAS Marginopora REEFS Middle Miocene - Holocene Sorites Oligocene - Holocene Cyclorbiculina BACK-REEFS Oligocene - Holocene COSMOPOLITAN TETHYS Puteolina Oligocene - Archaias Holocene ? Orbitolites Early - Middle Eocene Spiroloculina SEA-GRASS ENVIRONMENTS Peneroplis Late Cretaceous - Holocene Eocene - Holocene Figure 6.8. The progression from a peneropline shell to a completely annular test with chambers divided by internal partitions/ crosswise- obliquely arranged endoskeletal elements in the Soritoidea (Middle Eocene to recent) with Sorites - Amphisorus – Marginopora, and to the archaiasines with radial endoskeletal elements. The latter have no marginal subdivision of the chambers. The Cenozoic Larger Benthic Foraminifera: The Paleogene 405 • Raoia Matsumaru and Sarma, 2008 (Type species:  Raoia indica Matsumaru and Sarma, 2008). Interior of chambers undivided, and marginal zone of chambers sub- divided by rudimentary radial septula. Paleocene (Plate 6.11, fig. 12). • Spirolina Lamarck, 1804 (Type species: Spirolina cylindracea Lamarck, 1804). The test is elongate, with early chambers that are planispiral, but which later are uncoiled and simple. The aperture is single, rounded. Eocene to Holocene. Family Soritidae, Ehrenberg, 1839 The test is involute, planispiral to uncoiled evolute, flaring, annular discoid with par- tial or complete partitions. The aperture is multiple. Late Cretaceous (Cenomanian) to Holocene. • Amphisorus Ehrenberg, 1839 (Type species: Amphisorus hemprichii Ehrenberg, 1839). The test is large biconcave, circular with thickened edges and two partial layers of chamberlets. The megalospheric apparatus consists of a large deuteroconch embrac- ing a protoconch, and its wide-o pen flexostyle has a hemicylindrical to almost cylin- drical frontal wall bearing numerous apertures. Cyclical chambers are divided by septula. Microspheric forms have a peneropline early stage with six undivided cham- bers, followed by ten chambers subdivided by septula that then become annular. Axial sections show an abrupt increase in the irregularity and volume of the cham- berlet cavity in the last few chamberlet cycles. Crosswise-o blique stolons connect successive chambers. The aperture consists of numerous pores aligned in two alter- nating rows. Median apertures between the double rows of apertures may be present. Latest Oligocene to Holocene (see Chapter 7). • Archaias de Montfort, 1808 (Type species:  Archaias spirans de Montfort, 1808 = Nautilus angulatus Fichtel and Moll, 1798), as defined by its type species Archaias angulatus (Fichtel and Moll), extensively emended by Rögl and Hansen, 1984. The test is compressed, planispiral and involute, and may be partially evolute in the last whorls, with a thickened middle part and radial endoskeletal elements. Multiple apertures are flanked by irregular free and interseptal pillars. The subepi- dermal partitions are incomplete, and tests lack exoskeletal structures and no mar- ginal subdivision of the chambers. Middle Eocene to Holocene (Fig. 6.8; Plate 6.10, Figs 8, 11, 12; Plate 6.12, Fig. 4). • Cyclorbiculina Silvestri, 1937 (Type species: Orbiculina compressa d’Orbigny 1839.). The test is large, and discoidal with a peneropline early stage becoming annular through most of its adult growth in both microspheric and megalospheric forms, with regular partitions. Oligocene to Holocene (Fig. 6.8; see Chapter 7). • Cyclorbiculinoides Robinson, 1974 (Type species:  Cyclorbiculinoides jamaicensis Robinson, 1974). The test is large discoid. The peneropline stage is followed by cycli- cal chambers with partitions aligned in successive chambers rather than alternating as in Cyclorbiculina. Middle to Late Eocene (Plate 6.10, Fig. 13). • Mardinella Meriç and Çoruh, 1991 (Type species: Orbitolites shirazensis Rahaghi, 1983). This form has vertical septula dividing the chambers into small chamberlets instead of the oblique ones in Orbitolites, while the diagonal stolons connecting the chambers in different rows are not divided by horizontal septula. Late Paleocene (Thanetian). 406 Evolution and Geological Significance of Larger Benthic Foraminifera • Neorhipidionina Hottinger, 2007 (Type species: Rhipidionina williamsoni Henson, 1948). A planispiral test which has an uncoiled flaring adult part. Chambers are subdivided by septula that are perpendicular to the outer chamber wall and end proximally with a slightly thickened rim. It has cribrate apertures on the apertural face with additional apertures appearing in the median plane as one or several rows. The undivided median zone of the chamber is restricted to a median annu- lar passage. Neorhipidionina has a single foramen. This in contrast to Rhabdorites which between the septula has a radial row of foramina that almost reaches the periphery of the discoidal chambers. Middle Eocene (Plate 6.6, figs 8-9 ; Plate 6.10, fig. 14). • Neotaberina Hottinger, 2007 (Type species Neotaberina neaniconica Hottinger, 2007). The test is elongate to conical. The adult chambers are saucer- shaped and subdivided by radial partitions that are interpreted by Hottinger as “septula of an endoskeleton because they alternate regularly with radial rows of foramina”, and are arranged in an uncoiled, uniserial sequence. The apertural face is strongly convex and covered by numerous areal apertures. Late Middle Eocene (Plate 6.10, fig. 14). • Opertorbitolites Nuttall, 1925 (Type species: Opertorbitolites douvillei Nuttall, 1925). The test is lenticular, later chambers becoming cyclical and divided into numerous small chamberlets. Compact umbonal thickening, made of lateral laminae, covers the umbilical regions. Late Paleocene to Early Eocene (Plate 6.7, figs 1- 5, 8). • Orbitolites Lamarck, 1801 (Type species:  Discolites concentricus de Montfort, 1808  =  Orbitolites complanatus Lamarck, 1801). The test is a large discoid, very slightly concave, with a large proloculus and inflated nucleoconch, followed by cyclic chambers divided into small numerous chamberlets with curved thickened walls. Adjacent chambers are not interconnected with stolons, but the connections are between the obliquely adjoining chamberlets. According to Hottinger (2006) this superposition is clearly visible where the transverse section is tangential to an annu- lar septum. Early to Middle Eocene (Plate 6.6, fig. 1; Plate 6.7, figs 6- 7, 9; Plate 6.8, figs 1- 3B; Plate 6.12, fig.3;). • Praerhapydionina Van Vessem, 1943 (Type species:  Praerhapydionina cubana Van Vessem, 1943). The test is elongate, sub- conical, planispiral, and later uncoiled and circular in thin section. The interior of the chambers is subdivided by radial parti- tions that are aligned from one chamber to the next. Additional radial partitions are intercalated between the septula, and are shorter than the primary septula, and do not reach the aperture. A single terminal aperture, on a strongly convex face, has a petaloid to stellar outline with four to six rays. Between the petals of the aperture, peristomes protrude to form toothlike features. Middle Eocene (Lutetian) to Early Miocene (Plate 6.8, figs 4-6 ; Plate 7.3, fig. 18). • Rhabdorites Fleury, 1996 (Type species Rhapydionina malatyaensis Sirel, 1976). A soritid test, with a short planispiral early part followed by a cylindrical to conical uniserial part with radial septula, and a large central pre- septal space. The primary septula of the uniserial chambers are very long, aligned in successive chambers and appear to meet in the center of the chamber, but the remainder of the central por- tion of the chambers remains free of partitions. The genus shows multiple apertures. Middle Eocene (Plate 6.6, figs 3- 5, 7; Plate 6.8, fig. 7). The Cenozoic Larger Benthic Foraminifera: The Paleogene 407 • Somalina Silvestri, 1939 (Type species: Somalina stefaninii Silvestri, 1939). The test is lenticular with cyclic chambers divided into chamberlets. The prominent lateral laminae enclose numerous cavities/c hamberlets connected with stolons to the main equatorial chamber layer. Eocene (Lutetian) (Plate 6.11, figs 1- 2). • Sorites Ehrenberg, 1839 (Type species: Nautilus orbiculus Forsskal, 1775). The test is a large, discoid, with an early peneropline stage. Annular chambers are divided into numerous curved to rectangular small chamberlets, which are connected to each other and to those in adjacent chambers by stolons. The aperture is a single row of paired apertures. Oligocene to Holocene. (Plate 6.12, fig. 6). • Taberina Keijzer, 1945 (Type species: Taberina cubana Keijzer, 1945). The test is elon- gate, planispiral becoming uncoiled. All chambers are subdivided by radial parti- tions, while pillars occupy the central areas of the chambers. Early Paleocene. • Twaraina Robinson, 1993 (Type species: Twaraina seigliei Robinson, 1993). The test has a planispiral, compressed, early stage peneropliform, with a later part having flaring chambers crossed by irregular pillars, but lacking a cyclical stage. Eocene (Plate 6.11, figs 3- 6). • Yaberinella Vaughan, 1928 (Type species: Yaberinella jamaicensis Vaughan, 1928). The test is large, operculine to discoid, with a rapidly flaring peneropline early stage that may become cyclical in later stages. Chambers are numerous, and sub- divided by oblique septula into small chamberlets that communicate through stolons. Middle to Late Eocene (Plate 6.6, fig. 11; Plate 6.10, fig.16; Plate 6.11, figs 9- 11). ORDER ROTALIIDA Delage and Hérouard, 1896 The tests are multilocular with a calcareous wall, of perforate hyaline lamellar cal- cite. They have apertures that are simple or have an internal tooth- plate. Triassic to Holocene. Superfamily NUMMULITOIDEA de Blainville, 1827 The test is planispiral or cyclic, lenticular multicamerate, with septal flap and canalicu- lated septa. A spiral marginal cord and spiral canal system is present in early forms, but is modified in advanced forms or replaced by intraseptal canals (Figs 6.9-12). Paleocene to Holocene. Family Pellatispiridae Hanzawa, 1937 A planispiral test, having no marginal cord, but radial and vertical canals or fissures are present. Spiral and umbilical sides are not differentiated. Planispiral- evolute chambers are connected by a single intercameral foramen. The thickened shell mar- gin produced by a marginal canal sytem, as in Pellatispira (Fig. 6.10), may be over- grown by supplemental chamberlets either on the lateral flanks alone (Biplanispira, Fig. 6.11) or on all sides of the shell (Vacuolispira, Plate 6.14, fig. 7)). Paleocene to Eocene. • Biplanispira Umbgrove, 1937. (Type species: Heterospira mirabilis Umbgrove, 1936). Biconvex, early chambers, with a planispiral involute test in the early stage, and 408 Evolution and Geological Significance of Larger Benthic Foraminifera marginal cord septum c a b f alar prolongaon canals e D d P filaments reculate filaments radial filaments radial to meandrine granulaons between meandrine septa i involute tests h high chambers canal system in marginal cord l g j low chambers k planispiral coil trabeculae granulaon on septa reculaons m n intraseptal canal system o p evolute tests r straight,s canaliculated septa q t chamber divided into chamberlets inially involute involute cyclic chambers nepionic morphology like Heterostegina Figure 6.9. Figured by BouDagher- Fadel and Price (2014). a, Nummulites deserti De La Harpe, Egypt, P5, UCL coll., 4mm longest diameter (LD); b, Nummulites irregularis Deshayes, France, P8, UCL coll., 4mm LD; c, Nummulites sp., axial section Barton Bed, Chewton Bunny, Highcliffe, P8, UCL coll., 1.5mm LD; d-e , Nummulites fichteli Michelotti: d, France, P11, axial section, NHM P49522 (also figured by BouDagher- Fadel 2008), 3mm LD: e, solid specimen, Tang- i- Puhal area, P11, UCL coll., 5mm LD; f, enlargement of Nummulites sp. showing protoconch (P), deuteroconch (D), UCL coll., scale bar on figure; g- j, Nummulites sp., g, Egypt, Eocene, UCL coll., 10mm LD, j, Gernona, Spain, Eocene, 8mm, UCL coll. (also figured by BouDagher- Fadel 2008); h, Nummulites intermedius (d’Archiac), India, Eocene, NHM P30148, (also fig- ured by BouDagher- Fadel 2008),12mm LD; i, Nummulites gizehensis (Forskal), Spain, Late Lutetian, UCL coll., (also figured by BouDagher- Fadel 2008),17mm LD; k, Nummulites fichteli- intermedius (d’Archiac), Lower Nari Formation, Pakistan, Oligocene, UCL coll. (also figured by BouDagher- Fadel 2008), 2.5mmLD; l, enlargement of chambers of Nummulites sp., France, Middle Eocene, width of field view 1.5mm; m- n, Operculina aegyptiaca Hamam, m, solid specimen, Egypt, Early Eocene, UCL coll., 2mm LD, n, axial section, megalospheric form, latest Early Eocene, Gebel Gurnah, Luxor, Egypt, paratype, NHM P49827 (also fig- ured by BouDagher- Fadel 2008), 2.2mm LD; o- p, Assilina daviesi de Cizancourt, Lower Bhadrar Beds (Salt Range), Pakistan, Early Eocene, NHM coll. (also figured by BouDagher- Fadel 2008): o, equatorial setion, NHM P41529, 2.4mm LD: p, axial section, NHM P41524, 3.4mm LD; q, Heterostegina (Heterostegina) sp., Brazil, Eocene, UCL coll., 2mm LD; r, Heterostegina (Vlerkina) borneensis van der Vlerk, axial section, Borneo, Late Oligocene, UCL coll., 6mm LD; s, Cycloclypeus eidae Tan Sin Hok, Kinabatangan River, Sabah, North Borneo, Early Miocene, NHM coll., NB9067, enlargement of early part of test, width of field view 1mm; t, Cycloclypeus carpenteri Brady, off Jutanga, Holocene, UCL coll., 6mm LD. The Cenozoic Larger Benthic Foraminifera: The Paleogene 409 Radial canals 4 5 3 1 6 2 7 8 11 9 10 Marginal crest Chambers Enveloping canals Figure 6.10. Bilamellar Pellatispira fulgeria Whipple, Late Eocene, Sumatra, equatorial section showing chamber arrangement, marginal crest and lateral chamberlets, early spiral chambers numbered 1- 11 (1 being the protoconch), UCL coll., AS33 spiraling into two evolute spirals, with thick lateral laminae, one on each side of the equatorial plane. The thickened shell margin is produced by a marginal canal sys- tem, as in Pellatispira (below), that may be overgrown by supplemental chamberlets on the lateral flanks. The equatorial plane is covered by pillars with pores or canals. The later stage may become annular, with narrow marginal interlamellar cavities, formed by mostly imperforate walls suspended on radial spikes (Hottinger et  al., 2001). Middle to Late Eocene (Fig. 6.11; Plate 6.13, figs 1-2 ). • Pellatispira Boussac, 1906 (Type species: Pellatispira douvillei Boussac, 1906). The test is a flattened lenticular to discoidal, evolute planispiral. Double septa enclose the intraseptal canals. A supplemental skeleton forms a thick canaliculated crest, with densely grouped parallel, radial canals, extending in the equatorial direction (see Hottinger et al., 2001). Walls have coarse perforations and thick pillars perpendicu- lar to the surface. Late Middle Eocene to Late Eocene. (Fig. 6.10; Plate 6.5, fig. 7; Plate 6.14, figs 1- 6). • Serraia Matsumaru, 1999 (Type species:  Serraia cataioniensis Matsumaru, 1999). A  pellatispirid test, with secondary and tertiary spiral chambers of intercalary whorls in the early growth stage. Late Middle Eocene. • Vacuolispira Tan Sin Hok, 1936 (Type species: Pellatispira inflata Umbgrove, 1928). A thickly lenticular to globular test, with early evolute, planispiral chambers that support a heavy lateral supplemental skeleton, that is pierced by numerous radial canals. Later the spiral chambers are replaced by concentric arrangements of iso- lated chamberlets or interlamellar cavities, supported by pillars and covered by strong secondary lamination to form thick, perforate walls supported by pillars. The 410 Evolution and Geological Significance of Larger Benthic Foraminifera Spiral main chamber A Proloculus Piles of lamellae C Marginal crest B Lateral chamberlets Protoconch E Spiral main chamber D G Lateral chamberlets Peripheral Radial canals supplemental skeleton H F Spiral main chamber Marginal crest Figure  6.11. Bilamellar Biplanipira mirabilis (Umbgrove) growth A) Sterodiagram of Biplanispira mirabilis (Umbgrobe) (after Umbgrove, 1936). B-H ) Type material, Utrecht University coll. The original description of B. mirabilis is precise except for what he called “tubules = pores, originating from the primary chambers” are in fact the tubular element of the enveloping canal system (Hottinger et al., 2001). B) External lateral view, the spiral main chambers cannot be identified from outside; C- F) Equatorial sections of megalospheric specimens showing the main spiral chambers, the lateral chamberlets/i nterlamellar spaces and the full extension of the marginal crest. G-H ) Axial sections, G) of a lenticular specimen, H) of an almost biplanar discoidal test. Scale bars = 1mm. The Cenozoic Larger Benthic Foraminifera: The Paleogene 411 thickened shell margin produced by a marginal canal system, as in Pellatispira, may be overgrown by supplemental chamberlets on all sides of the shell. (see Hottinger et al., 2001). Late Eocene (Plate 6.14, fig. 7). Family Nummulitidae de Blainville, 1827 The test is planispiral involute or evolute with septal, marginal and vertical canals (Fig. 6.12). Paleocene to Holocene. Subfamily Heterostegininae Galloway, 1933 Planispiral, with a canaliculate marginal cord, and septal canal trabeculae, but with true secondary septa, developed right across the chamber, forming chamberlets. Paleocene to Holocene. • Grzybowskia Bieda, 1950 (Type species:  Grzybowskia Bieda, 1950.). The test is wholly involute with favosely shaped, polygonal chamberlets. Late Eocene (Plate 6.14, fig. 12). • Heterostegina d’Orbigny, 1826, sensu stricto emended Banner and Hodgkinson, 1991. The test is thick, planispiral, involute to evolute with chambers divided by secondary septa to form small chamberlets. No alar prolongations, but with raised sutures. Banner and Hodgkinson (1991) suggested three genera based on coiling: 1. Heterostegina (Heterostegina) (Type species: Heterostegina depressa d’Orbigny, 1826). The test is initially involute, but wholly evolute in mature whorls. Chamberlets are in equatorial view rectangular and restricted to those parts of the chambers which did not embrace earlier ones. Late Eocene to Holocene (Plate 6.14, fig. 11; Plate 6.15, figs 9- 10). Chamber Marginal cord Chamber suture Trabecule Lateral intraseptal canal Figure 6.12. Enlargement of the lateral surface of some chambers of Nummulites showing part of the sep- tal sutures between alar prolongations and their extensions, the trabeculae. 412 Evolution and Geological Significance of Larger Benthic Foraminifera 2. Heterostegina (Vlerkina) = Heterostegina (Vlerkinella) (Type species: Heterostegina (Vlerkina) borneensis). The test is completely involute (at least in the megalospheric forms), with rectangular lateral chamberlets, even in the chambers which embrace the preceding whorl, connected by Y- shaped intercameral stolons. Undivided sutural canals. Late Eocene to Late Miocene (Messinian) (Plate 6.15, figs 11, 14- 16; Plate 6.15, fig. 2). • Spiroclypeus Douvillé, 1905 (Type species: Spiroclypeus orbitoideus Douvillé, 1905). The test is planispiral, involute, with numerous narrow chambers increasing rapidly in height, divided into alternating chamberlets and tiers of lateral chamberlets on either side. Adjacent chamberlets of the same primary chamber, and adjacent cham- berlets of successive chambers communicate with pore- like apertures. There are no alar prolongations. Late Eocene (Letter Stage “Tb”) to Early Miocene (Letter Stage “Te5”) (see below in section 6.3 for discussion of Letter Stages) (Plate 6.15, figs 6, 7, 11- 13). • Tansinhokella Banner and Hodgkinson, 1991 (Type species:  Tansinhokella tatauensis Banner and Hodgkinson, 1991). The test is planispiral, involute, with a marginal cord, and bilamellar and canaliculated septa. It possesses groups of embracing alar prolongations, which are divided into chamberlets when seen in axial section, however, unlike Spiroclypeus, it lacks cubiculae (according to Banner and Hodgkinson, 1991 cubiculae are intralaminar chamberlets). Late Eocene (Letter Stage “Tb”) to Early Miocene (Letter Stage “Te5”) (Plate 6.15, figs 3- 5; Plate 6.16, fig. 13). Subfamily Nummulitinae de Blainville, 1827 The test is planispiral involute or evolute, with a canaliculate marginal cord, and septal canal trabeculae but without secondary septa forming chamberlets. They may become annular in more advanced forms. Late Cretaceous to Holocene. • Assilina d’Orbigny, 1839 (Type species:  Assilina depressa d’Orbigny, 1850 = Nummulites spira de Roissy, 1805 = Operculina (Assilina) Schaub. 1981). The test is tightly coiled, flat to biumbilicate, evolute with numerous chambers per whorl. The rate of whorl growth is slow resulting in a test without alar prolongations. The marginal cord is thick with a coarse canal system. Septa are radial, and trabeculae are absent. Late Paleocene (planktonic zone P4) to Middle Eocene (planktonic zone P14) (Plate 6.5, figs 11-1 2; Plate 6.14, figs 8-1 0; Plate 6.17, figs 2-8 ; Plate 6.18, figs 1-4 ; Plate 6.19, fig. 9; Plate 6.20, figs 6, 9- 11, 14). • Chordoperculinoides Arni, 1965 (Type species: Oerculina bermudezi Palmer, 1934). The test is characterised by coarse vertical canals and a massive marginal cord. Paleocene (Selandian to Thanetian) (Plate 6.20, Figs 15- 16) • Nummulites Lamarck, 1801 (Type species: Camerina laevigata Bruguière, 1792). The test is lenticular, planispiral, involute, and tightly coiled with numerous whorls. The chambers are simple with a distinct marginal cord on the periphery and ramified canals within the sutures, which may be radial, sigmoid, meandrine or reticulate. Trabeculae are present, and there are pronounced alar prolongations. The extinction The Cenozoic Larger Benthic Foraminifera: The Paleogene 413 of Nummulites can be correlated with the Rupelian stage and planktonic zones P18- P21a. Middle Paleocene to Early Oligocene (Plate 6.5, figs 11- 12; Plate 6.12, figs 1- 2; Plate 6.16, fig. 11; Plate 6.18, figs 8,10; Plate 6.19, figs 1- 8; Plate 6.20, figs 1- 5; Plate 6.21, fig. 1- 16). • Operculina d’Orbigny, 1826 (Type species: Lenticulites complanatus Defrance, 1822). The test is planispiral, evolute lenticular to compressed and loosely coiled. The chambers are simple. Trabeculae are absent. Sutural canals are forked or branching. Late Paleocene to Holocene (Plate 6.1, fig. 1; Plate 6.12, fig. 1; Plate 6.18, figs 5-6 , 9, 11- 12). • Operculinella Yabe, 1918 (Type species:  Amphistegina cumingii Carpenter, 1860). The last true Nummulites spp. became extinct at the top of the Td “Letter Stage” with Nummulites fichteli Michelotti 1841 from the upper Early Oligocene of Italy. Contrary to the opinions of S. Cole (in Loeblich and Tappan, 1964) and Loeblich and Tappan (1988), Nummulites can be distinguished from Operculinella. Eames et al. (1962) illustrated a simple Nummulites vascus Joly and Leymerie (plate 1, fig- ures A, B) to compare with Operculinella cumingii (Carpenter) (Palaeonummulites nomen oblitum). The strong dimorphism seen between microspheric and megalo- spheric forms of Oligocene specimens of Nummulites is never seen in Operculinella (where the microspheric and megalospheric generations are externally identical). The presence of trabeculae in Nummulites and their absence from Operculinella is note- worthy, but, most importantly, the diameter of the megalospheric protoconch of Nummulites (in both simple and complex forms) is much greater than the diameter of the proloculus of Operculinella. The megalospheric loosely coiled Operculinella (e.g. Operculinella cumingii) persists to the Holocene but the large protoconch of true Nummulites does not occur beyond the Early Oligocene. Oligocene to Holocene (See Chapter 7). • Palaeonummulites Schubert, 1908 (Type species: Nummulina pristina Brady, 1874). Palaeonummulites are here attributed to all involute forms, but having a tight spire and lacking the developed, highly extended later chambers of Operculinella (see Haynes et al. 2010; BouDagher-F adel and Price, 2014). Operculinoides, the American genus (type species of Nummulites willcoxi Heilprin, 1883; see BouDagher- Fadel and Price, 2014)  is an involute, tightly coiled simple nummulitid, and is a synonym of Palaeonummulites. (Plate 6.19, figs 7- 8; Plate 6.28, figs 18- 19). • Planocamerinoides Cole, 1958 (Type species:  Nummularia exponens de Sowerby, 1870). Biumbilicate test with multilamellar thickenings over the umbonal area. Late Paleocene to Middle Eocene (Plate 6.20, fig. 14) • Planostegina Banner and Hodgkinson, 1991 (Type species:  Heterostegina oper- culinoides Hofker, 1927). A  totally evolute, laterally more compressed form of Heterostegina, with chambers divided by septula into complete or incomplete sub- rectangular chamberlets connected by Y-s haped intercameral stolons. The test has strong ornamentation and undivided sutural canals. Late Paleocene to Holocene (Plate 6.15, fig. 1; Plate 6.18, fig. 7). • Ranikothalia Caudri, 1944 (Type species:  Nummulites nuttalli Davies, 1927). The test is lenticular with alar prolongations, initially involute, becoming evolute in the last whorls. There is a thick marginal cord with a coarse canal system that connects 414 Evolution and Geological Significance of Larger Benthic Foraminifera to simple vertical septal canals. Trabeculae are present. Late Paleocene (Plate 6.13, fig. 3; Plate 6.22, figs 1- 5). Family Cycloclypeidae Galloway, 1933 emend. BouDagher- Fadel, 2002 This family is distinguished by the development of concentric annular, wholly evolute chambers, each chamber being divided into numerous chamberlets in a median plane, and each chamberlet separated from adjacent chamberlets by straight, canaliculated walls. There is no marginal cord, except in the early stages of the microspheric genera- tion. Eocene to Holocene. • Cycloclypeus Carpenter, 1856 (Type species: Cycloclypeus carpenteri Brady, 1881). A  nummulitid with a nepionic morphology like Heterostegina, but with a final growth stage with cyclic chambers. No alar prolongations occur. Early Oligocene to Holocene (Plate 6.14, fig. 4; Plate 6.17, fig. 1, see Chapter 7). Family Orthophragminidae Vedekind, 1937 The orthophragmines have a Cycloclypeus- like or operculinid microspheric juvenile form. Middle Paleocene to Eocene. Subfamily Discocyclininae Galloway, 1928 Megalospheric forms have a subspherical protoconch enclosed by a larger reniform deuteroconch. Microspheric forms have an initial spiral of small chambers, and later stages with cyclical chambers subdivided by septula into small rectangular cham- berlets connected by annular and radial stolons. There is a fine equatorial layer and small lateral chamberlets. A small, intraseptal and intramural canal system is present (Fig. 6.13). Middle Paleocene to Late Eocene. • Actinocyclina Gümbel, 1870 (Type species:  Orbitolites radians d’Archiac, 1850). This form differs from Discocyclina (below) in having distinct rays formed by a pro- liferation of broad and low lateral chamberlets. Middle to Late Eocene (Plate 6.22, fig. 6). • Asterophragmina Rao, 1942. (Type species:  Pseudophragmina (Astemphragmina) pagoda Rao, 1942). Stellate in outline, with rays radiating from the centre. Radial walls are absent after five to ten annuli of cyclic chambers, which remain undivided. Late Eocene (Plate 6.22, fig. 7). • Athecocyclina Vaughan and Cole, 1940 (Type species:  Pseudophragmina (Athecocyclina) cookei Vaughan and Cole, in Cushman, 1940). Athecocyclina has no incipient septula. Eocene • Discocyclina Gümbel, 1870 (Type species:  Pseudophragmina (Astemphragmina) pagoda Rao, 1942). The test is discoidal, flat, with an equatorial layer composed of concentric rings of rectangular chamberlets, those of successive cycles alternating in position. Lateral chamberlets are connected with the equatorial layer by verti- cal stolons. Annular stolons occur at the proximal end of the radial walls and con- nect adjacent chamberlets. Middle Paleocene to Late Eocene (Plate 6.5, fig. 12; Plate The Cenozoic Larger Benthic Foraminifera: The Paleogene 415 Pores Eggholders 100μm D P B P D A Corona D Equatorial P chamberlets Lateral chamberlets D Auxiliary C chamberlets Figure 6.13. Features of Discocyclina; A, B) SEM photos of Discocyclina, the internal surface of the lateral chamber wall bears egg- holders which might have harbored symbionts; C) Equatorial section; D) Axial sec- tion of Discocyclina sp., France, UCL coll. P= protoconch; D = Deuteroconch 6.12, fig. 1; Plate 6.16, figs 3, 12; Plate 6.22, figs 5, 6, 8-1 5; Plate 6.23, figs 1-2 , 4-1 1; Plate 6.24, figs 1, 6- 8, 11; Plate 6.25, Fig. 7). • Hexagonocyclina Caudri, 1944 (Type species:  Orbitoclypeus? cristensis Vaughan, 1924). The test is similar to Discocyclina but with two symmetrical auxiliary cham- bers on each side of the nucleoconch and four spirals, and predominantly hexagonal equatorial chambers. Early to Middle Eocene (Plate 6.24, figs 2- 5). • Nemkovella Less, 1987 (Type species:  Orbitoides strophiolata Gümbel, 1870). The test has a circular outline, with ribs constructed (as in Actinocyclina) exlusively from lateral layers. It differs from Discocyclina in lacking annular stolons. Late Paleocene (Thanetian) to Early Eocene (Ypresian) (Plate 6.16, fig. 9). • Proporocyclina Vaughan and Cole, 1940 (Type species:  Discocyclina perpusilla Vaughan, 1929). This form is characterized by the presence of well- developed, radial septula with distal annular connections. Eocene • Pseudophragmina Douvillé, 1923 (Type species: Orthophragmina floridana Cushman, 1917). The test is circular to subquadrate in outline. An eulepidine embryo is sur- rounded by a single ring of large nepionic chambers, which are followed by smaller equatorial chambers. The septa are irregular. Numerous irregular layers of lat- eral chamberlets occur on both sides of the single equatorial layer. Although, the subgenera P. (Proporocyclina) Vaughan and Cole, in Cushman, 1940 and P. (Athecocyclina) Vaughan and Cole, in Cushman, 1940 are both considered as 416 Evolution and Geological Significance of Larger Benthic Foraminifera synonyms to Pseudophragmina by Loeblich and Tappan, 1988, they are However, in they are proved to be different genera, as they in fact exhibit important, distinguish- ing, stratigraphically- characteristic, morphological features; thus Athecocyclina has no incipient septula, Pseudophragmina has irregular septula, while Proporocyclina has well- developed, radial septula with distal annular connections (see BouDagher- Fadel and Price, 2017). Eocene (Plate 6.7, figs 10- 12; Plate 6.26, fig. 9). Subfamily Orbitoclypeinae Brönnimann, 1946 Microspheric tests have an early planispiral coil, while megalospheric tests have a globular protoconch, enclosed by a larger reniform deuteroconch. Members of this subfamily may occur with or without ribs. There is a single equatorial layer of cham- berlets and several layers of small lateral chamberlets, and cyclical chambers are not subdivided into chamberlets. Chambers have four stolons. Middle Paleocene to Late Eocene. • Asterocyclina Gümbel, 1870 (Type species: Calcarina? stellata d’Archiac, 1889). The test is stellate with five to six rays which are caused by thickening and out- ward growth of the median layer, where normally it is thin, as in Discocyclina. Lateral chambers, made of axial subdivision of annuli, are present on both sides of median layer. Middle Paleocene to Eocene (Plate 6.23, figs 1-2 ; Plate 6.24, fig. 9- 11). • Neodiscocyclina Caudri, 1972 (Type species: Discocyclina anconensis Barker, 1932). The test is flat and lenticular. Equatorial chambers are irregular, enlarging from the proloculus to the periphery, and radial walls are thin. Up to twenty layers of lateral chambers occur on both sides of the equatorial layer. Numerous pillars are visible in vertical section. Middle Eocene (Plate 6.27, fig. 1). • Orbitoclypeus Silvestri, 1907 (Type species: Orbitoclypeus himerensis Silvestri, 1907). The test is inflated centrally. The megalospheric embryo has a deuteroconch that completely encloses the protoconch. Complete cycles of periembryonic chambers surround the embryo. Equatorial chambers are spatulate, and in concentric rings, without stolons connecting adjacent chamberlets. Ribs may be present, and con- structed from lateral layers, but with a slight axial enlargement. Late Paleocene to Late Eocene (Plate 6.27, fig. 2). • Stenocyclina Caudri, 1972 (Type species: Orthophragmina advena Cushman, 1921). The test is circular in outline. The embryo is followed by small square equatorial chamberlets. Radial walls are aligned in successive annuli. Numerous layers of low lateral chamberlets occur on both sides of a thin, but well developed, equatorial layer. Middle Eocene. Superfamily NONIONOIDEA Schultze, 1854 Unlike in BouDagher-F adel (2008), the miscellaneids are separated from the num- mulitoids in this book and considered as belonging to the superfamily Nonionoidea because of the planispiral-i nvolute coiling combined with an interiomarginal position of the foramina (see Hottinger, 2009). The Cenozoic Larger Benthic Foraminifera: The Paleogene 417 Family Miscellaneidae Sigal in Piveteau, 1952 Members of this family have a planispiral, evolute test. The aperture is single or multi- ple, interiomarginal and symmetrical to the equatorial plane of the test (see Hottinger, 2009). Paleocene and Earliest Eocene. Subfamily Miscellaneinae Kacharava in Rauzer- Chernoussova and Furzenko, 1959 Members of this subfamily have a single, interiomarginal, intercameral foramen that is symmetrical with respect to the equatorial plane of the shell (see Fig. 6.14). • Miscellanea Pfender, 1935 (Type species: Nummulites miscella d’Archiac and Haime, 1853). A lenticular angular test, having few whorls, is strongly ornamented with very coarse perforation, raised rounded pustules, pillars, raised ridges and raised bars, Umbilical radial canal Pillars Foramen A Spiral canal Marginal cord Foramen Alar prolongation B C Embryonic wall Distal end of intraseptal canal system spiral sheet and supplementary skeleton together Figure 6.14. Illustrations of the differences between Nummulites: A) N. masiraensis Carter, Masira Island, Oman; B) N. cf. irregularis Deshayes, Kenya; C) Miscellanea miscella (d’Archiac and Haime), Paleocene, Indonesia. The straight wall between protoconch and deuteroconch indicates the biconch quality of the nepiont. UCL coll. Scale bars = 1mm. 418 Evolution and Geological Significance of Larger Benthic Foraminifera dividing the space between longitudinal ridges. Bilamellar septa are quite distinct. The embryonic apparatus is composed of a protoconch and a slightly smaller deu- teroconch, separated by a straight wall and surrounded by a common wall with canals forming a biconch. The biconch is followed by a third chamber with much smaller volume. Paleocene (Plate 6.13, figs 4-1 2). Subfamily Miscellanitinae Hottinger, 2009 Members of this subfamily have a planispiral, involute test with multiple intercam- eral foramina. Their apertures are a single interiomarginal row in a comparatively low chamber. Late Paleocene (Thanetian). • Bolkarina Sirel, 1981. (Type species: Bolkarina aksarayi Sirel, 1981) The test is large and discoidal. The proloculus is small, and followed by numerous rectangular peri- embryonic chambers. Septa are doubled with intraseptal canals. Late Paleocene (Thanetian) (Fig. 6.15). • Miscellanites Hottinger, 2009 (Type species: Miscellanea iranica Rahaghi, 1983). The test is globular, planispiral and involute with low but elongate chambers reaching from pole to pole. The outer surface of the chamber wall is covered by a simple and shallow enveloping canal system with pustules. Late Paleocene (Thanetian). Figure. 6.15 Schematic figure from Hottinger (2009) illustrating the structure of Bolkarina aksarayi Sirel, 1981. af: apertural face; apa: annular passage (in pre- septal position); expch: expanse chamber; f: foramen; puch: penultimate (expanse) chamber; rpa: radial passage; s: septum; sut: (chamber) suture; tpa: transverse tubular passage. Arrow: direction of growth. (not to scale). The Cenozoic Larger Benthic Foraminifera: The Paleogene 419 Superfamily PLANORBULINOIDEA Schwager, 1877 The test is trochospiral in the early stages, but later may be uncoiled and rectilin- ear, or biserial, or with many chambers in the whorl. They are found with intra- to extra- umbilical apertures, and additional equatorial apertures may be present. Early Cretaceous (Berriasian) to Holocene Family Eoannularidae Ferrández- Cañadell and Serra- Kiel, 1998 The test is bilamellar perforate, with a bilocular megalospheric embryo followed first by orbitoidal chamberlets, which change later to cyclical chambers. The cyclical cham- bers are subdivided into rectangular chamberlets, or are not subdivided. Microspheric initial chambers are arranged in a peneroplid- like spire that changes into annular chambers by a progressive increase in chamber width. Middle Eocene. • Eoannularia Cole and Bermudez, 1944 (Type species: Eoannularia eocenica Cole and Bermúdez, 1944). The test is discoidal, flat with subdivided annular chambers and lamellar thickening only in the early stage. The megalospheric form has a proto- chonch that is completely enclosed by the deuteroconch, and chambers following the bilocular embryo occurring as an annular series in a single layer. Middle Eocene (Plate 6.27, fig. 3; Fig. 6.16). • Epiannularia Caudri, 1974 (Type species:  Epiannularia pollonaisae Caudri, 1974). The test is discoidal, centrally thin and thickened towards the periphery. The bilocu- lar embryont is followed by cyclical divided chambers, later becoming undivided. Middle Eocene (Fig. 6.16). LINDERINIDAE EOANNULARIDAE Tetralocular embryonic apparutus Bilocular embryonic apparutus Orbitoidal growth Orbitoidal growth Divided cyclic chambers into Eoannularia arcuate alternating Middle Eocene chamberlets Thickening Epiannularia on both sides of the test Middle Eocene Thickinings splitting into lateral chambers Linderina Undivided cyclic chambersCaudriella Middle -Late Eocene Middle Eocene Figure 6.16. Schematic figures highlighting the differences between the Linderinidae and the Eoannularidae. Drawings not to scale. 420 Evolution and Geological Significance of Larger Benthic Foraminifera Family Linderinidae Loeblich and Tappan, 1974, emended Ferrández- Cañadell and Serra- Kiel, 1998 The test is bilamellar with a lobate outline. The megalospheric embryo has a quadril- ocular early stage, with three initial chambers separated by flattened walls, followed by a fourth arcuate chamber with apertures at both sides. Later chambers occur in orbitoi- dal growth, with crosswise- oblique stolons. Forms may have varying amounts of calcite deposited on both sides of the central embryont, but no lateral chambers. Microspheric intial chambers are arranged in Planorbulina- like spire. Middle to Late Eocene. • Caudriella Haman and Huddleston, 1984 (Type species: Margaritella ospinae Caudri, 1974). The test is lenticular with lateral chamberlets, a small proloculus followed by two larger chambers and an arcuate median chamber, with lateral, irregularly arranged chambers with thick walls and open lumina. Caudriella seems to differ from Linderina (below) by having well- developed lateral chamberlets. Caudri (1974) discussed the systematic position of her new genus. After considering and discard- ing the relationship with Linderina and with Pseudolepidina (see below), she decided to leave it “standing alone and isolated from all other groups and evolutionary lin- eages”. Loeblich and Tappan (1987) included Caudriella into the Lepidocyclininae (together with Astrolepidina, Eulepidina, Lepidocyclina and Pseudolepidina, all of which having a bilocular megalospheric embryo). Ferrández-C añadell and Serra- Kiel (1998) in their revision of Linderina decided to put the genus Caudriella within their emended Linderina on the basis of the presence of the embryonic apparatus, which is similar to that of Linderina. To date, only one species, Caudriella ospinae (Caudri, 1974), from the Middle Eocene of Margarita Island (Venezuela), is known, and more work is needed on this genus. Middle Eocene (Fig. 6.16). • Linderina Schlumberger, 1893 (Type species: Linderina brugesi Schlumberger, 1893). The test is large, discoid but without lateral chamberlets, and with considerable thick- ening in the early stage on both sides of the test, formed by the superposition of the successive involute outer lamellae. Orbitoidal chambers, consisting of small arched chamberlets, occur in concentric series, with successive layers alternating in position. There are no annular stolons, nor stolons in the distal wall of chamberlets. The aper- tures correspond to the stolon system, with rows of apertures between chamberlets in the last chamber, from which the orbitoidal growth starts. Microspheric initial chambers are arranged in a Planorbulina-l ike spire. Middle to Late Eocene (Plate 6.27, figs 4-6 ; Fig. 6.16). Family Planorbulinidae Schwager, 1877 The test is free or attached, with an early stage that is trochospiral, later becoming discoid, cylindrical or conical. The aperture is single or multiple. Eocene to Holocene. • Neoplanorbulinella Matsumaru, 1976 (Type species: Neoplanorbulinella saipanensis Matsumaru, 1976). The test is attached, conical to concavo- convex. The microspheric proloculus is small, while the megalospheric protoconch consists of a spherical pro- loculus enclosing a reniform second chamber, followed by a third chamber with only a proximal stolon and numerous trochospirally arranged chambers. The adult form The Cenozoic Larger Benthic Foraminifera: The Paleogene 421 has equatorial chambers that occur in an annular series and with two or more lat- eral chambers on the concave side. Multiple apertures are present. Late Eocene tarly Miocene. • Peelella Matsumaru, 1996 (Type species: Peelella boninensis Matsumaru, 1996). The test is small, concavo- convex. The megalospheric apparatus has a globular protoconch and a reniform deuteroconch, surrounded by a thick wall followed by a trochospiral to planispiral symmetrical later stage. The microspheric gener- ation has a small proloculus, followed by a trochospiral stage. Equatorial cham- bers form an annular series, with successive series alternating in position. Lateral chambers are irregular, curved, and well differentiated from the equatorial layer through stolons and coarse perforations on the ventral side of the test. Late Oligocene. • Planolinderina Freudenthal, 1969 (Type species:  Planolinderina escornebovensis Freudenthal, 1969). The test is attached, single- layered. Microspheric forms have a trochospiral early stage; megalospheric forms occur with a protochonch and deu- teroconch. Later chambers are added in a cyclic series. The aperture is a series of basal openings. Late Oligocene to Early Miocene (Burdigalian). • Planorbulina d’Orbigny 1826. (Type species: Planorbulina mediterranensis d’Orbigny, 1826). The test is attached, single- layered with a long initial spiral and an irregular pattern of a small number of orbitoidal chambers, with supplementary apertures in sutural positions. The protoconch and the two following chambers forming a tri- conch. The aperture is made of single slits, bordered with a narrow lip. Eocene to Holocene. • Planorbulinella Cushman, 1927 (Type species: Planorbulina vulgaris d’Orbigny var. larvata Parker and Jones, 1865). The test is attached, discoidal, with alternating chambers built over the apertures of the previous ring. They may have scattered pillars dorsally and ventrally (e.g. P. larvata, Plate 7.10, Fig.  10) or heavy lateral thickening (e.g. P. solida, Plate 7.11, Figs 2-5 ). Chamber walls are coarsely perforate. Relapse chambers are present. The embryont has three chambers with thick walls. The aperture is made of single slits, bordered with a narrow lip. Eocene to Holocene (Plate 6.27, fig. 7). • Tayamaia Hanzawa, 1967 (Type species:  Gypsina marianensis Hanzawa, 1957). A dome- like test with the ventral hollow, filled with numerous irregular chamberlets. Only very few chambers cover the dorsal side of the dome. The median layer has a globular protoconch and arcuate chambers, interconnecting by four stolons. Upper Late Oligocene (Te) to Early Miocene. Family Cymbaloporidae Cushman, 1927 Chambers occur in a single layer (Fig.  6.17). Late Cretaceous (Cenomanian) to Holocene. • Cymbalopora von Hagenow, 1851 (Type species: Cymbalopora radiata von Hagenow, 1851). The test is low and conical, with an open umbilicus, thickened wall and sec- ondary thickening, obscuring sutures and chambers on spiral side. The walls are lamellar. Late Cretaceous to Middle Paleocene (Fig. 6.17). 422 Evolution and Geological Significance of Larger Benthic Foraminifera • Eofabiania Küpper, 1955 (Type species: Eofabiania grahami Kopper, 1955). The test is conical, with a trochospiral early stage. Later chambers have no subdivisions and may be added in annular series. Early to Middle Eocene. • Fabiania Silvestri, 1924 (Type species: Patella (Cymbiola) cassis Oppenheim, 1896). The test is conical, with a deeply excavated centre. The early stage has two globose thick- walled and perforate chambers, later chambers are cyclical with horizontal and vertical partitions. The aperture is a single row of pores opening into the large umbil- icus. Late Paleocene to Late Eocene (Plate 6.27, figs 8- 11; Fig. 6.17). • Gunteria Cushman and Ponton, 1933 (Type species:  Gunteria floridana Cushman and Ponton, 1933). The test is compressed and flabelliform. The early stage has large, globular, undivided chambers, later chambers are concentric and subdivided, with radial and vertical partitions. Middle Eocene (Fig. 6.17). Figure 6.17 The evolution of the conical forms of the Paleogene some Planorbulinoidea (Cymbaloporidae) and Rotalioidea (Chapmaninidae). Drawings not to scale (modified after Deloffre and Hamaoui, 1973). The Cenozoic Larger Benthic Foraminifera: The Paleogene 423 • Halkyardia Heron- Allen and Earland, 1918 (Type species:  Cymbalopora radiata von Hagenow var. minima Liebus. 1911). The test is biconvex, with an embryont consisting of a large protoconch and deuteroconch, and two primary auxiliary chambers. The umbilicus is filled with horizontal bilamellae and connecting pil- lars. Middle Eocene (Lutetian) to Middle Oligocene (Rupelian) (Plate 6.27, fig. 12; Fig. 6.17). Family Victoriellidae Chapman and Crespin, 1930 The test is attached or may be free in the juvenile stage, with a trochospiral early stage, later becoming an irregular mass of chambers. Late Cretaceous (Santonian) to Holocene. Subfamily Carpenteriinae Saidova, 1981 The test is attached, trochospiral throughout, planoconvex with a large aperture, open in the umbilicus. Paleocene to Holocene. • Carpenteria Gray, 1858 (Type species: Carpenteria balaniformis Gray, 1858). The test has a carinate periphery, and is planoconvex with a flat spiral side and distinct rims or keels, a strongly convex, distinctly perforate umbilical side surrounded by thick pillars. Late Eocene to Holocene (See Chapter 7). • Neocarpenteria Cushman and Bermúdez, 1936 (Type species:  Neocarpenteria cubana Cushman and Bermúdez, 1936).The test is a low trochospiral and bi- evolute, with a flattened spiral side and a periphery, broadly carinate, with a keel. Late Eocene. Subfamily Rupertininae Loeblich and Tappan, 1961 The test is attached, with a distinct flattened disk, but coiling grows out away from the site of attachment. Late Cretaceous to Holocene. • Biarritzina Loeblicah and Tappan, 1964 (Type species: Columella carpenteriaeformis Halkyard, 1918). The test is attached by the flaring basal disk, with an early stage that is trochospirally enrolled with a distinctly loose coiling style, but later tending to become uniserial. The walls have fine perforations. The aperture is terminal with a distinctly raised lip, and may be present on one or two chambers of the final whorl. Middle Eocene to Holocene (Plate 6.26, fig. 5; see Chapter 7). Subfamily Victoriellinae Chapman and Crespin, 1930 The juvenile stage may be free living, but later stages are attached. High spired forms develop around a hollow axis, with pillar-l ike thickenings in the walls. The aperture is an umbilical slit bordered by a lip. Middle Eocene to Holocene. • Eorupertia Yabe and Hanzawa, 1925 (Type species: Uhligina boninensis Yabe and Hanzawa, 1922). The test is highly trochospiral, enrolled about an axial hollow. The walls are coarsely perforate with small pillars between the perforations. Middle to Late Eocene (Plate 6.26, figs 2- 4; Plate 6.28, figs 3-4 ). 424 Evolution and Geological Significance of Larger Benthic Foraminifera • Korobkovella Hagn and Ohmert, 1971 (Type species:  Truncatulina grosserugosa Gümbel, 1870). The test is low trochospiral with the spiral side flattened as an attached area. Walls have coarse pits and irregular vermiform channels. Middle Eocene. • Maslinella Glaessner and Wade, 1959 (Type species: Maslinella chapmani Glaessner and Wade, 1959). The test is low trochospiral, and semi-i nvolute with curved sutures and fine pillars surrounding the umbilicus. Late Eocene. • Victoriella Chapman and Crespin, 1930 (Type species: Carpenteria proteiformis Goës var. plecte Chapman, 1921). The test is conical, usually with a free juvenile stage and a small attachment area near the apex, consisting of a few inflated chambers with pillar-l ike thickenings in the wall. In the adult stage the coiling is high-s pired, with three to four subspherical chambers per whorl (not enclosing), either with an umbili- cal depression or arranged round an axial hollow. Septa are trilamellar. Late Eocene to Early Miocene (Plate 6.26, fig. 1). • Wadella Srinivasan, 1966 (Type species:  Carpenteria hamiltonensis Glaessner and Wade, 1959). The test is subconical, with a free living juvenile stage, which later becomes fixed on the substrate. Three to five chambers occur per whorl. The walls are coarsely perforate with pores opening at small mounds. Late Eocene. Superfamily ACERVULINOIDEA Schultze, 1854 The test is trochospiral to discoidal and encrusting, consisting of numerous irregularly formed chambers. Paleocene to Holocene. Family Acervulinidae Schultze, 1854 The test is free or attached. A low trochospiral in the initial early stage is followed by inflated chambers spreading over the substrate in more than one layer, forming an irregular mass, that appears to possess no apertures other than the pore-l ike cribrate openings in their upper, distal surfaces. Paleocene to Holocene. • Discogypsina Silvestri, 1937 (Type species: Discogypsina vesicularis Silvestri, 1937). The test is lenticular and flattened, with an equatorial layer of slightly larger thicker chambers, separating lateral layers of smaller irregularly arranged chambers with successive layers showing no alignment. The walls are coarsely perforate with no sto- lon system. Late Eocene to Holocene (Plate 6.26, figs 6-7 ). • Gypsina Carter, 1877 (Type species:  Polytrema planum Carter, 1876). The test is attached, formed by encrusting polygonal, inflated and closely appressed chambers. Chambers of successive layers alternate in position. Late Oligocene to Holocene (Plate 6.26, fig. 10A). • Orbitogypsina Matsumaru, 1996 (Type species: Orbitogypsina vesivularis Matsumaru, 1996). The test is concavo-c onvex. The megalospheric apparatus has a protoconch and deuteroconch followed by nepionic and equatorial chambers, connected by a stolon system as in the Lepidocyclinidae. Microspheric forms have a small prolocu- lus followed by spreading and regular chambers in a globular or discoidal mass, with mural pores and a cribrate aperture in the upper part of the chambers. Successive layers of equatorial chambers are only aligned at the periphery. Late Eocene to Oligocene. The Cenozoic Larger Benthic Foraminifera: The Paleogene 425 • Protogypsina Matsumaru and Sarma, 2008 (Type species:  Protogypsina indica Matsumaru and Sarma, 2008). The test is spherical. The megalospheric apparutus has a spherical protoconch and a kidney- shaped deuteroconch followed by large ovoid chambers. Paleocene (Plate 6.11, fig. 8). • Solenomeris Douvillé, 1924 (Type species: Solenomeris ogormani Douvillé, 1924). The test is large, composed of numerous branching chamberlet layers with fruticose pro- tuberances. The fruticose branches are free from one another, or laterally coherent to varying degrees. The embryonic apparatus consists of a large proloculus surrounded by a whorl of subspherical chambers. The aperture is multiple. Eocene to Holocene. • Sphaerogypsina Galloway, 1933 (Type species: Ceriopora globulus Reuss, 1848). The test is small, almost spherical, with chambers added in numerous layers, those of successive layers being aligned. Chamber roofs are perforate but the walls are thick and imperforate. Paleocene to Holocene (Plate 6.26, fig. 8). • Wilfordia Adams, 1965 (Type species: Wilfordia sarawakensis Adams, 1965). This form has no true initial spire and a complex embryont with a short nepionic spiral, relatively weak ‘pseudo- pillars’ form in the walls of the lateral chambers, with no massive thickenings, and its chambers are clearly rectangular in sec- tion. It differs from Sphaerogypsina in having few spines and many pseudop- illars (radial thickenings of chamberlet walls), and from Schlumbergerella (see Chapter 7) by having finer spines and fewer chamberlets in each tier and a less complicated embryonic apparatus. Loeblich and Tappan (1988) placed this form in the Acervulinidae because of the apparent absence of a canal system, with the communication based on wall perforation. Late Eocene (Plate 6.9, figs 12-1 3; Plate 6.12, fig. 7). Family Homotrematidae Cushman, 1927 The test is attached with a trochospiral early stage, later chambers grow in a massive branching structure. Eocene to Holocene. • Sporadotrema Hickson, 1911 (Type species:  Polytrema cylindricum Carter, 1880). The test occurs with a planispiral early stage, but later chambers spiral upwards around a core of irregular tubes that open distally on the branches. Early upright spiralling chambers have a terminal aperture that remains a foramen to subsequent chambers, and develop subsequently into a complex network of stolons that in turn open to the exterior at the ends of the chambers. Multiple, fine pores at the inner surface fuse to form a coarse perforation in the outer wall. This form differs from Victoriella (see above) in having less inflated chambers and no pillars or pustules. Eocene to Holocene (See Chapter 7). Superfamily ASTERIGERINOIDEA d’Orbigny, 1839 The test is trochospiral to planispiral, with a closed umbilicus. Chambers occur with internal partitions. Supplementary chamberlets develop around the umbilicus. The aperture is umbilical, and may extend up the apertural face. Late Cretaceous (Santonian) to Holocene. 426 Evolution and Geological Significance of Larger Benthic Foraminifera Family Amphisteginidae Cushman, 1927 Chambers are numerous, with interseptal pillars. The aperture is a narrow slit. Eocene to Holocene. • Amphistegina d’Orbigny, 1826 (Type species: Amphistegina quoyii d’Orbigny, 1826). The test is trochospiral, asymmetrically lenticular, involute with an angular, carinate periphery and lobed sutures. Chambers are strongly curved back at the periphery. Eocene to Holocene (Plate 6.18, fig. 7; Plate 6.26, fig. 10). Family Boreloididae Reiss, 1963 Tests occur with only the later chambers divided into chamberlets. The protoconch is bilocular and is followed by an early trochospiral stage, the later stages are planispiral and involute. Late Paleocene to Middle Eocene. • Boreloides Cole and Bermudez, 1947 (Type species:  Boreloides cubensis Cole and Bermudez, 1947). The test is subspherical with a trochospiral early stage and a biloc- ular embryont. The later stage is annular and planispiral. The spiral wall is thick and pitted. Middle to Late Eocene. • Eoconuloides Cole and Bermudez, 1944 (Type species:  Eoconuloides wellsi Cole and Bermudez, 1944). The test is lenticular with a rectangular axial section, and involute with a bilocular embryonic stage, the final chambers are subdivided into chamberlets on the umbilical side, pillars are present over the spiral side with a thick wall. The bases of the ventral septa have multiple stolons. This genus is dis- tinguished from Amphistegina (see Hanzawa, 1957, pp. 60-6 1, pl.6, fig. 11; Vaughan and Cole 1941, p.77, pl.45, fig.3) in having counter-s epta. A  counter- septum is, according to Hottinger (2006), “a kind of lower lip of an interiomarginal- basal aperture appearing in appropriate sections as a forward directed hook below the foramen and glued to the previous shell whorl”. Late Paleocene to Eocene (see Chapter 6). Family Lepidocyclinidae Scheffen, 1932 The test is discoidal, involute, and biconvex with a broad centrum, which grades into a narrow flange. Adauxiliary chambers may be present. The primary spire per- sists into the equatorial layer, or with annular rings of chamberlets that follow the embryont immediately. Stacks of “lateral chamberlets” (cubiculae) occur on each side of the median chamberlets. Pillars may be present between adjacent vertical stacks of cubiculae or scattered in the central region. The chamber walls are per- forated by stolons, but there is no canal system. Middle Eocene to Late Miocene (Early Pliocene?). Subfamily Helicolepidininae Tan, 1936 Members of this subfamily have tests in which the spiral arrangements completely sur- round the bilocular embryo, which is surrounded by a thickened wall, and is lacking adauxiliary chambers. Middle Eocene to Middle Miocene (Serravallian). The Cenozoic Larger Benthic Foraminifera: The Paleogene 427 • Eulinderina Barker and Grimsdale, 1936 (Type species:  Planorbulina (Planorbulinella) guayabalensis Nuttall, 1930). The test is lenticular, with an eoco- nuloid early stage, followed by a trochoid coil with a thick wall and many rows of arcuate median chambers that are connected by stolons. Counter-s epta, as in Eoconuloides (see above), are present. Middle Eocene (Plate 6.26, figs 11- 12; Plate 6.29, figs 2- 4). • Helicolepidina Tobler, 1922 (Type species:  Lepidocyclina (Helicolepidina) spiralis Tobler, 1922). The test is lenticular, with a small eoconuloid early stage followed by a loose planispiral coil (the helicolepidine string) and a series of large imbricate cham- bers outside the helicolepidine string. The median layer has small arcuate cham- bers connected by single or double apertures. Lateral chambers are well developed. Middle to Late Eocene (Plate 6.29, fig. 5). • Helicostegina Barker and Grimsdale, 1936 (Type species:  Helicostegina dimor- pha Barker and Grimsdale, 1936). The test is lenticular, with an eoconuloid early stage which constitutes the larger part of the test. The chambers in the last stage are subdivided into subsidiary chamberlets which consist of two or three rows of arcuate chamberlets growing around the last eoconuloid whorl. The counter- septa of Eoconuloides, which forms hooks below the foramen, is said to develop into “complete septal walls” in Helicostegina (see Hottinger, 2006). Helicostegina differs from Eulinderina in the increase of the size and number of whorls of the eoconuloid stage and in the decrease of the number of rows in the median cham- bers. Eocene (Plate 6.29, figs 6- 11). • Helicosteginopsis Caudri, 1975 (Type species: Helicostegina soldadensis Grimsdale, 1941). The test is lenticular, with an early stage eoconuloid and later numerous arcu- ate chamberlets, forming two or three rows just beneath the outer wall of the whorls. No lateral chamberlets are developed, nor are there any counter- septa. Late Eocene (Plate 6.29, figs 14- 16). • Polylepidina Vaughan, 1924 (Type species: Lepidocyclina (Polylepidina) chiapas- ensis Vaughan, 1924). The embryo consists of a protoconch and deuteroconch surrounded by a thick wall. The equatorial layer is formed by chambers arranged in two or more embryonic spires, followed by a cyclical phase. All chambers have one basal aperture from which the next chamber is formed. Later chambers have a second, or retrovert aperture. All chambers with two apertures then produce two new chambers, which eventually gives rise to cyclical growth. The equatorial chambers are arcuate with only radial stolons. Late Middle Eocene (Plate 6.23, fig. 3). Subfamily Lepidocyclininae Scheffen, 1932 Representatives of this subfamily have a bilocular or multilocular embryonal stage, surrounded by a thickened wall and adauxiliary chambers. Microspheric tests have an early planispiral coil, while megalospheric tests have a globular protoconch, enclosed or followed by a larger reniform deuteroconch. Post- embryonic chambers evolve from cyclical, arcuate to hexagonal in shape, usually with two or more apertures. The lateral chambers are well differentiated from the equatorial layer and in the advanced forms they are arranged in tiers on either side of the equatorial layer. Surface ornaments 428 Evolution and Geological Significance of Larger Benthic Foraminifera and development of pillars seem to be of specific importance. Middle Eocene to Late Miocene (Early Pliocene?), see BouDagher- Fadel and Price (2010a). • Astrolepidina Loeblich and Tappan, 1988 (Type species:  Lepidocyclina asterodisca Nuttall, 1932). Stellate in outline, with four broad arms. The protoconch is almost equal to the deuteroconch and separated by a straight wall. Equatorial layers have ogival early chambers, that later become hexagonal. The equatorial layer increases in thickness towards periphery. Oligocene. • Eulepidina Douvillé, 1911 (Type species: Orbitoides dilatata Michelotti 1861). The test is discoidal and biconvex, and can be extremely large with very broad lateral flanges. In the megalospheric form the small protoconch is completely enclosed by the larger deuteroconch, both are surrounded by a thick wall with many stolons con- necting the embryont to the first series of auxiliary chamberlets. It has two auxiliary chambers and numerous adauxiliary chambers. In the microspheric form, the pro- toconch and deuteroconch are very small. The median layer is very thick and with multiple stolons. Oligocene (Rupelian, P18 in America, P19 in Tethys) to Miocene (middle Burdigalian) (Fig. 6.18; Plate 6.16, figs 7-8 ; Plate 6.18, fig. 7; Plate 6.25, figs 8, 10; Plate 6.26, figs 1, 10B; Plate 6.29, figs 17-1 9). • Lepidocyclina sp. sensu lato Gümbel, 1870 emend. BouDagher-F adel and Banner, 1997. The nomenclatural revision of Lepidocyclina by BouDagher- Fadel and Banner (1997) makes the genus group name Lepidocyclina sensu lato available for the generic naming of microspheric forms. Megalospheric forms, however, are divisible into the typical Paleogene Lepidocyclina (Lepidocyclina) and L. (Nephroleidina) which is essentially Miocene in the Far East. The test is microspheric, biconvex with a periph- eral flange. In vertical section, there are numerous columns of “lateral chambers” (cubiculae) with cubicular lumina surrounding a thin median layer. In equatorial or oblique sections, the cubiculae are arcuate, hexagonal or polygonal, each with two apertures. The roof and floors of the chamberlets are perforate. It possesses a complete periembryonic ring of primary auxiliary and interauxiliary chambers. Middle Eocene (P10) to Early Miocene (N7) in America; Early/L ate Oligocene to late Miocene (Early Pliocene?) in Tethys (see BouDagher- Fadel and Price, 2010a). (Plate 6.15, fig. 6). • Lepidocyclina (Lepidocyclina) Gümbel, 1870 emend. BouDagher-F adel and Banner, 1997 (Type species: Nummulites mantelli Morton, 1833). The test is “isolepidine”, having a protoconch and deuteroconch of nearly equal size, separated by a straight wall. It has two auxiliary chambers and no adauxiliary chambers. The periembryonic chambers are arranged in four spirals. Butterlin (1981) reintroduced the subgenus Neolepidina (Bronnimann), to forms with larger protoconch than deuteroconch as Neolepidina. However, this genus is not valid as the Late Eocene species assigned by Butterlin to Neolepidina do not always have a larger protoconch than deuteroconch. Middle Eocene (Lutetian, P10) to Early Miocene (Burdigalian, N7) in America, Oligocene in Tethys (P18– P22) (Plate 6.16, fig. 1; Plate 6.25, figs 1- 6; Fig. 6.44). • Lepidocyclina (Nephrolepidina) Douvillé, 1911 (Type species: Nummulites marginata Michelotti, 1841). The megalospheric test is biconvex, with an embryonic appa- ratus consisting of a small proloculus followed by a much larger, reniform deutero- conch. The latter forms have quadrate proloculi in the later stages of many lineages (see Chapter 7). The two chambers are separated by a thin imperforate wall with The Cenozoic Larger Benthic Foraminifera: The Paleogene 429 ac D P A Hexagonal chambers B Spatulate chambers D P C Stolons D Figure 6.18. A) Eulepidina ephippioides (Jones and Chapman), Rupelian (early Oligocene), Greece, SFN coll, 2001z0150/ 0011, showing the completely enclosed protoconch (P)  within the deuteroconch (D)  and the numerous small adauxiliary chambers (ac); B- D) Eulepidina dilatata (Michelotti), with typical embryo- nal development with thick- walled D almost fully embracing the thin- walled P, Chattian (late Oligocene), France, UCL coll. Scale bars = Figs A, D = 1mm; B-C  = 0.5mm. a central foramen, and surrounded by a common thick tabulated wall. The equa- torial layer of chamberlets has a basal stolon, that is arcuate in older species, but pointed or hexagonal in younger forms. Lateral chamberlets form on each side of the median layer. Middle Eocene (Lutetian, P10) to Early Miocene (Burdigalian, N7 in America), Late Oligocene in the Mediterranean (early Chattian, P21b) and Indo- Pacific (late Chattian, P22) to Late Miocene (Early Pliocene?) in the Tethyan province. (Plate 6.16, fig. 4; Plate 6.25, fig. 9; Plate 6.29, figs. 12–1 3; Fig. 6.45, see Chapter 7). 430 Evolution and Geological Significance of Larger Benthic Foraminifera • Pseudolepidina Barker and Grimsdale, 1937 (Type species: Pseudolepidina trimera Barker and Grimsdale, 1937). A  lenticular test having a protoconch and deutero- conch that are almost equal in size. The median layer consists of irregularly arcuate chambers, doubled beyond the protoconch and communicating with lateral cham- berlets by stolons. Middle Eocene (Plate 6.16, Fig. 2). Superfamily ORBITOIDOIDEA Schwager, 1876 The test is discoidal to lenticular with prominent dimorphism, as in most orbitoidal species both megalospheric and microspheric generations are found. Microspheric specimens have a distinctly small protoconch (usually about 20 microns), while megalo- spheric forms have a distinctive embryonic stage, enclosed in a thicker wall. Equatorial and lateral chambers may be differentiated or indistinguishable. Late Cretaceous (Santonian) to Oligocene. Family Lepidorbitoididae Vaughan, 1933 The Lepidorbitoididae differ from the Orbitoididae (see Chapter 5) by the different character of the embryonic apparatus and by the form of the median chambers. An embryonic stage with two chambers is followed by hexagonal or arcuate equatorial chamberlets and by differentiated lateral chambers. Late Cretaceous (Santonian) to Oligocene. • Actinosiphon Vaughan, 1929 (Type species: Actinosiphon semmesi Vaughan, 1929). The test is lenticular with well- developed polygonal to hexagonal equatorial and lateral chambers. The embryont consists of a large subspherical protoconch and a smaller reniform deuteroconch, and is followed by a spire of about eleven chambers. Equatorial chambers of the same cycle communicate via median stolons, while lat- eral chambers communicate through pores. Pillars are present forming surface papil- lae. Late Paleocene (Plate 6.17, fig. 10). • Neosivasella Meriç and Çoruh, 1998 (Type species: Neosivasella sungurlui Meriç and Çoruh, 1998). The test is conical. In the megalospheric embryo the protoconch is partly enveloped by the deuteroconch. Equatorial chambers are arcuate in shape, interconnected with stolons. Late Paleocene. • Orbitosiphon Rao, 1940 (Type species: Lepidocyclina (polylepidnal) punjabensis Davies, 1937). The arrangement of embryonic and periembryonic chambers is similar to the Cretaceous genus Orbitoides (see Chapter  5) in having epi- auxiliary chambers, but it lacks the thick embryonic wall. Paleocene (Plate 6.17, fig. 9). • Sirelella Özgen- Erdem, 2002 (Type species:  Sirelella safranboluensis Özgen- Erdem, 2002). A  trochospiral early orbitoidal test having a primitive stolon system (van Gorsel, 1978, pl.6, fig. 5a) and an umbilical plug without vertical canals. The early orbitoidal stage is similar to Orbitokathina (see Chapter  5) and Neosivasella; in Orbitokathina, the umbilical plug is pierced by numerous vertical canals, while Neosivasella possesses lateral chambers. Middle Eocene (Lutetian). The Cenozoic Larger Benthic Foraminifera: The Paleogene 431 Family Orduellinidae Sirel, 1999 The test is free, dimorphic, spherical with sub-r ectangular chambers arranged in mul- tiple spirals in the early part of the test. Later chambers are arcuate and connected by stolons. They are added in concentric series in an orbitoidal manner. The aperture is a single, simple, basal slit in the early stage. Paleocene. • Orduella Sirel, 1999 (Type species: Orduella sphaerica Sirel, 1999). The test is glob- ular, with a large megalospheric proloculus enclosed by a thick wall, followed by sub- rectangular large chambers. The microspheric tests have a small proloculus with numerous small chambers. Late Paleocene (Thanetian). Superfamily ROTALIOIDEA Ehrenberg, 1839 The test is involute to evolute, initially trochospiral or planispiral, commonly with many chambers in numerous whorls. As new chambers are added septal flaps attach to the previous apertural face and enclose radial canals, fissures, umbilical cavities, and intraseptal and subsutural canals. The wall is made of perforate hyaline calcite, and is generally optically radial in structure. Primary apertures occur singly or as multiples. Small opening into the canal system may occur along the sutures. Late Cretaceous (Coniacian) to Holocene. Family Rotaliidae Ehrenberg, 1839 The test is built of radially- fibrous calcite and deposited in successive laminae. It forms a trochospiral with an evolute spiral side and an involute umbilical side. The umbilicus is filled with plugs, and throughout it has radial canals or fissures and intra- septal and subsutural canals. The aperture is umbilical, basal, and single to multiple. Late Cretaceous (Maastrichtian) to Miocene. Subfamily Laffitteininae Hottinger, 2013 The test is trochospiral, involute to planispiral, with a enveloping canal system and vertical fissures on both sides. Umbilical and spiral sides not differentiated in structure. Umbos on both sides are pierced by numerous funnels. The aperture is a narrow slit extending obliquely over apertural face. Late Cretaceous to Miocene. • Cuvillierina Debourle, 1955 (Type species:  Cuvillierina eocenica Debourle, 1955 = Laffteina vallensis Ruiz de Gaona, 1948). The lenticular involute test is auriculate in outline with an angular periphery. The surface shows reticulate ornamentation and pillars. The enveloping canal system is derived from heaving feathering of the septal sutures. Vertical canals are present in the umbonal region. Bilamellar with intrasep- tal space widening towards the periphery. The periphery is sharp but not keeled. The aperture is an areal, comma-s haped slit. Early Eocene (Ypresian) (Fig. 6.19; Plate 6.28, figs 20-2 1). • Laffitteina Marie, 1946 (Type species: Laffitteina bibensis Marie, 1946). The test is trochospiral, almost planispiral, involute, and covered on both sides with an envel- oping canal system. The periphery is rounded. A single areal aperture/ foramen forms a slit in the septum. Late Cretaceous to Early Paleocene (Maastrichtian to Danian) (Fig. 6.19; Plate 6.25, figs 11- 12; Plate 6.28, figs 1- 2). 432 Evolution and Geological Significance of Larger Benthic Foraminifera Intraseptal canal system 1 2 Foramen Proloculus Side view, Funnels rounded periphery 3 4 Side view, sharp periphery 5 6 Figure  6.19. 1- 4) Laffiteina vanbellini Grimsdale, 1952. Early Eocene, Mushoral Well 1, Iran, NHM P40691, 40692-4 ; 5- 6) Cuvillierina vallensis (Ruiz). Early Eocene, Punta Iruarriaundiete, Guipuzcoa, Spain, NHM P4640. Scale bars = 0.5mm. Subfamily Daviesininae Hottinger, 2013 This subfamily is characterized by trochospiral, bilamellar- perforate, heavily ornate tests. Middle Paleocene to Late Oligocene. • Daviesina Smout, 1954 (Type species: Daviesina khatiyahi Smout, 1954). The test is flat- tened, trochospiral and large, with slightly unequal alar prolongations. The walls are thick, perforated and ornamented with thick pillars. Septa are secondarily doubled with no marginal cord. The umbilicus has a flap or plate. Prominent umbilical pillars, fis- sures and complex intra-s epta canals are distributed unequally on both sides of the test. Middle Paleocene to Late Oligocene (Plate 6.6, fig. 10; Plate 6.27, figs 13-1 4; Plate 6.28, fig. 26). Subfamily Rotaliinae Ehrenberg, 1839 The test is trochospiral with an umbilical plug formed by superposed folia fused at their tips, or by a single undivided secondary deposits or piles. Radial canals, intrasep- tal and subsutural canals are present. Late Cretaceous (Coniacian) to Early Miocene. The Cenozoic Larger Benthic Foraminifera: The Paleogene 433 • Medocia Parvati, 1971 (Type species: Medocia blayensis Parvati, 1971). The test is len- ticular and trochospiral; septa are doubled with septal passages. The spiral chambers are dorsally evolute and ventrally involute. Secondary deposits fused to a compact mass covering the umbilical area and are pierced by few tubular large funnels that are not always parallel, but connect earlier parts of the spiral side to the surface of the umbilical side. The spiral side has a thick lamellar wall. Middle Eocene (Lutetian). • Rotorbinella Bandy, 1944 (Type species:  Rotorbinella colliculus Bandy, 1944). A rotaliid with a single undivided umbilical pile with short mostly free folia (see also Revets, 2001; Hottinger, 2013). Late Cretaceous to Early Paleocene. • Rotalia Lamarck, 1804 (Type species: Rotalites trochidiformis Lamarck, 1804). The spiral side is evolute and smooth, while the umbilicus side is filled with a columellar structure produced by the fusing of the tips of the folia. Eocene (Plate 6.30, fig. 7). Subfamily Redmondininae Hottinger, 2013 This subfamily is characterized by coarsely perforated walls with a canal system that tends to extend onto the spiral side and a reduced umbilical filling. • Redmondina Hasson, 1985 (Type species: Redmondina henningtoni Hasson, 1985). The spiral side is evolute and inflated. The folia are short, free or fused at the tips to form a ring of imperforate umbilical piles. Paleocene to Middle Eocene (Bartonian). Subfamily Kathininae Hottinger, 2013 The test is lenticular or conical with chambers arranged in single or multiple spiral. The folia are small with fused folia forming a solid mass perforated by numerous funnels. Late Paleocene to Early Eocene (Thanetian to earliest Ypresian). • Dictyokathina Smout, 1954 (Type species:  Dictyokathina simplex Smout, 1954). The test has a multiple spire, as in Dictyoconoides. The umbilical side is covered by numerous pillars. Vertical radial canals penetrate the umbilical region and extend from the umbilical apertures of the chambers to the external pores. Paleocene to earliest Eocene (Plate 6.27, fig. 21). • Kathina Smout, 1954 (Type species: Kathina delseota Smout, 1954). The test is lenticu- lar in shape with a sharp un-k eeled periphery. The spiral side is smooth. The ventral side is filled by a solid umbilical mass pierced by numerous slits or parallel funnels. Middle Paleocene to earliest Eocene (Selandian to early Ypresian) (Plate 6.28, figs 24- 25). Subfamily Lockhartiinae Hottinger, 2013 This subfamily is characterized by a heavily ornamented spiral side and a complex umbilical structure, where the umbilical cavities are delimited by successive foliar walls and numerous parallel umbilical piles (Late Paleocene to Middle Eocene) (Fig. 6.20). • Dictyoconoides Nuttall, 1925 (Type species: Conulites cooki Carter, 1861). The test is conical with multiple intercalated spires of small rectangular chamberlets. The 434 Evolution and Geological Significance of Larger Benthic Foraminifera umbilical side is filled by pillars separated by spaces and cavities of equal sizes. Intraseptal and subsutural canal systems are present. Middle Eocene (Plate 6.28, figs 7-8 ). • Lockhartia Davies, 1932 (Type species:  Dictyoconoides haimei Davies, 1927). The test is conical to lenticular with a simple spire of numerous chambers, similar to Dictyoconoides but lacking the intercalated spires. The dorsal side is ornamented with nodes, and the umbilicus is filled with numerous pillars with numerous cavi- ties communicating with the chambers. Paleocene to Middle Eocene (Plate 6.28, figs 9- 15). Prolocuclus Chamber Piles Folia C B Folia Piles Prolocuclus A Folia Piles Figure 6.20. A) Lockhartia haimei (Davies), Paleocene, Qatar, Smout coll., NHM P40156; B) Dictyoconoides kohaticus Davies, Early Eocene, Kohat Shales, India, NHM P22632; C) Sakesaria dukhani Smout, paratype, Paleocene, Qatar, NHM P40203. Scale bars = 0.5mm. The Cenozoic Larger Benthic Foraminifera: The Paleogene 435 • Rotaliconus Hottinger, 2007 (Type species:  Rotaliconus persicus Hottinger, 2007). The test is trochospiral with a coarsely perforate, evolute and strongly convex dorsal side. The ventral side is involute and flattened, with a smooth or slightly pustulose, weakly perforate umbilical face. The umbilicus is covered by umbilical plates and a canal system is absent. There is a single interiomarginal aperture. Late Middle Eocene. • Sakesaria Davies, 1937 (Type species:  Sakesaria cotteri Davies, 1937). The test is elongate, with a very high trochospiral. The umbilicus is filled with pillars; cavities communicate between the chambers as in Dictyoconoides. Paleocene to Early Eocene (Plate 6.28, figs 1- 2). Subfamily Pararotaliinae Reiss, 1963 The test is trochospiral, with an enveloping canal system but with umbilical cavities. Late Cretaceous (Coniacian) to Holocene. • Camagueyia Cole and Bermudez, 1944 (Type species: Camagueyia perplexa Cole and Bermudez, 1944). The test is conical with an elevated spiral side and a flattened umbilical side. Thick walls result in reduced chamber lumen. Few, but massive pillars fill the umbilical region. Middle Eocene. • Neorotalia Bermüdez, 1952 (Type species: Rotalia mexicana Nuttall, 1928). The test is low trochospiral, with a simple umbilical boss and pillared walls, both ventrally and dorsally. The aperture is single, areal, with no complicated umbilical canal sys- tem. Early Oligocene (P18) to Late Oligocene (Burdigalian) (Plate 6.30, fig. 17; see Chapter 7). Family Miogypsinidae Vaughan, 1929 The test is flattened to biconvex. The microspheric form has a trochospiral or planispi- ral early spire, while the megalospheric form has a bilocular embryonal stage followed by a fan of median chamberlets. Middle Oligocene to Middle Miocene. Several related genera can be identified, thus: (A) Uniserial coil around the megalospheric proloculus. 1. Initial coil around the megalospheric proloculus: (a) Fan of only one or two additional chambers attached to the spire in the median equatorial plane: • Americogypsina BouDagher-F adel and Price, 2010 (Type species:  Americogypsina braziliana BouDagher- Fadel and Price, 2010b). The test is moderately small, tro- chospiral with a tendency to become uncoiled and biconvex in axial section with a thick granulated wall. In equatorial section, the chambers increase rapidly in size but their internal and external periphery are covered by small incomplete cham- berlets. The periphery of the test is covered with pronounced fissures and granula- tions. Pillars cover the umbilical and spiral parts of the test. Americogypsina differs from Paleomiogypsina in having one row of small incomplete and irregular cham- berlets on the periphery of the first whorl and two to three rows of mainly incom- plete thick-w alled chamberlets on the periphery of the last whorls. Americogypsina 436 Evolution and Geological Significance of Larger Benthic Foraminifera is the ancestral form of the American Miolepidocyclina, while Paleomiogypsina is the ancestor form of the typical Miogypsina. Middle Oligocene (see Fig. 6.21, Plate 6.30, figs 4- 5, 18). • Paleomiogypsina Matsumaru, 1996 (Type species:  Paleomiogypsina boninensis Matsumaru, 1996). There is a fan of only one or two additional chambers attached to the spire in the median equatorial plane. Paleomiogypsina differs from Neorotalia (see above) in being low trochospirally coiled and in having small chamberlets scattered on the periphery of the last whorl. This the evolutionary beginning of the fan of cham- berlets which was to develop in Miogypsinella. Early Late Oligocene (Plate 6.30, fig. 6). (b) Fan extends, producing a broad equatorial layer of ogival chambers: • Miogypsinella Hanzawa, 1940 (Type species:  Miogypsinella borodinensis Hanzawa 1940). Miogypsinella differs from Paleomiogypsina in having a fan of equatorial chamberlets. Loeblich and Tappan (1988) considered Miogypsinella to be a synonym of Miogypsinoides (see below), but the two taxa are easily distinguished and they have very different stratigraphical ranges. Miogypsinella differs from Miogypsinoides in having a weak trochospiral initial coil, and the lateral walls of the initial spire and the succeeding fan of ogival median chamberlets are much thinner. Late Oligocene to Early Miocene (Plate 6.30, figs 3, 8-1 0; Plate 7.14, fig. 3). • Boninella Matsumaru, 1996 (Type species: Boninella boninensis Matsumaru, 1996). Shows two spires of subquadrate chambers. Late Oligocene. 2. Embryont coils that are virtually planispiral, with only one whorl around the megalospheric proloculus, and a septal canal system that is weakly present: (a) The lateral walls of the initial spire and the succeeding fan of ogival median chamberlets become very thick and solid: • Miogypsinoides Yabe and Hanzawa, 1928 (Type species:  Miogypsinoides dehaarti Van der Vlerk, 1924). The septa of all the chamberlets of Miogypsinoides also pos- sess a clear intraseptal canal system. Oligocene (Rupelian, P19, to Chattian, P22, in America), Late Oligocene to Miocene (Chattian, P21b, to Burdigalian, N7, in the Mediterranean, P22 to Early Langhian N8, in the Indo- Pacific province) (Plate 6.30, fig. 11; see Chapter 7). (b) Lateral walls have cubiculae between the lamellae which begin to split apart: • Miogypsinodella BouDagher- Fadel et  al., 2000 (Type species:  Miogypsina (Miogypsina) primitiva Tan Sin Hok, 1936). The embryont coil is similar to that of Miogypsinoides (see Chapter 6), it is virtually planispiral, but there is only one whorl around the megalospheric proloculus, and a septal canal system is present. However, the lateral walls have gaps between the lamellae, which begin to split apart and form the beginnings of lateral chamberlets. This splitting results in thick-w alled irregu- lar chamberlets, unlike the regularly formed, stacked chamberlets of Miogypsina (see Chapter 6). Late Oligocene (late Chattian) to Middle Miocene (Langhian) (see Chapter 7). The Cenozoic Larger Benthic Foraminifera: The Paleogene 437 (B) Microspheric specimens which possess a uniserial coil around the proloculus, and a megalospheric embryont that has no coil around the proloculus, but two bidirec- tional coils around the proloculus, with lateral cubiculae regularly stacked on each side of the median layer, but lacking a canal system. 1. Embryont near the apex: • Miogypsina Sacco, 1893 (Type species Nummulina globulina Michelotti, 1841). Early species have megalospheric nepionts in which two series of chambers surround the deuteroconch unequally (e.g. M. borneensis Tan Sin Hok). In advanced forms (e.g. M. indonensis Tan Sin Hok) the series become equal, and both surround the megalospheric deuteroconch by means of equal half whorls. Microspheric speci- mens (e.g. the syntypic specimens of M. borneensis Tan Sin Hok) possess a uniserial coil around the proloculus, as in megalospheric Miogypsinella, the ancestral genus. Latest Early Oligocene (Rupelian, P21, in America), Late Oligocene (Chattian, P22, in Mediterranean), Early Miocene (Aquitanian, N4, in the Far East) to Early Miocene (Burdigalian, in America) and Middle Miocene (middle Serravallian, in the Indo- Pacific province) (Plate 6.30, figs 13, 15; see Chapter 7). 2. Embryont about midway between the centre of the test and the periphery: • Miogypsinita Drooger, 1952 (Type species:  Miogypsina Mexicana Nuttall, 1933). The embryont is peripheral in the microspheric test, but half way between the periphery and the centre of the megalospheric test. The embryont is divided into two unequal chambers with a straight septum and two unequal principal auxiliary chambers. Equatorial chambers have a diamond shape. Late Oligocene to Early Miocene. • Miolepidocyclina Silvestri, 1907 (Type species: Orbitoides (Lepidocyclina) burdigalen- sis Gümbel, 1870). The embryonic apparatus, consisting of a large protoconch and deuteroconch, is surrounded by a thick wall. The megalospheric nepiont is similar to that of Miogypsina, with no coil around the proloculus but has two bidirectional coils around the proloculus. However, the nepiont is centrally placed, instead of being at the edge of the test, as in Miogypsina. Early Oligocene (Rupelian, P20) to Early Miocene (Burdigalian) (Plate 6.30, figs 12, 14, 16)). Family Chapmaninidae Thalman, 1938 The test is conical with a trochospiral initial part, followed by a uniserial part and a tubular apertural system. Septa are invaginated into tube pillars. Late Paleocene to Late Miocene (Tortonian). • Angotia Cuvillier, 1963 (Type species:  Angotia aquitanica Cuvillier, 1963). The embryonic apparatus is bilocular, and is followed by a series of chambers form- ing a high cone with a flattened base. Hollow pillars and tunnels, perpendicular to the outer margin, fill the centre of the test. Middle Eocene (late Lutetian) (Fig. 6.17). • Chapmanina Silvestri, 1931 (Type species: Chapmanina gassinensis Silvestri, 1905). A bilocular embryonic apparatus is followed by a reduced trochospiral stage, followed 438 Evolution and Geological Significance of Larger Benthic Foraminifera Oligocene Early Miocene Middle Miocene Epoch RUPELIAN CHATTIAN AQUITANIAN BURDIG. LANGHIAN SERRAVALLIAN Stage Planktonic Zones Neorotalia Paleomiogypsina Miogypsinella Miogypsinoides Miogypsinodella Miogypsina Americogypsina Miolepidocyclina Figure 6.21. Phylogenetic chart showing the evolutionary lineages of the American Miogypsinidae (from BouDagher-F adel and Price, 2010b). by rapidly enlarging discoidal chambers in an uniserial arrangement. Peripheral sec- ondary septa are fissured at the umbilical surfaces. The central parts of the cham- bers have concentric rings of tubular pillars of identical size within a chamber. The aperture consists of multiple openings with internal tubes to the previous septum, resulting, as in Angotia, in the invagination of the septum. Middle Eocene to Late Miocene (Tortonian) (Fig. 6.17). • Crespinina Wade, 1955 (Type species:  Crespinina kingscotensis Wade, 1955). The microspheric test has an early planispiral coil, followed by embracing and annular chambers. The megalospheric embryo has a globular protoconch and a reniform deuteroconch, followed by a few annular undivided chambers, and later by rectilin- ear chambers forming a low conical test. The inside of the test is filled by unaligned hollow pillars. Late Eocene to Early Oligocene (Fig 6.17). • Ferayina Frizzell, 1949 (Type species: Ferayina coralliformis Frizzell, 1949). The surface has saucer- like rectilinear chambers with additional ribs and costae. There is a flat imperforate apertural face, and an aperture consisting of multiple pores at the end of hollow pillars extending to the previous septum. Middle Eocene (Fig. 6.17). (Ma) (Md) The Cenozoic Larger Benthic Foraminifera: The Paleogene 439 • Sherbornina Chapman, 1922 (Type species: Sherbornina atkinsoni Chapman, 1922). The test is large, discoidal, with annular chambers in the adult. Lateral walls have corrugations near the sutures, which alternate from chamber to chamber. Septal and radial canals are present and terminate in coarse pores at the outer surface. Late Paleocene to Middle Miocene. Family Calcarinidae Schwager, 1876 The test is enrolled with protruding spines. Late Cretaceous (Maastrictian) to Holocene. • Silvestriella Hanzawa, 1952 (Type species: Calcarina tetraëdra Gümbel, 1868). The test is large, with three to four large radial spines, resulting in a tetrahedral form. Spines arise from the early whorl of the chambers and widen rapidly. Spine canals arise from the interseptal spaces of the early chambers. The spine surface has numer- ous large pores. Solid pillars may be present between the outermost chambers. Middle Eocene (Lutetian to Bartonian) (Plate 6.15, fig. 10). • Meghalayana Matsumaru and Sarma, 2008 (Type species:  Meghalayana indica Matsumaru and Sarma, 2008). The test is ovoidal to ellipsoidal with four spines. It resembles the Cretaceous Siderolites (see Chapter 5), but is different from the latter in having the raspberry- like arrangement of the megalospheric form in early stage. Late Eocene (Plate 6.11, fig. 7). Family Elphidiidae Galloway, 1933 The test is planispiral to trochospiral and uncoiled with sutural pores, canals and supplementary apertures. The aperture is single or multiple. Paleocene to Holocene. • Elphidium de Montfort, 1808 (Type species:  Nautilus macellus var. 3 Fichtel and Moll, 1798). The test is lenticular, planispiral and involute or partially evolute with deeply incised sutures. The umbilical plug has vertical canals communicating with the spiral canals. Cellanthus de Montfort, 1808 is similar to Elphidium, but with a fully developed septal flap. Eocene to Holocene (Plate 6.29, fig. 1). • Pellatispirella Hanzawa, 1937 (Type species: Camerina matleyi Vaughan, 1929). The test is lenticular, planispiral and involute with an umbilicus perforated by canals. Septa are folded at the base and have numerous transverse canals. Middle Eocene (Plate 6.16, figs 5-6 ). 6.3 Biostratigraphy and Phylogenetic Evolution Larger benthic foraminifera are widely distributed in Paleogene carbonates (Fig. 6.22). They evolved gradually, forming a succession of biometrically large populations within important phylogenetic lineages, and have been essential in the development of Tethyan carbonate stratigraphy. Their systematics and biostratigraphy have been exten- sively studied and a number of zonations based on their occurrence have been erected 440 Evolution and Geological Significance of Larger Benthic Foraminifera Figure 6.22. The biostratigraphic range and diversity of the main superfamilies (as shown by the horizontal scale of the spindles) found in Tethys during the Paleogene. The Cenozoic Larger Benthic Foraminifera: The Paleogene 441 Larger benthic foraminifera zonaon Backreef/reef Forereef/reef First Occurrences Epoch - Stage Miogypsinoides complanatus Nephrolepidina sumatrensis L Miogypsinella/ Paleomiogypsina H. (Vlerkina) borneensis Nephrolepidina Neorotalia Eulepidina Miogypsina(American E province) Borelis pygmaeus Nummulites fichteli Cycloclypeus Lacazinella Heterostegina gracilis L Borelis Heterostegina Nummulites fabianii sensu lato Pseudofabularia Nummulites boulangeri Alveolina elongata Nummulites lyelli M Praerhapydionina Nummulites herbi Alveolina prorrecta Nummulites crassus Alveolina munieri Nummulites benehamensis Alveolina s pes Nummulites laevigatus Halkyardia Alveolina violae Nummulites manfredi Alveolina dainellii Nummulites cantabricus E Alveolina schwageri Nummulites burdigalensis Alveolina trempina Assilina pomeroli Alveolina corbarica Nummulites atacicus Alveolina moussoulensis Nummulites robus formis Alveolina ellipsoidalis Nummulites minervensis Amphistegina Alveolina vredenburgi Assilina prisca L Glomalveolina levis Miscellanea meandrina Glomalveolina primaeva Lockhar a condi Nummulites M Daviesina Miscellanea globularis Discocyclina/ Daviesina E Rotorbinella skourensis Laffi eina bibensis Dictyokathina/ Lockhar a Figure 6.23. Paleogene larger benthic foraminifera biozones, as refined in this study, with diagnostic first and last occurrences. (BouDagher- Fadel and Price, 2016). A larger foraminifera zonation of the Paleocene and Eocene of the Tethyan realm was published by Serra-K iel et al. (1998) as one of the results of the IGCP 286 Early Paleogene Benthos study. Cahuzac and Poignant (1997) proposed a similar larger foraminifera zonation for the Oligocene to Miocene of 66.0 56.0 33.90 23.03 Age (Ma) Paleocene Eocene Oligocene Danian Selandian Thanetian Ypresian Lutetian Bartonian Priabonian Rupelian Chattian PLANKTONIC ZONES "Letter Stages" SB-Zones 442 Evolution and Geological Significance of Larger Benthic Foraminifera the west European basins. In the biostratigraphy of the Far East, the use of the “Letter Stages” is well established, although correlation with global standard bio- and chro- nostratigraphical scales have until recently proved difficult, but which are now inte- grated in Fig 6.23. These difficulties were due to the fact that the larger foraminiferal assemblages are facies controlled, and because of taxonomic confusion between the Far and Middle East assemblages, where taxonomic overlap (synonyms) is known to exist. The distribution and ranges of the Paleocene superfamilies and genera are plot- ted in Charts 6.1 and 6.2, and the their correlation with the global Planktonic Zones (PZ) are given in Fig 6.23. In the following sections, the biostratigraphic “Letter Stages” of Far East, and the biostratigraphic and phylogenetic evolution of the Paleogene textulariides, miliolides, and rotaliides are discussed. 6.3.1 The “Letter Stages” of SE Asia and provincial biostratigraphy The “Letter Stages” subdivision of the Indo- Pacific Cenozoic (Leupold and van der Vlerk, 1931; Adams, 1970; Chapronière, 1984; BouDagher-F adel and Banner, 1999; BouDagher-F adel, 2008; Advocaat et al., 2014; Sun et al., 2015; Wang et al., 2015; An et al., 2015; BouDagher-F adel et al., 2015; Hu et al., 2015; Li et al., 2016; BouDagher- Fadel et al., 2017) is based on larger foraminifera, some of which have also been used as range fossils in western hemisphere stratigraphy (Barker and Grimsdale, 1936; Caudri, 1996). In the Far East, where the Cenozoic sedimentary sequences are dominated by warm water, shallow marine carbonates, these deposits are largely biogenic in ori- gin, primarily of benthic faunas and algae. The 19th and early 20th Century workers attempted to utilize new stratigraphic schemes involving benthic faunas to date and study these rocks. Martin (1880), studying the mollusc faunas of Java and surrounding areas was able to follow the work of Charles Lyell in Europe in utilizing the ratios of extant to extinct faunas to indicate relative ages. Two major contributions to Cenozoic stratigraphy came from the work of Martin. Firstly, he showed that the mollusc faunas of the East Indies developed separately from those of Europe, and secondly, as a con- sequence of this, the stage names of Europe could not be correlated with confidence to the Indonesian region. Martin studied a number of Javanese marine sections in detail, determined the ratios of extant to extinct molluscs, and predicted the stratigraphic order of these strata. Mapping, and studies on the other regionally significant fossil group, the larger foraminifera, supported the mollusc stratigraphy, but there was still little idea on how it fitted into the European stratigraphic column. By the late 1920s, the larger foraminifera had become the preferred fossil group for biostratigraphy in the Indonesian area (Sharaf et al., 2014; Li et al., 2015; Gold et al., 2017a, b; Breitfeld et al., 2016; Advocaat et al., 2017; White et al., 2017). They had the advantage that they were more abundant than molluscs, and also a scheme was devel- oped that utilised assemblage zones rather than percentages of extant forms. Using molluscs to identify and correlate sections required extensive knowledge of both living and fossil species. The larger foraminifera assemblage zones could be identified by the presence of a few key taxa, usually with a hand-l ens in the field. They were proposed The Cenozoic Larger Benthic Foraminifera: The Paleogene 443 as alternatives to the European stages and they still are recognizable as fundamental stratigraphic units over the Indo- Pacific area. Van der Vlerk and Umbgrove (1927) published the Letter Classification of the Indonesian “Tertiary”, based on larger foraminifera. This scheme subdivided the “Tertiary” into seven parts. Six parts were labeled “a” to “f” (e.g. “Tertiary a”, or “Ta” for short), and a seventh part characterized by non-o rbitoidal “Tertiary” fora- minifera between Tf and Tg was noted; Tg was considered Late Miocene. They equate to the Cheribonian and Sondien regional stages, which are roughly equivalent to the Early and Late Pliocene. The “Quaternary” or Pleistocene is represented in the mollusc scheme by the Bantamien regional stage, with 70 % or more of extant species. This per- iod therefore has no equivalent in the Letter Stages (Leupold and van der Vlerk, 1931). This scheme was immediately successful and was rapidly adopted as a worka- ble standard biostratigraphic scheme for the SE Asian region, even if the correla- tion with the European epochs and stages was not adequately known. The authors pointed out that the taxonomy and detailed biostratigraphy of the genera used in this initial scheme was not adequately known, and suggested that future work would probably increase the number of recognizable subdivisions. The difference between Tf and Tg was originally based on a faunal turnover, with the common taxa Lepidocyclina and Miogypsina, and the less frequent forms Austrotrillina and Flosculinella, all disappearing from the carbonate facies, whereas no new forms appeared in “Tertiary g”. The importance of this faunal turnover as a biostrati- graphic event is underlined by other carbonate facies organisms showing a change at the same time as the non- Lepidocyclina larger foraminifera extinctions. In brief, shallow marine biohermal carbonates changed from mixed coral and coralline algal boundstones with larger foraminiferal grainstones to more dominant coral reefs with a marked increase in Halimeda green algae. The latter are often preserved in recognizable forms, or as an increase in micrite and fine bioclastic products from the early breakdown of aragonitic platelets. Larger foraminiferal grainstones are only rare after this event, as deeper photic Cycloclypeus facies or minor Operculina/ Amphistegina or Alveolinella calcarenites. Most well and field sections encountering limestones of this later Miocene or Pliocene age in SE Asia do not sample rocks composed of larger foraminifera tests. Although, the use of the “Letter Stages” is widespread, correlation with global standard bio- and chronostratigraphical scales were for a long time uncertain and problematic (Berggren, 1972; Blow, 1979; Adams, 1984). The difficulties are due to the larger foraminiferal assemblages being facies controlled, and so changes in relative sea level affecting the photic zone, wave base, clastic sediment supply, influence the larger foraminifera (and associated algal) assemblages. However, the major planktonic foraminiferal bio-e vents for the Late Mesozoic and the Cenozoic were recently calibrated against the biostratigraphical time scale and the radio- isotopes by BouDagher- Fadel (2013). Those of the Cenozoic were compared with the earlier zonations of Berggren and Pearson (2005) and Wade et  al. (2011). The biostratigraphic resolution of the alpha- numeric biozonations (Planktonic Zones, PZ) were determined by distinct evolutionary lineages of morpho-s pecies, and were deduced through extensive industrial work, calibrated against Sr- isotopes and other biostratigraphic markers (see BouDagher- Fadel 2015). The resulting PZ stages, have 444 Evolution and Geological Significance of Larger Benthic Foraminifera been used to calibrate and correlate the provincial larger benthic zones, and hence link the “Letter Stages” to other bioprovinces. In this work, the “Letter Stages” of the Paleogene are assigned as follows (see Fig. 6.23, and BouDagher- Fadel, 2015): • The Paleocene [Ta 1] is divided into two parts: o Ta 1(a) (corresponds to P1-P 2 PZ, 66.0-6 1.6Ma), containing rare larger ben- thic foraminifera, is based on the appearance of the conical forms Dictyokathina and Lockhartia in the Western Tethys, and corresponds to Early Paleocene. o Ta 1(b) (corresponds to P3-P 5a PZ, 61.6- 56.0Ma), Middle to the Late Paleocene, is based on the presence of Fallotella in Western Tethys and the global appearance of Discocyclina, and Miscellanea globularis and Daviesina. The upper part of Ta 1(b) cor- responds to the appearance of Nummulites, Alveolina and the presence of Ranikothalia in Tethys. This zone is also equivalent to the appearance of the first orbitoids, Actinosiphon in the American province. • Ta 2 (corresponds P5b- P9 PZ, 56.0- 47.8Ma), Early Eocene, is based on the appearance of the cosmopolitan form Amphistegina, and the Tethyan Orbitolites. It is equivalent to the presence of the conical form, Verseyella in the American province. • Ta 3 (corresponds to P10-P 15a PZ, 47.8- 35.1Ma), Middle Eocene, corresponding to the appearance of the broadly conical cosmopolitan form, Halkyardia and the Tethyan Chapmanina and Somalina in Tethys. The top of Ta 3 is defined on the appearance of Pellatispira in Tethys. This zone corresponds to the appearance of L. (Lepidocyclina), L. (Nephrolepidina) and Lepidocyclina sp. in the American province. • Tb, Late Eocene (correspomds to P15b-P 17 PZ, 37.8- 33.9Ma), is based on the appearance of the cosmopolitan Heterostegina sensu lato. • Tc– Td (Tc corresponds to P18-P 19 PZ, 33.8- 30.3Ma; Td corresponds P20- P21a PZ, 30.3-2 8.1 Ma):  base of Tc, Early Oligocene, is based on the appearance of Cycloclypeus in Tethys and corresponds to the appearance of Eulepidina and Astrolepidina in the American province, and Td corresponds to the appearance of Miolepidocyclina and Miogypsina in the American province. • Early Te (Te1– Te4) (corresponds to P21b – P22 PZ, 28.1- 23.0Ma), Late Oligocene, is based on the appearance of Miogypsinella and L. (Nephrolepidina) in the Far East, Miogypsina in Europe and corresponds to the appearance of Miogypsinita in the American province. In Tethys, the boundary between Te1 and Te2 is based on the disappearance of Paleomiogypsina and the appearance of Miogypsinoides compla- natus. Amphisorus martini marks the beginning of Te3, while Te4 is marked by the appearance of Miogypsinoides formosensis (see Chart 7.1). Parallel to this zonation, new schemes for the Mediterranean region, based on the biogeographic zonation of shallow benthic larger foraminifera and their direct corre- lation with magnetostratigraphy, were defined by Serra-K iel, et al. (1998). They pre- sented a system of numbered units for the “Tertiary”; SBZ 1-2 3 for the Paleogene (see Fig. 6.23; Chart 6.2). These zonations were subsequently correlated (Less, 1998; Less et al., 2007) with the Orthophragminae (orbitoidal larger foraminifera) for the Late The Cenozoic Larger Benthic Foraminifera: The Paleogene 445 Paleocene- Eocene of the Mediterranean; OZ 1a- 16. These have all now been cross cor- related by the work described in BouDagher- Fadel (2015) and shown in Fig 6.23 and Chart 6.2. Below are described the three main groups that dominated the larger foraminifera assemblages of the Paleogene, (see Chart 6.1 and 6.2), namely: • the agglutinated textulariides (which evolved from small surviving Cretaceous forms), • the miliolides (which became abundant in the Eocene, imitating in their evolution that seen in Cretaceous), and • the rotaliides (mainly the orbitoids). 6.3.2 The Textulariides of the Paleogene In the Paleogene, most of the stratigraphically important agglutinated foramin- ifera are included in three different superfamilies (see Fig 6.2, 6.3 and Charts 6.1 and 6.2), which can be divided into two morphological groups; the elongated forms with a trochospiral, biserial or triserial in the early stage, the Textularioidea, and the conical forms, which are homeomorphs of the Cretaceous orbitolinids, such as the Coskinolinoidea and Orbitolinoidea. The Textularioidea are mainly found in rocks from the Late Paleocene to Eocene formed in shallow water environments of Tethys, while the conical coskinolinids and orbitolinids are rare in the Middle Eocene, but absent in the Late Eocene shallow and extremely specialized facies of Tethys (Hottinger, 2007). • The Textularioidea evolved in parallel lineages, most probably from different ances- tors. The cosmopolitan forms, Pseudochrysalidina and Vacuovalvulina followed the same trends as their ancestors, the chrysalinids of the Cretaceous, by developing interior pillars in a multiserial test. The Tethyan form, Pfendericonus, evolved from a simple trochospirally enrolled ataxophramid, Arenobulimina, by developing internal pillars (Fig. 6.2). • The Coskinolinoidea in their turn evolved from the same ancestor, Arenobulimina, but followed different evolutionary trends, which involved reducing the arenobu- liminid spire with the latest part becoming uniserial in Coleiconus (Fig. 6.3), with chambers subdivided by scattered pillars, such as in the Tethyan Coskinon, or tubu- lar pillars as in the cosmopolitan Pseudolituonella (Fig. 6.2). In the Early to Middle Eocene, the cosmopolitan Coskinolina acquired, in addition to the pillars, vertical partitions. • The Orbitolinoidea, which thrived in the Early to mid- Cretaceous of Tethys, survived the Cretaceous boundary event with a single genus, Dictyoconus, that give rise to many descendants. The initial trochospiral had disappeared completely in this super- family and the chambers are subdivided by vertical and horizontal partition with pillars filling the centre. The foraminiferal apertures are disposed (as in Orbitolina, see Chapter 5) in alternating positions from one chamber to the next one. Many of the Orbitolinoidea were short ranged and are very important biostratigraphically 446 Evolution and Geological Significance of Larger Benthic Foraminifera in Tethys. Only Verseyella is unique to America. The peripheral tiered rectangu- lar chamberlets of Dictyoconus seem to belong only to this genus. The rest of the Paleogene forms evolved with different kinds of partitions and different ways of scattering their pillars (see Figs 6.2, 6.3). The larger benthic forms of the textulariids ranged throughout the Eocene, many disappearing at the Middle- Late Eocene boundary. Only Dictyoconus crosses the Eocene to Oligocene boundary, but then died out at the top of the Oligocene. 6.3.3 The Miliolides of the Paleogene The stratigraphically important miliolides of the Paleogene are mainly divided into three superfamilies: • the Alveolinoidea • the Milioloidea • the Soritoidea. The Alveolinoidea of the Paleogene include two main larger benthic families, the Alveolinidae and the Fabulariidae (see Fig. 6.24). The porcelaneous Alveolinidae evolved from a form with a simple milioline test, coiled in varying planes, with the axis of coiling normal to the apertural axis. The axis in the primitive forms is rotated so that an angle of 72° exists between the median planes of consecutive chambers in a quinqueloculine test, or 120° in a triloculine test or 180° in a spiroloculine or biloculine test. These forms evolved in turn to genera with streptospiral early coiling (e.g Pseudonummuloculina) (see Hottinger, Drobne and Caus 1989), but that were planispiral in the adult part and spherical to fusiform in shape (see Fig.  6.25). These alveolinids show an example of morphological conver- gence with those of the Cretaceous. However, a considerable gap of about 20 Ma exists in the fossil record between the Alveolinidae of the Paleocene and the mid- Cretaceous Praealveolina group (see Chapter 5). The oldest Paleogene form, Glomalveolina (with a streptoloculine origin and with a fixed axial coiling throughout ontogeny) is considered the ancestor of Alveolina. Glomalveolina is small, globular, and non- flosculinised with a spherical proloculus in a streptospiral coil. The streptospiral coil still exists in the B forms of Alveolina (see Fig. 6.26). Alveolina is a planispiral- fusiform form with a single tier of chamber- lets in each chamber and considerable thickening of the basal layer (flosculinisation) (Fig.  6.26). In the Eocene the amount of flosculinisation varies in different species of Alveolina and might be considered to be of specific value only. Alveolina evolved into larger forms through the Eocene and reached gigantic sizes before the end of the Middle Eocene. It become an essential member of Eocene carbonate fossil assem- blages. However, Alveolina forms became smaller again before completely disappearing at the end of Eocene. (see Fig. 6.27). The Cenozoic Larger Benthic Foraminifera: The Paleogene 447 ALVEOLINOIDEA SORITOIDEA Epoch - Stage L E L M E L M E Figure 6.24. Ranges of the main genera of the Paleogene Alveolinoidea and Soritoidea (also see Charts 6.1 and 6.2 for more details). Borelis, an extant form and the only survivor of the terminal Middle Eocene extinc- tion of alveolinids, is common in the Early Oligocene. A gradual stratigraphic succes- sion of alveolinid species marked its very slow evolution in the Oligocene (Fig. 6.28). Species evolved with an incremental increase in the length of the test from pole to pole, from globular through ovoid to elongated spindle-s haped. In the Te “Letter Stage” of 66.0 56.0 33.90 23.03 Age (Ma) Paleocene Eocene Oligocene Danian Selandian Thanetian Ypresian Lutetian Bartonian Priabonian Rupelian Chattian PLANKTONIC ZONES Lacazina Pseudolacazina Glomalveolina Alveolina Lacazinella Fabularia Pseudofabularia Borelis Bullalveolina Haymanella Mardinella Opertorbitolites Twaraina Orbitolites Peneroplis Spirolina Dendritina Yaberinella Archaias Globoreticulina Cyclorbiculina Cyclorbiculinoides Neorhipidionina Somalina Praerhapydionina Penarchaias Neotaberina Archiacina Sorites Amphisorus "Letter Stages" SB-Zones Figure 6.25. The evolution of Alveolina from a simple streptospiral origin (Pseudonummuloculina). roof pre-septal passage Glomalveolina post-septal passage proloculus flosculin regular coiling B form streptospiral nepiont Alveolina proloculus A form flexostyle Figure 6.26. Equatorial sections showing the original stage of the Alveolinidae, modified after Hottinger (1960). “A form” is the megalospheric form and “B form” is the microspheric form. The Cenozoic Larger Benthic Foraminifera: The Paleogene 449 psp p prp apf cht C ss dg flt fl pi B psp s b prp pp A ch pspprp cl D psp psp prp aps dg cht sp prp F G ch E 21 3 Figure  6.27. Reconstructions of different genera of the Paleogene alveolinids, modified from Reichel (1936). A:  Alveolina; B:  A flosculinized Alveolina; C:  Stereo-d iagram showing portion of two whorls; D: Glomalveolina with minimal flosculinisation; E: Borelis with only a pre-s eptal passage; F: Diagram of an internal cast showing the chamberlet form in Borelis; G: Bullalveolina with a large pre- septal passage (psp) and final chamber with multiple rows of small apertures. Abbreviations in figure: psp = post-s eptal passage, prp = pre-s eptal passage, flt = flexostyle, ch = chamber, cht = chamberlet, apf = apertural face, p = prolocu- lus, fl = flosculine, ss = secondary septa, s = septum, b = basal layer, pp = main apertures, pi = supplementary apertures, dg = direction of growth, 1 = secondary small tier of chamberlets which alternates with the larger ones producing “Y” shaped septa in axial section of Borelis, 2, normal chamberlet, 3, displaced chamber (from Hottinger, 2006). the Indo- Pacific, the alveolinids included Borelis pygmaeus (which persists to the top of the Te stage; BouDagher-F adel and Banner, 1999), where it is suddenly replaced by a more advanced form Flosculinella bontangensis (see Chapter 7). In the Mediterranean, a smaller form is recorded in the Oligocene, Bullalveolina, which evolved from being streptospiral to having a planispiral quinqueloculine test, but unlike the extant Borelis, Bullalveolina disappears at the top of the Oligocene. Hottinger (2006, 2007)  gath- ered forms that exhibit streptospiral nepionts in both generations and a planispiral- involute chamber arrangement in the adult stage of growth, such as Subalveolina, Praebullalveolina, Bullalveolina, and Globoreticulina, into the same group within his new subfamily Malatyninae. Sirel (2004, p. 38) classified the genus Malatyna into the milioline family Rivieroinidae, but according to Hottinger (2006, 2007) Malatyna can not be a milioline because it lacks the milioline pattern of growth, with two chambers per whorl and the apertural axis perpendicular to the coiling axis. Globoreticulina, clas- sified by Rahaghi (1978) in the milioline subfamily Fabulariidae, has an architecture similar to Malatyna and was consequently transferred by Hottinger (2006, 2007)  to the family Alveolinidae, and assigned together with Malatyna to the his subfamily Malatynae (which is not adopted here). 450 Evolution and Geological Significance of Larger Benthic Foraminifera ALVEOLINOIDEA Epoch - Stage P15 18 P14 P13 17 M P12 16 15 P11 14 P10 13 12 P9 11 P8 E P7 10 9 P6 8 7 6 P5 5 L 4 P4 3 M 2 P3 Figure 6.28. Stratigraphic ranges of some key species of alveolinids. 5 5 . 8 0 A g e ( M a ) E o c e n e P a l e o c e n e S e l a n d i a n T h a n e t i a n Y p r e s i a n L u t e t i a n B a r t o n i a n P L A N K T O N I C Z O N E S G l o m a l v e o l i n a p r i m a e v a G l o m a l v e o l i n a l e v i s A l v e o l i n a g l o b o s a A l v e o l i n a v r e d e n b u r g i A l v e o l i n a p a s t i c i l l a t a A l v e o l i n a e l l i p s o i d a l i s A l v e o l i n a m o u s s o u l e n s i s A l v e o l i n a c o r b a r i c a A l v e o l i n a t r e m p i n a A l v e o l i n a s c h w a g e r i A l v e o l i n a d a i n e l l i i A l v e o l i n a v i o l a e A l v e o l i n a e l l i p t i c a n u t t a l l i A l v e o l i n a e l l i p t i c a A l v e o l i n a m u n i e r i A l v e o l i n a p r o r r e c t a A l v e o l i n a e l o n g a t a " L e t t e r S t a g e s " S B - Z o n e s The Cenozoic Larger Benthic Foraminifera: The Paleogene 451 The Fabulariidae are all linked to the Cretaceous fabulariids via Lacazina, which survived the across Cretaceous- Paleocene boundary (see Charts 6.1 and 6.2). Lacazina evolved from a miliolid- bilocular ancestor, such as the quinqueloculine to biloculine Idalina (see Fig. 6.7) with its overlapping chambers and fixed apertural and coiling axes (see Chapter 5). No direct phylogenetic relationship is traced between the Santonian pil- lared Pseudolacazina and the unpillared, much smaller Middle Eocene Pseudolacazina (Hottinger et  al., 1989). All Paleogene representatives of Pseudolacazina are distin- guished from the Cretaceous forms in having lower chambers, subdivided by continu- ous chamber partitions (septula) (see Drobne, 1988; Hottinger, 1989). Fabularia is distinguished from Lacazina in lacking the monolocular stage and radial pillars. According to Haynes (1981), Fabularia evolved from the small, prima- tive Eocene form Miliola in the Middle Eocene by acquiring the involute bilocu- line last stage after the early quinqueloculine stage and the thick vertical partitions, which subdivide the chambers into elongated chamberlets with two tiers in outer whorls. On the other hand, Miliola may have given rise earlier, in the Late Paleocene, to Lacazinella, which has completely overlapping chambers. While many of the Alveolinoidea are still extant, most of the Fabulariidae did not survive the Eocene- Oligocene boundary, with the exception of Lacazina, which however died out within the Oligocene. Middle to Late Paleogene Tethyan shallow- water foraminiferal communities are dominated by Soritoidea (see Fig. 6.24), which exhibit a completely new line of evolu- tion for the miliolides (see Fig. 6.8). They appear to have evolved in the Eocene from forms with a simple planispiral, Spirolina, to a flat, evolute and flaring planispiral, Peneroplis. These porcelaneous, spiroliniform and peneropliform foraminifera devel- oped uniserial cylindrical chambers with radial partitions, which varied from simple, as in Praerhapydionina, to combined radial partitions with pillars, as in Neotaberina, with apertures varying from single, stellar- shaped apertures, such as Praerhapydionina, to others with multiple apertures, such as Rhabdorites. They are very common in the Middle East, where they were originally classified by Henson (1948) as “Rhapydionina” and “Rhipidionina”. However, both of these forms, previously described by Stache (1913) and later confirmed as megalospheric and microspheric forms by Reichel (1984), are different from those of the Eocene (Hottinger, 2007). In his Ph.D. thesis, Henson (1950) transferred his Eocene Rhipidionina species to the genus Meandropsina. However, the Meandopsinidae show similar morphological trends, but are confined to the Cretaceous and may have arisen from Praepeneroplis (see Chapter 5). Henson (1948) recognized in his Eocene material, mainly from Iraq, three taxa with similar structures: • “Rhapydionina” urensis (Plate 6.6, figs 3- 5) having a spiroliniform test, • “Rhipidionina” macfadyeni (Plate 6.6, figs 7- 8), and • “Rhipidionina” williamsoni (Plate 6.6, fig. 9) exhibiting a peneroplid test. All three taxa represent megalospheric forms and have the same basic patterns of chamber subdivision and a planispiral- involute nepiont (Hottinger, 2007). Hottinger emended and recognized a new genus, Rhipidionina designating “Rhipidionina” 452 Evolution and Geological Significance of Larger Benthic Foraminifera williamsoni as its type species. The genus Haymanella was originally described from the Paleocene of Turkey by Sirel (1999). However, it was shown by Hottinger (2007) that Sirel’s species is the only genus of the Peneroplidae and of the Soritidae, that combines a porcelaneous wall with an agglutination of coarse grains. In this respect, Haymanella is similar to agglutinating miliolines such as Agglutinella, Schlumbergerina or Siphonaperta (see Hottinger et  al., 1993). Eocene peneroplids apparently also show a progression through many lineages, leading to Archaias, Sorites and the rest of the Miocene discoid miliolines, such as Marginopora and Amphisorus (see Chapter 7). In the Oligocene, forms with a peneropline early stage became strongly flaring, as in Puteolina, becoming completely annular with Cyclorbiculina. The peneropline nepiont still occupies more than half of the test, but the annular chambers are by now divided into small rectangular chamberlets, as in Cyclorbiculina. These latter forms became flat and completely discoid, as in Sorites, with ogival chamberlets connected by a single open space or stolon, occurring in the equatorial plane. This lineage, after adopting the circular trend, became very successful and is still living in modern seas. Another extant form evolved in the Eocene from a peneropline ancestor, Archaias, an involute planispiral thickened form with pillars throughout the test, but with no annular stage of growth. A short trend occurred in Tethys during the Early and Middle Eocene, in which successive stolon layers formed a pile of discs (Hottinger and Drobne, 1980), as in Orbitolites. However, unlike the lineage of Sorites, Orbitolites does not have an penero- pline early stage, but rather only has oblique stolons connecting the adjoining cham- berlets (Hottinger, 2006) and chambers subdivided into small curved chamberlets with thick walls (see Fig. 6.8). In the Middle Oligocene of the Far East, one widespread group of Milioloidea had an alveolar exoskeleton, Austrotrillina, in order, presumably to harbour symbiotic algae. Austrotrillina was revised by Adams (1968), presenting Austrotrillina paucialveo- lata Grimsdale (Plate 6.9, Figs 14-1 6; Plate 6.10, figs 1- 2) as the most primitive species of the group, with a range starting in the Early Oligocene. However, in recent years, it has been recorded from the Late Eocene of Iran by Rahaghi (1980), and more recently Hottinger (2007) erected a new species for Rahaghi’s form, Austrotrillina eocaenica. According to Hottinger, this species in Rahaghi’s material is “primitive”, but cannot be interpreted as a direct ancestor of A. paucialveolata, because it is larger and has a more differentiated exoskeleton. Another doubtful occurrence of Austrotrillina was recorded from the Late Eocene of off- shore Tunisia (Bonnefous and Bismuth, 1982). The questionable phylogenetic relationship will not be resolved until more material is found that relates the Oligocene Far East taxa with those of the Eocene of the Middle East. The evolution of these forms in the Indo-P acific realm will be discussed in detail in Chapter 7. In the Caribbean, the ecological equivalents of the Tethyan Soritoidea are rep- resented by the discoidal Yaberinella (Plate 6.10, fig.  16; Plate 6.11, figs 10- 11), of miliolide-f abulariid origin (Hottinger, 1969), and Cyclorbiculinoides (Plate 6.10, fig. 13). The latter exhibits an unusual architecture with vertically superposed radial pillars (see Hottinger, 1979) of unknown origin (Hottinger, 2001). The Cenozoic Larger Benthic Foraminifera: The Paleogene 453 6.3.4 The Rotaliides of the Paleogene In the Paleogene, the rotaliides thrived in the warm shallow waters of Tethys. They became very large, reaching up to ~15cm diameter in forms such as Cycloclypeus. The problems associated with large size were solved by adding different structures to the foraminiferal test, such as filling the umbilical area with pillars, becoming annular by subdividing their chambers into small chamberlets, and by suspending their lamel- lae from a thickened marginal cord. Many rotaliides have canal systems (Fig.  6.29) within the walls of their plurilocular calcareous tests. Such constructions are called supplemental skeletons, representing infolded outer lamellae. The canals became very complicated in the Eocene with forms such as Pellatispira (see Hottinger et al., 2001). Their biological function has been investigated by many authors (Röttger et al., 1984; Hottinger et  al., 2001). In extant nummulitids, elphidiids, calcarinids, pellatispirids, they were found filled with protoplasm with permanently differentiated microtubuli (Hottinger and Dreher, 1974; Leutenegger, 1977; Hottinger and Leutenegger, 1980; Hottinger et al., 2001). These apparent morphological changes in the rotaliines of the Paleogene provided the foundation for the stratigraphically important superfamilies (and families) discussed below: • Planorbulinoidea • Nummulitoidea • Acervulinoidea • Asterigerinoidea (especially the Lepidocyclinidae) • Rotalioidea (especially the Miogypsinidae). As explained in Chapter 5, nearly all groups of large rotaliine foraminifera were derived, via separate lineages, from planispiral or trochospiral small rotaliine ances- tors, with a primitive canal system (Fig. 6.9), such as Rotalia or Cibicides. Figure  6.29. The canal system in a simple rotaliine test (after Hottinger, 2006). Abbreviations in dia- gram:  up  =  umbilical plate, spc  =  spiral canal, sis  =  spiral interlocular space, isc  =  intraseptal canal, ch = chamber, is = interlocular, f = foramen, ih = loop- hole. 454 Evolution and Geological Significance of Larger Benthic Foraminifera Cibicides shows a number of evolutionary trends (Fig. 6.30) leading to the develop- ment of attached forms in the Paleogene, irregularly uncoiled biserial small forms in the Eocene (Dyocibicides) and annular discoid small forms in the Paleocene to Miocene (Cycloloculina). Cibicides also gave rise to the Planorbulinoidea in the Eocene (Fig. 6.30) by developing rings of alternating chamberlets with peripheral apertures. The name of the single layered planorbulinids is derived from the genus Planorbulina in the family Planorbulinidae, which has a long trochoid, Cibicides- like keeled spiral and a small num- ber of irregular later chambers in an orbitoidal pattern. It has been suggested that both the extant forms Planorbulina and Planorbulinella derived from some Cibicides-l ike ances- tor (Drooger, 1993) in the Early Eocene. In both forms (Planorbulina and Planorbulinella) retrovert apertures of the later spiral chambers originated from irregular orbitoidal forms in the Cibicides-l ike ancestor. A particular feature belonging to the Planorbulinella mor- phology is the occurrence of relapse chambers, where after the occurrence of chambers with two stolons, one or more of the later chambers had reverted to having one stolon. Freudenthal (1969) introduced a morphometric analysis for Planorbulinella. His analysis concentrated on “Y counts” (the number of chambers with a basal opening) and measure- ments of the embryon expressed in diameter “d” and height “h”. Y was found to vary from 8 to 2. In the Oligocene a form similar to Planorbulinella, but with the apertures consisting of smaller basal openings instead of single slits, Planolinderina, might have also evolved from a Cibicides- like ancestor. The relapse chambers of Planorbulinella are not observed in Planolinderina, and Drooger (1993) noted that unlike in the phylogeny of Planorbulinella there is no Y=2 barrier in the phylogenetic development of Planolinderina. Immediately following the appearance of Planorbulinella, genera with thick lamellar walls and secondary deposits of calcite on both sides of the central embryont appeared in the Middle Eocene; the cosmopolitan Linderina and the American Eoannularia, with incipient orbitoidal morphology (see Figs 6.16, 6.30). Another lineage that evolved from Cibicides, with conical forms having umbilical cover plates or plugs and walls thickening on the spiral side by addition of lamellae, included the Late Cretaceous- Paleocene form, Cymbalopora and the Eocene Halkyardia (Figs 6.17, 6.30). The latter has also been reported from the Early Oligocene platform and periplatform limestones of Jamaica, as well as being recorded in Cuba, Ecuador, and Mexico (Robinson, 1996). The branching homotremids (Fig. 6.30), with planispiral and very high coiling, seem to have arisen from the Planorbulinidae in the Eocene. The extant form Sporadotrema, with flattened sides and cylindrical branches, first appeared in the Eocene. It has no pillar- like lamellar thickening as would be expected in Victoriella. The victoriellids, with three subfamilies, the Carpenteriinae, the Rupertininae and the Victoriellinae, also arose from a Cibicides- like ancestor. The Carpenteriinae are the first, primitive forms, and evolved from Cibicides by acquiring a distinct keel, extend- ing to the attachment area. The Rupertininae, with a distinctly flat spiral side, had the adult part of the test growing away from the attachment site. Members of this family are known from the Late Eocene, with Carpenteria still living to the present day. The Victoriellinae is a group in which the early stage was probably free living, but later became attached. These forms range from Late Eocene to Early Miocene, and are rep- resented by high spiral forms, such as Victoriella and Wadella, and low spral forms such as Maslinella. newgenrtpdf The Cenozoic Larger Benthic Foraminifera: The Paleogene 455 Figure 6.30. The different parallel lineages evolving from Cibicides in the Paleogene. 456 Evolution and Geological Significance of Larger Benthic Foraminifera Members of the planispirally-c oiled Nummulitoidea became very abundant in the Paleogene (see Fig.  6.31). They exhibited rapid evolutionary rates, and developed complex, large tests, which make them a very important index fossil group for shal- low marine environments (Schaub, 1981; Pignatti, 1998; Serra- Kiel et al., 1998). They probably evolved from a small rotaliine, via the the Late Cretaceous Sulcoperculina (see Chapter 5), by developing a complex system of a marginal cord and well- developed spi- ral and subsutural septal canals. The development of the canaliculate marginal cord, replacing the primary aperture of the nummulitids, was essential for growth, loco- motion, reproduction, excretion and protection (Röttger, 1984). This complex canal system is characteristic of all living and fossil nummulitids and gave rise to special Figure 6.31. The ranges of the Nummulitoidea in the Paleogene. The Cenozoic Larger Benthic Foraminifera: The Paleogene 457 lamination in the tests (Hottinger, 1977). The lamellar tests were formed during the process of chamber construction, in which each chamber wall covers the total test, including all former chambers (Hoheneger et al., 2001). Many authors have studied the morphology of the nummulitic tests, with some detailed descriptions as early as D’Archaic (1850), Carpenter (1850), Galloway (1933), Davies (1935) and Glaessner (1945). Among the more recent studies are those of Hottinger (1977, 1978), Hottinger et al. (2001), Racey (1995) and BouDagher- Fadel and Price (2014). The morphology of Nummulites of the family Nummulitidae is illus- trated in Figs 6.32 and 6.33. Figure 6.32. Test structure of Nummulites (megalospheric form, modified from Carpenter, 1850; Golev, 1961; and Barnett, 1969). Abbreviations:  A  =  aperture; AP  =  alar prolongation of lumen of chambers; AX = axial plug; CH = chambers; CS = canal system in marginal cord and septa; F = filaments; G = gran- ules on and between filaments; IP = interseptal pillars; M = marginal cord; P = proloculus; PP = coalescing granules forming polar pustule; S = septum; SP = septal pillars. Figure 6.33. The axial and equatorial sections of Nummulites (after Blondeau, 1972). 458 Evolution and Geological Significance of Larger Benthic Foraminifera Cole (1964) noted that certain features visible in axial section, such as the alar prolongations (chevron- shaped cavities on each side of the test) in involute forms, such as Nummulites, and the presence of thick lamellar walls on each side of the equatorial layer, as in Cycloclypeus, are useful in classification at the generic level. Hottinger (1977) proposed a generic classification based on the type of canal and stolon system, and the absence or presence of trabeculae (“imperforate shell material extending from an imperforate sutural zone into the perforate lateral chamber-w all and housing oblique, ramified trabecular canals opening between the pores on the surface of the lateral chamber wall” (Hottinger, 2006)). He considers the chamber formation and the type of stolon and canal system of greatest importance in the taxonomy of the nummulitic tests. Hottinger considered the stolon system and the canal structure as progressive, since they appear to become more complex with time. He distinguishes Heterocyclina (see Chapter 7) from Heterostegina on its L- shaped rather than Y- shaped stolons. However, Y- shaped stolons also occur in Spiroclypeus and Cycloclypeus (Haynes, 1981). Only the advanced Cycloclypeus has diagonal crossed stolons. Banner and Hodgkinson (1991, p. 105) defined the chamberlet as the division of the chamber as is seen in some species of Nummulites (e.g. Heterostegina (Vlerkina), etc.) and they called “cubicula” (sensu Banner and Hodgkinson, 1991) the separate little “chamberlets” that do not arise from the division of larger chambers and are called by others “lateral chamberlets”. This applies particularly to the lat- eral structures of orbitoids and miogypsinids. The median layer is composed of true chamberlets (which are known to be derived from the division of primary chambers phylogenetically or ontogenetically, or both). Within the family Nummulitidae, that includes the genera Nummulites and Assilina, a number of morphological evolutionary trends occurred (Racey, 1995; BouDagher- Fadel and Price, 2014), such as; • the degree of the involution or evolution of the test; • the degree of the development of the marginal cord; • the extent of the opening of the spire; • the division and subdivison of the chambers; • the overall size of the microspheric forms (i.e. the forms produced by sexual repro- duction), which increases over geological time for most lineages; • the size of the proloculus, which increases with time in the megalospheric forms (i.e. forms produced by asexual reproduction) for most lineages; • the development of embryonic chamber complexity tends to increase with time. Various biometric methods for studying the embryonic development of larger fora- minifera have been proposed. One of them (the “Factor E method” of van der Vlerk and Bannink, 1969), involves measuring the degree of enclosure of the second cham- ber by the third, to give a number referred to as Factor E.  It was used in studies reported by Racey (1992, 1995); • the shape of the septa (see Adams 1988; Racey 1995), as the septal complexity increases during time with: The Cenozoic Larger Benthic Foraminifera: The Paleogene 459 Figure 6.34. Range chart of some key Nummulites species with radial filaments. Nummulites: Filaments radial subreculate to reculate, becoming granulate Western Tethys Indo-Pacific "Le er Stages" Lower Te 23 L (Te 1-4) 22b Td 22a E Tc 21 20 L Tb 19 18 17 M Ta3 16 15 N. djokdjokartae 14 13 Nummulites uroniensis Nummulites laevigatus Nummulites benehamensis 12 11 E Ta2 10 9 8 Nummulites planulatus 7 6 5 L 4 3 Ta1 M 2 5 6 . 0 3 3 . 9 0 2 3 . 0 3 A g e ( M a ) E o c e n e O l i g o c e n e E p o c h S e l a n d i a n T h a n e t i a n Y p r e s i a n L u t e t i a n B a r t o n i a n P r i a b o n i a n R u p e l i a n C h a t t i a n S t a g e P l a n k t o n i c Z o n e S B - Z o n e s 460 Evolution and Geological Significance of Larger Benthic Foraminifera Figure 6.35. Range chart of some key Nummulites species with meandrine filaments. Nummulites: Filaments meandrine, not granulate, Septa meet roofs very acutely Western Tethys Indo-Pacific "Letter Stages" Lower Te 23 L (Te 1-4) 22b Td 22a E Tc 21 20 L Tb 19 18 17 M Ta3 N. javanus 16 Nummulites millecaput Nummulites gizehensis 15 14 Nummulites millecaput Nummulites gizehensis 13 12 11 E Nummulites irregularis Ta2 10 9 8 Nummulites irregularis 7 6 5 L 4 3 Ta1 M 2 5 6 . 0 3 3 . 9 0 2 3 . 0 3 A g e ( M a ) E o c e n e O l i g o c e n e E p o c h S e l a n d i a n T h a n e t i a n Y p r e s i a n L u t e t i a n B a r t o n i a n P r i a b o n i a n R u p e l i a n C h a t t i a n S t a g e P l a n k t o n i c Z o n e S B - Z o n e s The Cenozoic Larger Benthic Foraminifera: The Paleogene 461 Figure 6.36. Range chart of some key Nummulites species with sigmoidal to reticulate filaments. Nummulites: Filaments radial, becoming sigmoidal/ re culate, not granulate Western Tethys Indo-Pacific "Letter Stages" P22 Lower Te 23 L (Te 1-4) b Nummulites fichteli 22b P21 a Falciform N. fichteli Td 22a P20 filaments, E not twisted. SeptaP19 Tc 21 P18 N. vascus P17 Falciform 20 P16 filaments, L twisted at Tb 19 P15 poles. Septa N. hormoensis N. incrassatus 18 N. striatus P14 P13 17 N. striatus P12 N. pengaronensisM Ta3 16 P11 N. beaumonN. beaumon 15 14 P10 13 12 P9 Nummulites rotularius Nummulites rotularius 11 P8 Straight, radial E branching P7 10filaments; straight Ta2 septa N. globulus 9 P6 8 7 N. atacicus N. globulus Nummulites atacicus 6 P5 N. praecursor 5 L N. deser 4 P4 O. herber 3Ta1 M 2 P3 5 6 . 0 3 3 . 9 0 2 3 . 0 3 A e M a E o c e n e O l i g o c e n e E p o c h S e l a n d i a n T h a n e t i a n Y p r e s i a n L u t e t i a n B a r t o n i a n P r i a b o n i a n R u p e l i a n C h a t t i a n S t a g e P l a n k t o n i c Z o n e S B - Z o n e s 462 Evolution and Geological Significance of Larger Benthic Foraminifera o simple radiating (striate), falciform or curved septal filaments in Late Paleocene to Early Eocene species (see Fig. 6.34); o meandriform and complex branching in Middle Eocene forms (see Fig. 6.35); o reticulate septal filaments in Late Eocene forms (Fig. 6.36); o in the Early Oligocene the reticulate septa are still widespread (Fig. 6.36), but there is a tendency to the return to simple structures. Qualitative features such as the presence/ absence of granulation on the surface of the test and quantitative characteristics based on the morphometric system introduced by Drooger and Roelofsen (1982) (see Fig. 6.37) are commonly used in classifying num- mulitic species (see BouDagher- Fadel and Price, 2014, Fig. 6.37). They include: • P: the largest diameter of the proloculus in μm, excluding the thickness of the wall, • D: the largest diameter of the deuterconch in μm, excluding the thickness of the wall, • X: the number of the number of undivided, operculinid chambers before the appear- ance of the first partly divided planosteginid chamber, excluding the embryon (the first two chambers, protoconch and deuteroconch). This parameter indicates the degree of operculinid reduction. (see for example Fig. 6.37, in which X = 8), • l: the maximum diameter of the first whorl along the common symmetry axis of the embryon (including the protoconch and deuteroconch), • L: the maximum diameter of the first whorl and the subsequent half whorl as meas- ured in μm, along the common symmetry axis of the embryon (including the proto- conch and deuteroconch), • K: the index of spiral opening, where K = 100×(L–l )/ (L– P). 5 6 4 7 3 P D 8 2 1 l L Figure 6.37. Morphometric measurements in the equatorial section of megalospheric Planostegina africana BouDagher-F adel and Price (from BouDagher-F adel and Price, 2014). P = proloculus; D = deuteroconch; Pre- planosteginid chambers (X) are 8; l = the maximum diameter of the shell in the first whorl; L= the max- imum diameter of the first one and subsequent half whorl. The Cenozoic Larger Benthic Foraminifera: The Paleogene 463 In the American province, Sulcoperculina a proposed ancestor of the Paleogene nummulitoids (see BouDagher-F adel and Price, 2014)  died out at the top of the Maastrichtian. The first Paleogene nummulitoid may have evolved in the Americas from a pre-e xisting rotaliid form with a trochospiral test and intraseptal passages of a canal system, typified by a genus such as Pararotalia (see BouDagher- Fadel and Price, 2014; Fig. 6.38). Analyses of molecular data show a close relationship of nummulitids to the rotaliids (Holzmann et  al. 2003). However, the proposed process of evolving from a rotaliid such as Pararotalia would have been complicated and involved a ser- ies of coupled morphological changes. The trochospiral chambers of Pararotalia are arranged to expose the umbilical region that creates direct access to the ambient envi- ronment. The test possesses a spiral umbilical canal, formed by interconnected tooth- plates, with a free edge (Hottinger et  al. 1991). BouDagher- Fadel and Price (2014) suggested that the Nummulitidae may have evolved from a simple rotaliid test through forms which developed a moderately thick marginal cord and a rapidly widening coil leading to, in the Late Paleocene (P4a), the long-r anging completely planispiral evolute Operculina (see Fig. 6.38). The first occurrence of Chordoperculinoides bermudezi was recorded by BouDagher- Fadel and Price (2014) in the Middle Paleocene. This form is characterized by coarse vertical canals and a massive marginal cord, and is found throughout the Selandian (P3) and the Thanetian (P4). It gave rise to the sub-e volute Caudrina in the Late Thanetian (P5) of the American province (BouDagher-F adel and Price, 2014). Caudrina did not survive the Paleocene- Eocene boundary, while Chordoperculinoides died out at the top of the Early Eocene (end Ypresian, P9). Caudrina has not been recorded so far from the Eastern Hemisphere, but as BouDagher- Fadel and Price (2014) infered that other forms migrated eastward from the Americas, to populate the Eastern realms (see below). During the Thanetian (P4b), Operculina appears to have given rise to the evolute tightly coiled Assilina and the involute tightly coiled extant genus, Palaeonummulites. The latter had given rise to “lax” forms by the beginning of the Lutetian (P10), as seen in Operculinella. In a separate lineage, the Middle Eocene (P12b) also witnessed the development of an early involute test in some operculine species, which subsequently became fully involute, having chambers divided into chamberlets with secondary septa, giving rise to Heterostegina (see BouDagher- Fadel and Price, 2014). In the Eocene of the Western Tethyan province, Nummulites and Assilina witnessed a major radiation and increase in test size, which persisted up to the major extinc- tion of the last large species at the Middle- Late Eocene boundary. Few small species of Nummulites, and none of Assilina, survive into the Late Eocene, and Nummulites finally became extinct in the end of Early Oligocene. The Heterostegininae (Fig 6.38) evolved from the Nummulitidae in the Late Eocene and subsequently followed a clear morphological sequence of increasing involution of the test and the enclosure, by the alar (umbilical) prolongations, of the chambers. It also showed an increase in complexity of the chamberlet development (Banner and Hodgkinson, 1991). The pattern of increasing morphological complexity was repeated in many lineages from the Paleocene to the Early Miocene, as the more advanced forms appeared later in time. The flat evolute Operculina gave rise to the wholly evolute 464 Evolution and Geological Significance of Larger Benthic Foraminifera Nummulinae Cycloclypeidae Nummulidae Nummulites Pellaspiridae Middle Palaeocene - Early Oligocene Heterostegininae Cycloclypeus Early Oligocene - Holocene Palaeonummulites Late Paleocene - Holocene Pellatispira Spiroclypeus Late Middle Eocene - Late Eocene Late Eocene - Early Miocene Operculina Late Palaeocene - Holocene Heterostegina Latest Middle Eocene - Holocene ? ? Rotaliidae Pararotalia Late Cretaceous - Late Miocene Figure 6.38. The evolution of some nummulitid families from an rotaliid ancestor. The Cenozoic Larger Benthic Foraminifera: The Paleogene 465 Planostegina in the Late Paleocene, by developing partly developed chamberlets (Fig.  6.37). Planostegina persisted through the Neogene to Holocene. Heterostegina and Cycloclypeus evolved separately from Planostegina, the first descendant becom- ing initially involute, while the second lost its planispirality and became annular. This inferred sequence was corroborated by Holzmann et al. (2003) from their molecular genetic analyses. In the Late Eocene, the secondary chamberlets of Heterostegina sensu stricto became fully developed and the test became initially involute. This genus also per- sists into Holocene. In equatorial view the chambers are rectangular. However, the wholly involute Grzybowskia formed polygonal chamberlets in the Late Eocene. Heterostegina (Vlerkina) first appeared in the Late Eocene with a completely involute test and many rectangular lateral chamberlets. It disappeared within the Burdigalian (BouDagher- Fadel et  al., 2000, 2001). Also during the Late Oligocene and Early Miocene, Tansinhokella, a wholly involute genus, with abundant lateral chamber- lets and meandriform alar prolongations, became widespread. These lateral cham- bers proliferated and became polygonal (cubiculae, sensu Banner and Hodgkinson, 1991) and symmetrical on both sides of the spiral acquiring a sub-a nnular architec- ture, as in Spiroclypeus. In the Early Oligocene, the evolute Planostegina lost its planispirality and gave rise to Cycloclypeus, by becoming annular discoid with thick lamellar walls on each side of the equatorial layer. Cycloclypeus (Fig. 6.38) is a nummuliid with a nepionic morphology like Heterostegina, but with a final growth stage with cyclic chambers. In the course of the ontogeny of the megalospheric individuals, the early operculine chambers are reduced and the subdivided heterosteginid chambers get longer, with an increasing number of chamberlets, until they extend backwards over the previous part of the preceding whorl (see Fig. 6.38). The evolution of Cycloclypeus was stud- ied in great detail by Tan Sin Hok (1932). However, the potential for biostratigraphic application of his detailed biometrical descriptions is severely limited by the uneven distribution of suitable Cycloclypeus material over the geological column, and also by the complex nature of the data. The oldest species would appear to be C. koolho- veni (early Rupelian, Tc), possibly the direct ancestor of C. oppenoorthi (late Rupelian to early Chattian, Td) (see Fig. 6.39). It evolved gradually in the Oligocene to Early Miocene by reducing its nepionic whorl from two to a half, and the nepionic septa from 38 to just 2 stages in advanced forms such as C. cf. guembelianus and carpenteri (see Fig. 6.39). Over time, nepionic reduction in Cycloclypeus occured at a steady rate (Tan Sin Hok, 1932; MacGillavry, 1978; Haynes, 1981). However, it occurred within differ- ent lineages at different times. It should be noted that in general, the one Cycloclypeus lineage that could be continuous consists of species with very wide variations in inter- nal morphological characters, and that the individual species, if at all separable, are very long- ranging. Holocene forms of Cycloclypeus have been found to possess (X- shaped) stolons (Haynes, 1981). While Heterostegina and Cycloclypeus are still living, Spiroclypeus became extinct in the Early Miocene. The Pellatispiridae family (Figs 6.38 and 6.40) of the Nummulitoidea lived for a com- paratively short period at the end of the Eocene, during the Bartonian and Priabonian stages. They all have a planispiral-e volute test connected by a single foramen, with 466 Evolution and Geological Significance of Larger Benthic Foraminifera fns fns ns D P C A B anc anc fns anc fns ns P P D P D D D E F D P G H Figure  6.39. The variation in the (A)  microspheric and (B–G ) megalospheric embryonic and nepionic chambers of Cycloclypeus, modified after Tan Sin Hok (1932). Enlargement factors only refer to the nepionic stage. The last whorls shown are the first of the annular neanic stage. Here, P, stands for proloculus, and D, deuteroconch. The nepionic whorl is highlighted by a darker wall. Other abbreviation: anc, ‘‘ananepionic’’ chambers which are chambers of operculine shape, not subdivided into chamberlets; ns, nepionic (non- cyclic) septa; fns, first neanic (cyclic) septum. (A) An equatorial section of an early Cycloclypeus (Early to Middle Oligocene) with 2 anc followed by 27 heterostegine chambers, making a total of approximately 29 ns, e.g. C. koolhoveni Tan Sin Hok (with large central pillars) and C. oppenoorthi Tan Sin Hok (without central pillars). (B) C. eidae Tan Sin Hok with 1 anc and 18 ns (Late Oligocene to Early Miocene). (C) C. poste- idae Tan Sin Hok (Early Miocene) with 8 ns. (D) Cycloclypeus indopacificus Tan Sin Hok (Early to Middle Miocene) with 5 ns. (E) Cycloclypeus postindopacific Tan Sin Hok, 4 ns. (F) Advanced forms of Cycloclypeus, such as C. cf. guembelianus Brady with 3 ns. (G)  Enlargement of the embryonic apparatus to show the development of the radial canal system into marginal cord. (H) A microspheric Cycloclypeus with a small protoconch, 2 operculine stages (anc) with 11 nepionic stages (ns) followed by numerous cyclic chambers, Kalimantan, Early Miocene, Kalimantan. Scale bar = 1mm. The Cenozoic Larger Benthic Foraminifera: The Paleogene 467 interlocular spaces transformed into an enveloping canal system and flying covers of per- forate walls suspended on spikes or pillars (Hottinger et al., 2001). Pellatispira is con- sidered by Boussac (1906) and Umbgrove (1928) to have evolved from the thick-w alled assilid (planocamerinid) forms, by developing outer lamellae over the earliest chamber wall. However, the supplemental skeleton at the pellatispirid shell margin is not a marginal cord, as seen in nummulitids: it lacks the sulcus, and its canal system is radial and envelop- ing, not a tangential, polygonal network (Hottinger 1978, 2001). The work of Hottinger et al. (1991) clearly shows on the other hand, a close relationship between Pellatispira and Biplanispira, as emphasised by the adult of the former often forming the juvenile stage of the latter. He desribed the thickened shell margin produced by the marginal canal sytem in Pellatispira (Fig. 6.40) as being overgrown by supplemental chamberlets either on the lateral flanks alone (Biplanispira) or on all sides of the shell (Vacuolispira). The Orthophragminidae are bilamellar, perforate, orbitoidal larger benthic foramin- ifera, and are characterized by a discoidal, lenticular test with a fine equatorial layer and small lateral chamberlets (Fig. 6.41). The growth of orthophragminids was both cyclical and involute. Each chamber had a bilamellar structure with an inner lining and an outer lamella, with the subdivision of the equatorial chamberlets provided by dif- ferent inner linings (Ferràndez- Canãdell and Serra- Kiel, 1992; BouDagher- Fadel and Price, 2017). No canal system is evident (cf. the nummulitids), and connections between chambers are provided by a tridimensional stolon system (see Ferràndez- Canãdell and Serra- Kiel, 1992). Orthophragminids are classified based on the general shape of their tests, the pillar- lateral chamberlet network, the different kinds of stolons, and the size of their pillars. Moreover, their most important evolutionary parameters are associ- ated with the shape of the embryons in the megalospheric generations, the characters of nepionic stages in the microspheric orthophragminid juvenaria “their embryons”, Figure 6.40. A sketch, from Hottinger (2006), showing a spiral pellatispirid shell with an enveloping canal system and a marginal crest. According to Hottinger, the primary lateral chamber walls which are “emerg- ing” from the supplemental skeleton are covered by secondary lamellae but are perforated in continuation of the primary bilamellar wall and are not a part of the supplemental skeleton. The supplemental chamberlets are fed by canal orifices as they do not communicate directly with the spiral chambers by retral stolons. Abbreviations: ch = chamber; f = foramen; lh = loophole; mcr = marginal crest; pr = proloculus; schl = sup- plemental chamberlet; up = umbilical plate. 468 Evolution and Geological Significance of Larger Benthic Foraminifera and features of the equatorial chambers (Less, 1987; Brönnimann, 1951; Ferràndez- Cañadell and Serra- Kiel, 1992; BouDagher- Fadel, 2008; BouDagher- Fadel and Price, 2017). The Orthophragminidae are divided into two subfamilies, the Discocyclininae and the Orbitoclypeinae. Loeblich and Tappan (1987) considered Proporocyclina and Athecocyclina as synonyms of Pseudophragmina (see BouDagher- Fadel, 2008). However, BouDagher- Fadel and Price (2017) consider them to be different genera, as they exhibit stratigraphically characteristic, distinguishing morphological features. Dimorphism in the orthophragminids is common and is reflected in the size of the test, which is larger in sexually produced microspheric specimens than in the asexually produced megalospheric forms. The diameter of megalospheric specimens is ~3 mm or less, while microspheric specimens may reach >10 mm in diameter. As in occurrences of most larger benthic foraminifera, the megalospheric orthophragminids are more common than the microspheric ones (Ćosović and Drobne, 1995; Hottinger, 2001). The megalospheric forms of the discocyclinines have a subspherical protoconch enclosed by a larger reniform deuteroconch (Fig. 6.41, 1- 6). Their microspheric forms have a Cycloclypeus- like microspheric juvenile (see Fig. 6.41, 3- 9) with an initial spi- ral of small chambers. Later stages exhibit cyclical chambers (annuli) subdivided by septula into small rectangular chamberlets connected by annular and radial stolons. Externally, the test surface is either smooth, with scattered pillars, or has radially devel- oped ribs. An inflated central part or umbo (Fig.  6.41, 3- 4) is occasionally present. Adauxiliary chamberlets / Corona Deuteroconch Protoconch 1 2 Median/equatorial layer Lateral chamberlets 3 5 6 4 Megalospheric embryon Equatorial chamberlets Cycloclypeus-like intial spire of a microspheric form 7 8 9 Figure 6.41. The morphological features (from BouDagher-F adel and Price, 2017) of Discocyclininae. 1– 6 megalospheric forms: 1–3 highlighting the embryonic apparatus of the equatorial section, 1 schematic figure of the embryonic apparatus, 2 Discocyclina prattii (Michelin), France, 3 Discocyclina dispansa Sowerby, Tibet. 4–6 axial sections: 4 Schematic figure showing the three layers of the test, 5 Discocyclina sheppardi Barker, Soldado Rock, Trinidad, 6 Discocyclina sp., the Hecho Group, Ainsa Basin, south central Pyrenees, Spain. 7– 9 Microspheric forms: 7 schematic figure of the microspheric apparatus, 8 Discocyclina sp. France, 9 Oblique axial section of Discocyclina dispansa Sowerby, Tibet. Scale bars = 1mm, except for 1, scale bar = 0.25mm. The Cenozoic Larger Benthic Foraminifera: The Paleogene 469 Members of orbitoclypeines have an early planispiral microspheric coil, while meg- alospheric tests have a globular protoconch, enclosed by a larger reniform deutero- conch. Tests may occur with or without ribs. These taxa have a single equatorial layer of chamberlets and several layers of small lateral chamberlets, ranging from arcuate, spatulate to those hexagonal in shape, and cyclical chambers that are not subdivided into chamberlets. Biometric data are frequently used in specific definitions, and were first developed for Discocyclina by Neumann (1958), and subsequently expanded by Brolsma (1973), Fermont (1982), and Setiawa (1983). The degree of embryonic enclosure, the dimen- sions of the embryonic chambers, and the number of the periembryonic chambers have all been used as characteristic orthophragminid morphometrics (Brolsma, 1973; Fermont, 1982; Setiawan, 1983). More recently the morphological features of the whole group have been revised by authors such as Less (1987, 1998), Özcan et al. (2006) and BouDagher- Fadel and Price (2017). These authors focussed on internal features found in equatorial sections, and emphasised the description of the embryo (Fig 6.19), taking into consideration a number of parameters: • P1 and P2: the diameter of protoconch perpendicular and parallel to P-D  axis, • Dl and D2: diameter of deuteroconch perpendicular and parallel to P-D  axis, • A: the number of auxiliary chamberlets directly arising from the deuteroconch, • the number of adauxiliary chamberlets (see Fig. 6.41), or according to Hottinger (2006) corona - the first cycle of chamberlets enveloping an embryonal apparatus completely, at least in one plane of sectioning, as in Discocyclina, • the number of annuli within a 0.5mm wide stripe measured from the rim of deutero- conch along P- D axis, • H and W: the height and width of the equatorial chamberlets in the first annulus, • h and w: height and width of the equatorial chamberlets around the peripheral part of the equatorial layer. The development of annular stolons in the orthophragminids is deemed to be of generic significance (Samanta, 1967). Annular stolons occur at the prox- imal end of the radial walls and connect adjacent chamberlets in Discocyclina and Asterocyclina (BouDagher- Fadel, 2008), and at the distal end in Pseudophragmina and Asterophragmina (Haynes, 1981). The Discocyclininae have a Cycloclypeus-l ike microspheric juvenile (see Fig. 6.41) and might have evolved from an earlier nummulitid (Vaughan, 1945). However, many authors subsequently disputed the presence of true canals (Brönnimann, 1951; Samanta, 1967) but rather regarded them as the fissures of the Rotalioidea. The only reference to a primitive Discocyclina is that made by Caudri (1972), who describes a genus, Hexagonocyclina Caudri, 1944 with “two symmetric auxiliary chambers on each side of the nucleoconch and four spirals, and predominantly hexagonal equato- rial chamber”. However, the issue of the real ancestors of the orthophragmids is still not yet resolved as the earliest megalospheric or A-f orms, from the early Thanetian, are well developed and do not yield preorbitoidal stages (Less et al. 2007; Ferrandez- Caňadell and Serra- Kiel, 1992; BouDagher- Fadel and Price, 2017). 470 Evolution and Geological Significance of Larger Benthic Foraminifera Discocyclina was followed directly in the Late Paleocene (Thanetian) by Nemkovella, which differs from Discocyclina in lacking annular stolons. It only survived until the Early Eocene, while Discocyclina did not disappear before the Late Eocene. Parallel to this evolution was that of the Orbitoclypeinae, the first form of which to appear, in the Middle Paleocene, was the stellate Asterocyclina, which was directly followed by the non- ribbed Orbitoclypeus in the Late Paleocene. Both survived the Paleocene-E ocene boundary only to die out at the end of the Eocene. The study of the evolutionary lineages and stratigraphic distribution of the Tethyan orthophraminids, alveolinids and nummulitids led Serra-K iel et al. (1998) to propose 20 shallow benthic zones (SBZ) in Paleocene-E ocene time. Less (1998), and Less and Kovács (1995) documented the stratigraphic ranges of Tethyan orthophragminid fora- minifera with reference to these standard zones. They separated eighteen orthophrag- minid zones, marked OZ, by Less et al. (in press), from OZ 1a to 16, ranging from early Thanetian to late Priabonian. The Acervulinidae (Fig.  6.42) were another evolutionary lineage that occurred in the Eocene in parallel to the Planorbulinella lineage, that were based on globular and subglobular tests. In the Paleocene, small, spherical- concentric Sphaerogypsina evolved from a Cibicides-l ike ancestor by multiplying chambers in numerous layers (Fig. 6.42). The chamber arrangement strongly resembles that of Planorbulinella, but with the pat- tern being observed in every section irrespective of its direction. In the Middle Eocene, A lcht B eqch E a eqch lcht pax ech G pech cht C eqch f tlcht faix s pech F ech D lcht Figure 6.42. A-D sketches showing different Acervulinidae genera in axial sections, E- F in equatorial sec- tions, modified after Bursch (1947). (A) Planogypsina squamiformis Bermudez showing chambers added in a single layer (see Chapter 7). (B) Gypsina mastelensis Borsch with arcuate lateral chamberlets form on both sides of the equatorial layer. (C) Discogypsina discus Goes. (D) Discogypsina vesicularis Silvestri with truncate lateral chambers on one side and arcuate on the other side of the equatorial layer. (E) Sphaerogypsina globu- lus Reuss, with chambers aligned in successive layers. (F) The arrangement of the first-f ormed chambers in Discogypsina discus. (G) Alternating chamberlet arrangement three dimensions, producing a chessboard pat- tern, as in the spherical- concentric, globular shell of Sphaerogypsina globulus Reuss, after Hottinger (2006). The Cenozoic Larger Benthic Foraminifera: The Paleogene 471 a simple discoid, weakly symmetrical form evolved from a Cibicides-l ike ancestor into the three-l ayered Discogypsina (Fig. 6.41). In the Late Eocene, subglobular Gypsina developed a median layer, which can be recognized only in the central, earliest por- tion of the test, while a form similar to Sphaerogypsina, but with few spines and many pseudo- pillars, Wilfordia, is not yet known outside the type locality in Sarawak of Late Eocene age. The Asterigerinoidea originated from an Amphistegina- like form that itself was derived from a Cibicides-l ike ancestor in the Eocene, and evolved into the Lepidocyclinidae in the Middle Eocene in the Americas, but developed later in the Oligocene- Miocene in Tethys (see Fig. 6.43; BouDagher- Fadel, 2008; BouDagher-F adel and Price, 2010a). The Lepidocyclinidae comprises the subfamilies Helicolepidininae, in which adauxiliary chambers are absent and spirally arranged chambers completely sur- round the embryon, (e.g., Eulinderina, Helicolepidina, Helicostegina, Helicosteginopsis, Polylepidina), and the Lepidocyclininae, in which adauxiliary chambers are present and there is limited or no development of spirally arranged chambers (e.g., Astrolepidina, Eulepidina, Lepidocyclina). The origin of the Lepidocyclininae has been the subject of many studies (e.g. Barker and Grimsdale, 1936; Tan Sin Hok, 1936; Rutten, 1941; Hanzawa, 1964; 1965; Matsumaru, 1971, 1991; Frost and Langenheim, 1974; Sirotti, 1982; Butterlin, 1987; Drooger, 1993; BouDagher- Fadel and Price, 2010a; and others). They originated in the middle Eocene in the Americas (e.g., Butterlin, 1987), and gradually migrated westward, with their first appearances in the middle Eocene of West Africa, the early Oligocene in the Mediterranean part of the Tethys, and the latest early Oligocene in the Indo- Pacific (e.g., Adams, 1967; BouDagher- Fadel and Lord, 2000; BouDagher- Fadel and Price, 2010a). They reportedly went extinct in the middle Miocene (Serravallian) LEPIDOCYCLINIDAE Polylepidina with lateral chamberlets in the microspheric generation Eulinderina Late M. EoceneM. Eocene Lepidocyclina M.Eocene-L.Miocene Helicostegina M.-L.Eocene Helicolepidina M.-L.Eocene AMPHISTEGINIDAE Inner lamella Outer lamella . Figure 6.43. The origin and evolution of the lepidocyclinids. 472 Evolution and Geological Significance of Larger Benthic Foraminifera in Indonesia and the northern Indo-P acific (BouDagher- Fadel and Price, 2010a), but extended into the late Miocene (Tortonian) off Australia (Hallock and others, 2006). Reports of early Pliocene occurrences off Australia and the Pacific islands (Betzler, 1997; Adams et al., 1979) might represent reworked material and requires further bio- stratigraphic study to confirm their age (BouDagher- Fadel and Price, 2010a). Lepidocyclina evolved from an amphisteginid ancestor (Grimsdale, 1959), Eoconuloides via Helicostegina, Helicolepidinoides, Helicolepidina, Eulinderina and Polylepidina by simultaneously gradually decreasing the size and number of whorls of the eoconuloid stage while increasing the number of rows in the median chambers. In Helicostegina the annular growth is derived from the ecolunoid spiral, which is reduced in Eulinderina and completely missing in Polylepidina. The persistence of the spiral character in the parallel lineage of Helicolepidina is accompanied by differentiation of tb pg pi sp zn p fl mz d lc 1 fl 2 ec A 3 pp ec lc d d p pa lc ec B C Figure 6.44. (A) Three dimension cut away drawing of Lepidocyclina by Eames et al. (1968). (B) Equatorial section of a megalospheric Lepidocyclina. (C)  Axial section of a megalospheric Lepidocyclina. Abbreviations:  p  =  protoconch, d  =  deuteroconch, sp  =  stellate periphery, pp  =  polygonal periphery, ec =  equatorial chamber, lc =  lateral chamber, pi = pillar, fl = flange, tb =  tubercles, zn =  central zone, mz = marginal zone, pg = pillar forming an external tubercle, pa = auxiliary chambers. The Cenozoic Larger Benthic Foraminifera: The Paleogene 473 lateral and equatorial zones of chambers in both stocks. During ontogeny, the equato- rial chambers are mainly inside the ancestral eoconuloid wall, but later on this wall disappears and the genuine median layer is developed. The embryonic apparatus in the megalospheric generation of all lepidocycli- nids comprises the first and second chambers of the test (Fig. 6.44). These may be referred to as the proloculus and deuteroloculus when they form an integral part of the primary spire (e.g. Helicostegina gyralis, Eulinderina sp., see Fig. 6.43), and as the protoconch and deuteroconch when separated from the spire by a displaced third chamber (principal auxiliary) with two basal stolons. In this case, the second chamber (deuteroconch) does not form an integral part of the primary spire (e.g. Polylepidina chiapasensis). In all species of Helicostegina and Eulinderina, the deu- teroloculus tends to be smaller than the proloculus, and the third chamber smaller than subsequent spiral chambers (Adams, 1987). As in Discocyclina, the protoconch of the Lepidocyclinidae is always in the centre of the test. The embryonic apparatus consists of protoconch and deuteroconch, with a strong tendency for the deutero- conch to embrace the protoconch. Morphometric studies on the embryont of Lepidocyclina have been carried out by many taxonomists. Van der Vlerk (1959) proposed the parameters based on the dif- ference between the angle of attachment of the deuteroconch, expressed as an “alpha value”, and the percentage overlap, as the “Ai value”. DI and DII (Fig. 6.45) are the diameter of the protoconch and deuteroconch, measured at right angles to a line join- ing their centre (Haynes, 1981). Nepionic chambers are considered as auxiliary if their walls rest on the embryon, and interauxiliary if they are formed from the auxiliary chambers. Accessory chambers arising over the stolons in the deuteroconch are called adauxiliary, while those arising over the protoconch are the called protoconchal. The principle of nepionic acceleration seems to be the almost universal. Nepionic accel- eration is a general decrease over time in the number of steps in chamber formation, or budding steps, that is needed before orbitoidal growth is reached. It is seen in all known lineages of larger foraminifera with orbitoidal growth, and can also be inferred to have analogues in other larger foraminifera, e.g. Nummulites. Occasionally, lineages include discontinuities or sudden apparent relapses to more primitive forms. Examples of such relapses are the relatively primitive evolutionary stage, in terms of nepionic acceleration, of the first European Miogypsina (see below) compared with coeval and supposedly ancestral Miogypsinoides, and the sharp drop in embryon size from L. (Nephrolepidina) sumatrensis to N.  angulosa in the Indo- Pacific province. They can sometimes be explained by local environmental factors, or by invoking bottleneck events in which the populations are dramatically reduced and only less specialized and more primitive surviving forms provide the basis for further development. However, it may also be necessary to invoke local extinctions and re-c olonization by forms which evolved elsewhere, and reached different evolutionary stages, in order to explain these relapses. The stolon system of communication varies following the chamber shape. Arcuate chambers possess four stolons, while hexagonal chambers have up to six stolons. There have been many studies of European (Freudenthal, 1969; De Mulder, 1975) and Indo- Pacific (van Vessem, 1978) forms. However, variation of the stolon system is difficult 474 Evolution and Geological Significance of Larger Benthic Foraminifera Chambers ogival to elongate hegaxonal Adauxiliary chambers B d d DII p DI p Primary auxiliary chambers A (PAC) Protoconchal Figure 6.45. (A) Lepidocyclina (Nephrolepidina) morgani Lemoine and Douville, Formation bioclastique de Carry from the Nerthe area, near Marseille Petit Nid. (B)  Sketch showing the embryonic apparatus of L.(Nephrolepidina), modified after Haynes (1981). Abbreviations: p = protoconch, d = deuteroconch, DI = diameter of the protoconch, DII = diameter of the deuteroconch. to see in axial section and can vary in individual specimens, which prompted some authors such as Eames et al. (1962) and van Gorsel (1975, 1978) to doubt its taxonomic usefulness. Auxiliary chambers are primary auxiliary chambers (PAC) if they originate from stolons between the protoconch and the deuteroconch (Fig. 6.45). All subgenera of Lepidocyclina possess a complete ring of PAC, one on either side of the embryonic apparatus, and each PAC possess two stolons. L. (Nephrolepidina) develops one or more adauxiliary chambers, from a single stolon, on the deuteroconch and accessory chambers occasionally on the protoconch (protoconchal chambers). The biometric parameters of Lepidocyclina have been used by many scientists since van der Vlerk (1959) first introduced the factors “A” and “B” to characterize the nucle- oconch. The factor “A” measures the degree to which the protoconch is enclosed by the deuteroconch and the parameter “B” measures the extent to which the adauxil- iary chambers cover the circumference of the nucleoconch. The “B” parameter was later found to be inadequate to explain the morphometric variations, and hence was rejected (van der Vlerk, 1963). Other parameters were subsequently introduced to express the phylogenetic stage of the lepidocyclinids. These includes dc (degree of cur- vature of the common wall; van der Vlerk and Gloor, 1968), INT ratio (percentages The Cenozoic Larger Benthic Foraminifera: The Paleogene 475 of isolepidine- nephrolepidine- trybliolepidine types in an assemblage; van der Vlerk, 1973), E (degree of curvature of the embryonic chambers; Chaproniere, 1980) and F (form number; Chaproniere, 1980). Many of the lepidocyclinid taxa have been used by different authors in numerous ways, with different stratigraphic significance attached to each. This led to the reassess- ment of the genus- group names by BouDagher- Fadel and Banner (1997). These authors demonstrated that the subgenus Isolepidina Douvillé is a synonym of Lepidocyclina Gümbel, and that both are usable in the same and original sense, the latter being the senior name available for subgeneric use. They also neotypified the type species of the subgenus L. (Nephrolepidina) Douvillé, so that the subgenus can be used in the sense adopted by van der Vlerk and other Dutch workers. They proposed the abandonment of the name Trybliolepidina van der Vlerk, because a type species was subsequently designated making this genus a synonym of Eulepidina. (see BouDagher- Fadel and Banner, 1997). The characteristics intended for “Trybliolepidina” by its original author are more accurately characterized by the use of subspecific names in the nomenclature of each species with the appropriate nepionic morphology. Many authors, such as Lunt and Allan (2004), do not agree with removing the genus Trybliolepidina. Others such as Renema (2005), argues that in thin sections it is hard to distinguish between Lepidocyclina, Eulepidina and L. (Nephrolepidina) and a better way of citing the species is for example L. (Nephrolepidina)? rutteni, if the A-B species (megalospheric- microspheric) couple is known or “Lepidocyclinidae gen, indet.” if only the B- form (microspheric) of the species has been found. However, and as discussed by BouDagher-F adel and Banner (1997) the microspheric forms are subgenerically indis- tinguishable from the Eocene to the Middle Miocene, whilst the megalospheric forms are distinguishable subgenerically in a way which is stratigraphically valuable. It is true that in the same sample both megalospheric and microspheric forms are sometimes difficult to distinguish, but as L. (Nephrolepidina) is stratigraphically important, it is useful to differentiate between the two forms on a subgeneric level. The general ranges of the Asterigerinoidea are shown in Fig. 6.46. However, in many cases more detailed evolutionary trends have been determined. An example of the evolution of the American lepidocyclinids as developed by Barker and Grimsdale (1936) and others is given in (Fig. 6.47). In the Late Paleocene Eoconuloides lopeztrigoi evolved gradually from relatively simple to complex forms with many lineages in the Middle Eocene. The beginning of these lineages saw the appearance of Helicostegina gyralis, a primitive, single long spired form, equipped with counter septa, which gradually evolved into a shorter-s pire form with a complete flange of equatorial chamberlets in H. dimorpha. In the latter, the lateral chambers may appear in the last whorl, but a fully tiered arrangement is not attained. The lateral cham- bers in H. polygyralis were poorly developed and appeared only in later growth in H. paucip- ira, with fewer primary chambers in H. soldadensis. In the Middle Eocene, these forms gave rise to Eulinderina guayabalensis, by evolving more numerous arcuate equatorial chambers in a cyclical arrangement, with shorter primary coils and counter septa. Helicostegina poly- gyralis evolved into Helicolepidina spiralis by developing retrovert apertures, which in turn generated a second spire. This form evolved into H. trinitatis, which appears to be the most advanced member of the genus. It exhibits two subequal primary auxiliary chambers and 4 unequal periembryonic spires. According to Adams (1987) H. trinitatis indistinguishable from H. nortoni and of which it is a junior synonym. 476 Evolution and Geological Significance of Larger Benthic Foraminifera ASTERIGERINOIDEA Epoch - "Letter Stage Stages" P22 Lower L Te b P21 a Td P20 E P19 Tc P18 P17 L P16 Tb P15 P14 P13 P12 Ta3 M P11 P10 P9 P8 E P7 Ta2 P6 P5 L P4 M Ta1 P3 P2 E P1 Figure 6.46. The ranges of the Asterigerinoidea. 66.0 56.0 33.9 23.03 Age (Ma) Paleocene Eocene Oligocene Danian Selandian Thanetian Ypresian Lutetian Bartonian Priabonian Rupelian Chattian PLANKTONIC ZONES Eoconuloides Helicolepidinoides Amphistegina Helicolepidina Eulinderina Helicostegina Nephrolepidina Lepidocyclina sensu stricto Pseudolepidina Caudriella Polylepidina Helicosteginopsis Astrolepidina Eulepidina The Cenozoic Larger Benthic Foraminifera: The Paleogene 477 Figure  6.47. Evolution of the Asterigerinoidea during the Paleogene modified from BouDagher- Fadel (2008) and BouDagher- Fadel and Price (2010a). 478 Evolution and Geological Significance of Larger Benthic Foraminifera Polylepidina chiapasensis evolved in the Middle Eocene from Eulinderina guaya- balensis by increasing the number of embryonic spires from one to two (P. chiapasensis subplana) or more, and by developing two principal auxiliary chambers, with basal apertures and true lateral chambers. It gradually evolved from a 2- spiraled form to being a 4 spiralled one, and gave rise to the Lepidocyclina (Lepidocyclina) proteiformis group, by developing 4 equal spires, the 4 spires in Polylepidina being unequal (see Fig. 6.47). This group evolved into L. (Nephrolepidina) tournoueri (see Chapter 7) and, in the rest of the European and Indo-P acific, into the Lepidocyclina (Nephrolepidina) lineages. Quite independently, the Early Oligocene marked the first appearance of the earliest known species of Eulepidina (see Fig. 6.47) in America, e.g. E. favosa (see Chapter  7), which were the ancestors of the rest of the European and Indo- Pacific Eulepidina lineages (see BouDagher-F adel and Price, 2010a; Chart 7.1). Eulepidina is found throughout the Oligocene and ranges to the early Miocene, but they were most common during the early to middle Oligocene. In Eulepidina, the adauxillary chambers are particularly numerous, relatively small and irregularly distributed (Fig. 6.18). It is proposed that several lineages developed during the history of the evolution of Lepidocyclina s.l. (BouDagher- Fadel and Lord, 2000; BouDagher- Fadel and Price, 2010a), Miogypsina (BouDagher- Fadel et  al., 2000; BouDagher- Fadel and Price, 2013), Cycloclypeus and Austrotrillina, all of which have contrasting phylogenetic char- acteristics. In the early stages of evolution of Lepidocyclina s.l., the megalospheric and microspheric forms were similar in size and shape. However, as will be discussed in Chapter  7, in the Neogene Letter Stage interval late- Te to Tf3 there was a marked development of large microspheric forms, which bear few similarities with corre- sponding megalospheric forms that are often smaller and less distinctive. This evo- lutionary history curiously parallels that of the genus Nummulites. In the Early and Middle Eocene, the microspheric and megalospheric generations of Nummulites were similar in size and shape but in the Late Eocene to Early Oligocene specimens of the microspheric generation were very large in size and dominate their (often associated) megalospheric partners. A famous example of this the megalospheric Nummulites fich- teli and the microspheric Nummulites intermedius which often occur together but are thought to be the same biological species (see van der Vlerk, 1929, p.36, figs 30a-b , 31a,b; Eames et al., pl.1). The Rotalioidea (Fig. 6.48) can be traced back to Rotorbinella in the Late Cretaceous (see Hottinger, 2013), and possibly to a discorbid ancestor, such as Biapertobis (Fig. 6.49). The Rotaliidae, mainly with low-c onical tests (with the umbilical side cov- ered by numerous pillars) appear in the Early Paleocene, and include Dictyokathina, Lockhartia and Dictyoconoides in the Middle Eocene. On the other hand, the high conical Chapmaninidae appeared in the Late Paleocene and were common in the Late Paleogene (e.g. Angotia, Crespinina) (see Fig 6.17). The Miogypsinidae are the best known group of rotaliides foraminifera as they have been the subject of many micropaleontological investigations. Miogypsinids lived from Oligocene to Middle Miocene times, and form a largely coherent morphological and phylogenetical unit. As will be discussed in the next section, it seems that the miogy- psinid group shows distinct provincialism. Many independent lineages of Miogypsina have been recognised in the Mediterranean, Indo-P acific and central American bio- provinces (e.g., Van der Vlerk, 1966; Salmeron 1972; Raju, 1974; Raju and Meijer, 1977; The Cenozoic Larger Benthic Foraminifera: The Paleogene 479 ORBITOIDOIDEA - ROTALIOIDEA Epoch - "Letter Stage Stages" P22 L Lower Te b P21 a E Td P20 P19 Tc P18 P17 L P16 Tb P15 P14 P13 P12 Ta3 M P11 P10 P9 P8 E P7 Ta2 P6 P5 L P4 M Ta1 P3 P2 E P1 Figure 6.48. The ranges of the large rotaliides in the Paleogene. De Bock, 1976; Chaproniere, 1983, l984; Cahuzac, 1984; Ferrero, 1987; Wildenborg, 1991; Ferrero et al., 1994; Drooger, 1952, 1963, 1984, 1993; Drooger and Raju, 1973; Cahuzac and Poignant, 1991; BouDagher-F adel et  al., 2000; BouDagher-F adel and Price, 2010b, 2013). 66.0 56.0 33.90 23.03 Age (Ma) Paleocene Eocene Oligocene Danian Selandian Thanetian Ypresian Lutetian Bartonian Priabonian Rupelian Chattian PLANKTONIC ZONES Kathina Dictyokathina Lockhartia Sakesaria Daviesina Sherbonina Actinosiphon Neosivasella Elphidium Chapmanina Sirelella Dictyoconoides Pararotalia Ferayina Pellatispirella Angotia Crespinina Neorotalia Miogypsinoides Miogypsina Paleomiogypsina Miogypsinella Miogypsinita 480 Evolution and Geological Significance of Larger Benthic Foraminifera Dictyoconoides Middle Eocene Sakesaria Paleocene - Early Eocene Lockhartia Paleocene - Middle Eocene Rotorbinella (Coniacian - Early Paleocene) Biapertorbis (Turonian - Eocene) Figure 6.49. The evolution of the conical Dictyoconoides from a rotaliid ancestor. It is generally accepted that the group originated from a trochoid ancestor. In the Americas, miogypsinids appear in the in latest Early Oligocene (Rupelian) (BouDagher-F adel and Price, 2010b), while in Europe, Miogypsina spp. first appears in the Aquitanian (Drooger, 1963) and is not known in the underlying Oligocene. On the other hand, Miogypsinoides evolved independently from Miogypsinella in the Far and Middle East (BouDagher-F adel and Banner, 1999) at the base of the Late Te Letter Stage. Many of the ranges presented in this and the subsequent chapter for genera occurring the Far East and in the American provinces have been revised and modified by BouDagher- Fadel and Price (2010b, 2013) Tan Sin Hok (1936) described the evolution of the Miogypsinidae through Miogypsinoides, in the Early Miocene of Indonesia (then Netherlands, East Indies). The Cenozoic Larger Benthic Foraminifera: The Paleogene 481 He traced the morphological development on the basis of the megalospheric nepionic chamber arrangement, but he also used a typological species concept in his classifi- cation. This typological species concept was later used by other workers such as Cole (1957). From the study of Central American collections, Barker and Grimsdale (1931) proposed that Neorotalia was the root stock for all the Miogypsinidae. This evolution was demonstrated by them from the presence of a septal canal in the most primitive Miogypsinoides. Other authors have offered different opinions about the transition from Miogypsinoides to Miogypsina (see Drooger 1993, p.75- 76). Drooger (1952) in his thesis on American Miogypsinidae introduced the population concept for classification purposes, that is to say measurements of numerous specimens were used to define taxa. Since then he has performed many biometric studies on the Miogypsina nepiont, as seen in equatorial section. The biometric method appeared to be a useful tool in the study of the evolution of a single lineage of Miogypsina and of miogypsinid genera in general (see Fig. 6.50). Indices “X” and “V” are used as the primary characteristics for subdivision, where X is the total number of spirally coiled nepionic chambers, once a second spiral appears, while the biometric factor V, which is equal to 200α/ β, is the degree of symmetry between the two spirals (see Fig. 6.50). In Miogypsina, the biometric factor V is low in older species, and high in the nearly symmetrical embryon of the younger species (Drooger, 1993). In the Miogypsinoides sequence of assemblage-s pecies in the Indo-P acific (M.  complanatus - formosensis - bantamensis - dehaarti – indicus) there is an evolutionary lineage involving the reduc- tion of the nepionic spiral, and the increase in size of the protoconch (Drooger, 1993; see also Chart 7.1). Drooger (1993) set numerical limits between the X values to define the species, thus:  complanata - 17 - formosensis - 13 - bantamensis - 10 - dehaarti - 8 - indica. The biometric factor “Y” is the number of spiral chambers up to the first chamber with two foramen or stolons (see Fig. 6.50), while the factor “γ” is angle of deviation of the line of symmetry of the embryont from the apical line, which is the line of sym- metry of the adult test. For long- spiraled juveniles this value is considered as a negative number and positive when there is less than one juvenile whorl, and there is a second nepionic spiral. De Bock (1976) and Drooger (1993) found that Miogypsinoides pos- sesses a well- developed canal system, whereas in Miogypsina such a system is lacking. On the other hand, Miogypsina was believed to develop a more complex stolon system than in Miogypsinoides. Matsumaru (1996) in his studies of Japanese Paleogene foraminifera, recognised the Late Oligocene origin of Miogypsinella from his new genus Paleomiogypsina in the late Chattian, which he believed to be a descendant of Pararotalia (Matsumaru, 1996, p. 44). However, in common with the conclusions of Barker and Grimsdale (1931), BouDagher Fadel et al. (2000) traced the origin of the Miogypsinidae to the genus Neorotalia, and not to Pararotalia nor to Rotalia and its relatives (e.g. Medocia), which have a different umbilical structure and apertural system. Pararotalia and Neorotalia have very different apertural systems from Rotalia, as discussed by Loeblich and Tappan (1988, p. 659), however, Pararotalia and Neorotalia are not synonymous as suggested by these authors. Pararotalia first occurs in the Late Cretaceous and gives rise to Neorotalia in the Oligocene. They both have a rel- atively low trochospire, but Pararotalia is more strongly trochospiral and smooth, 482 Evolution and Geological Significance of Larger Benthic Foraminifera γ γ18 p p Stolon A B γ γ ax d d ax α ax p ax p β C D Figure 6.50. (A). A sketch showing the earliest chamber of an older Miogypsinoides, such as Miogypsinoides complanatus. There are 18 spirally coiled nepionic chambers (biometric factor X). There are 10 spiral chambers before the first chamber with 2 stolons (biometric factor Y). The angle of rotation between the apical line of symmetry and the central line crossing the protoconch and deuteroconch is around - 330o. (B) Miogypsinoides bantamensis Tan Sin Hok from Castlesardo section, northern Sardinia, with a biometric factor of X = 10 and a γ of approximately - 290o. (C) A sketch of the early chambers of a young Miogypsina, such as M. cushmani, with 2 auxiliary chambers (ax) with a biometric factor V (= 200α/ β), where γ, the positive angle of deviation of the line of symmetry of the adult test, is approximately 78o. (D) Miogypsina cushmani Vaughan, from Castlesardo section, northern Sardinia. while Neorotalia has a much lower trochospiral form and is pillared both dorsally and ventrally. Neorotalia, while sharing a basic architecture with Pararotalia, also differs from the latter by its enveloping canal system (Hottinger et al., l99l). After Neorotalia acquired its dorsal, ventral and peripheral pillars, it evolved into the miogypsinid line- age by the acquisition of an additional fan of small chamberlets in a median equatorial layer (see Fig. 6.51). Paleomiogypsina evolved from Neorotalia in the Late Oligocene, by acquiring little chamberlets scattered on the periphery of the last whorl (Matsumaru, 1996, p. 54). Miogypsinella evolved into Miogypsinoides (with lateral thickening, no cubiculae, and a planispiral embryont) in the Late Te stage (BouDagher- Fadel and Banner, 1999) with the initial coil becoming planispiral, and the lateral walls of the initial spire and the succeeding fan of ogival median chamberlets becoming very thick The Cenozoic Larger Benthic Foraminifera: The Paleogene 483 and solid (see Fig. 6.51). In forms such as Miogypsinoides abunensis pillars continue on both sides of the test. In Miogypsinoides dehaarti the proloculus is followed by a single whorl of auxiliary chambers (Van der Vlerk, 1966) and the pillars are lost. The evolution of the miogypsinids started with a trans-A tlantic migration of Neorotalia from the Americas, where miogypsinids originated. The eastward migration followed two paths: one to the south towards South Africa, where a distinct phylogenetic lineage, but similar to that found in America, developed but went extinct in the Burdigalian; the other to the north, through the Mediterranean corridor. During the Chattian and Aquitanian significant miogypsinid forms evolved in the Mediterranean from the morphologically distinct Mediterranean Neorotalia and migrated, within a few million years of their first appearance, into the Indo- Pacific, where they diversified further. The extinction of 1' 2' 3' 1'' 2'' 3'' 1''' 2''' 3''' 4' 5' 4'' 5'' 5''' 4''' 5'''' Figure 6.51. Silhouette diagrams of stages in the evolution of the Miogypsinids in the Far East, modi- fied from BouDagher- Fadel and Lord, 2000. (1) Paleomyogypsina. (2) Miogypsinella. (3) Miogypsinoides. (4) Miogypsinodella, (5) Miogypsina. 5’’ early Miogypsina, 5’’’ advanced Miogypsina, e.g. M. indonesiensis Tan Sin Hok. 1’ – 5’ Direction and approximate extent of chambers coiled around the proloculus. 1’ – 5’’, 5’’’ Equatorial extent of chambers coiled around the proloculus. 1’’’ – 4’’’, 5’’’’ Axial diagram showing form of the test. 484 Evolution and Geological Significance of Larger Benthic Foraminifera Mediterranean miogypsinids went extinct in the Langhian, but miogypsinids survived in the Indo-P acific into the Serravallian. This miogypsinids exhibit an example of ‘par- allel speciation’ as discussed by Schluter et al. (2004). As species became geographically isolated, colonizing new areas environmentally similar to each other, they thrived and evolved similar but distinct parallel lineages, taking advantages of empty niches and opti- mum conditions. This process probably reflects that they all shared a genetic predispo- sition to develop mutations of a specific, advantageous type, inherited from their last common ancestor (see BouDagher- Fadel and Price, 2010b, 2013). 6.4 Palaeoecology of the Paleogene Larger Foraminifera The Paleogene is a transitional time during which the Earth moved from a uniformly warm Cretaceous to a cooler, more climatically heterogeneous Neogene (Berggren and Prothero, 1992). The Early Paleocene was characterized by warm, generally ice- free conditions, but slightly cooler than the preceding Cretaceous. However, temperature rose again in the Late Paleocene with an anomalously warm global climate optimum spanning some 4- 5 Ma during the Early Eocene (Prothero and Berggren, 1992). In this brief period of extreme warming at the onset of the Eocene, called the Paleocene- Eocene Thermal Maximum (PETM) (Cramer and Kent, 2005), climates became hot and arid north and south of the Equator. The PETM has been attributed to a sudden release of carbon dioxide and/o r methane and coincided with a major perturbation Figure 6.52. Climate zones in the Middle Eocene. The Cenozoic Larger Benthic Foraminifera: The Paleogene 485 of the carbon cycle as indicated by a sharp negative δ13C excursion (CIE) (Dickens, 2001). This warm period continued through the Eocene (see Fig. 6.52), where a pole- ward expansion of coral reefs occurred, together with a broader latitudinal distribution of temperature- sensitive organisms, such as larger benthic foraminifera, mangroves, palms, and reptiles (Adams et al., 1990; Kiessling, 2002; Pearson et al., 2007). Coral reefs were able to grow up to 46oN, in comparison to today’s coral-r eef distribution that reaches to only 34oN (Kiessling, 2002). Like most benthic fossils, the larger benthic foraminifera of the Paleogene are biofa- cies bound, sometimes even on a regional scale. They have biotopes closely associated with carbonate environments that represent a variety of palaeoenvironments, ranging from backreef to open marine conditions. Large scale changes in these biotopes occur in response to eustatic sea level fluctuations. Larger foraminifera are extreme K- strategists (characterized by long individual lives and low reproductive potential, see Chapter 1) thriving in a stable, typically oligotro- phic environment (Hottinger, 1983). Although all larger foraminifera are considered K- strategists, Paleogene larger foraminifera exhibit an increasing trend in K- strategy, from alveolinids to nummulitids to discocyclinids (Hottinger, 1982). In the Cretaceous, a superheated (Fig. 6.53), hypersaline ocean-c limate zone favoured the proliferation of rudists over corals (Kauffman and Johnson, 1988; Johnson et al., 1996). Following the K-P event, corals appear to have been rare and of low diversity (Johnson et al., 1996). The warm sea- surface temperatures and enhanced CO2 levels of the Early Paleocene, would have caused symbiont loss and bleaching (Gattuso and Buddemeier, 2000), and thus prohibited the expansion of reef- building corals in the lower latitudes (Sheppard, 2003). The Early Paleocene is commonly thought to have been a time of reorganization of reefs and their communities (Newell, 1971; Talent, Variation of Sea Level and Temperature in the Paleogene 200 8 6 160 4 120 2 0 80 –2 40 –4 –6 0 20 25 30 35 40 45 50 55 60 65 70 –8 Oligocene Eocene Paleocene –40 –10 Age (Ma) Figure 6.53. Variation in sea-l evel and temperature during the Paleogene based on Zachos et al. (2001) and Miller et al. (2011). Sea Level Change(m) Temp Anomaly (degreeC) 486 Evolution and Geological Significance of Larger Benthic Foraminifera 1988; Copper, 1989), and it was not until the late Thanetian that coral- algal patch-r eefs became more diverse. Following the K- P event, larger benthic foraminifera began to stage a recovery in the Early Paleocene, as unlike reef-b uilding corals, rising summer sea-s urface tem- peratures do not cause symbiont loss in their test (Hallock, 2000). They became the most common constituents of Late Paleocene- Early Eocene carbonate platforms, fill- ing the empty niches left by the decline of the Cretaceous rudist-c oral assemblages. They thrived on shallow, oligotrophic, circum- Tethyan carbonate platforms (Buxton and Pedley, 1989). By the Late Paleocene, nummulitids (Nummulites, Assilina and Operculina), orthophragminids (Discocyclina) and alveolinids (Alveolina) became important components of the carbonate Tethyan platforms. Their distributional pat- tern in Paleogene microfacies indicate different ecological gradients based on different faunal associations (Fig. 6.54), which in turn influenced lamellar thickness and flatten- ing of larger foraminifera tests (Ćosović et al., 2004). Although the Late Paleocene witnessed the beginning of the diversification of the num- mulitids, it was during the Eocene that the Nummulites fulfilled their unique rock-f orming potential. They became abundant and formed the widespread nummulitic limestones of hydrocarbon reservoirs in offshore North Africa, India, and the Middle East. Their reservoir qualities are mostly due to the preservation of the intraskeletal porosity of the Nummulites test. Various depositional models have been proposed, and most of them described Nummulites accumulations as banks, bars or low- relief banks, sometimes related to palaeo- highs. The behaviour of Nummulites could explain the diversity of such deposi- tional models. Depending on local hydrodynamic conditions, autochthonous Nummulites deposits can be preserved as in situ winnowed bioaccumulations or can be accumulated offshore, onshore or alongshore, away from the original biotope (Jorry et al., 2006). Figure 6.54. The distribution of Paleogene larger benthic foraminiferal taxa on the Tethyan carbonate shelf. The Cenozoic Larger Benthic Foraminifera: The Paleogene 487 The nummulitids included genera with flattened to stoutly lenticular and even glob- ular species, with a periphery varying from sharp to rounded or somewhat undulose (Beavington- Penney and Racey, 2004). The globose to ellipsoidal-flat Nummulites thrived in warm, tropical mesophotic zone, while the large flat forms, such as N. mil- lecaput are common in the oligophotic zone, either in shaded shallow water zones or deeper on the shelf (Mateu-Vicens et al., 2012, Pomar et al., 2017). Of the nummulitids, Operculina is the most “primitive” genus known and the least specialised (Cole, 1957; Chaproniere, 1975). For this reason it is thought to be less dependant on coralline algae symbionts for food, and therefore it is inferred to have had a wider environmental range than the other nummulitids (Cole, 1957; Adams, 1965). Although the living nummulitids house symbiotic microalgae, they prefer calm water conditions and avoid highly illumi- nated areas near the water surface, since their flat tests could easily be damaged in a tur- bulent hydrodynamic regime (Hohenegger et al., 2000). The living nummuliid Operculina appears to be restricted by oceanic salinities (Chaproniere, 1975). It has been found at depths as shallow as 14m (Hottinger, 1983). It has also been reported from the quieter parts of lagoons (McKee et al., 1959), channels on reef flats (Maxwell et al., 1961) and in off- reef shelf areas (Maxwell, 1968). Operculina lives on soft- bottomed substrates in the Gulf of Aqaba at depths of 30-1 50m, with flatter forms most common between 60- 120m (Reiss and Hottinger, 1984). It dominates low-e nergy, muddy seabeds (Banner and Hodgkinson, 1991). In the Oligo-M iocene this genus is inferred to have inhabited environments ranging from high energy, shallow water forereef facies (BouDagher-F adel et al., 2000) to quiet waters near the base of the photic zone (Chaproniere, 1975). Similarly, Cycloclypeus has occupied a broader depth range in the past than at pres- ent. Modern species of Cycloclypeus are believed to live in deeper waters of 70-1 30m, down to the lower limit of the photic zone, crawling on firm substrates (Hottinger, 1983; Reiss and Hottinger, 1984; Hohenegger et al., 2000; Hohenegger and Yordanova, 2001; Yordanova and Hohenegger, 2002). They therefore tolerate lower light levels and temperatures than most other larger foraminifera (Cole, 1957). During the Oligo- Miocene, Cycloclypeus may have had a depth range of <12m to sub- euphotic depths (BouDagher- Fadel et al., 2000), and only later became restricted to the deeper range (Chaproniere, 1975). Thin, flat forms of larger benthic foraminifera are thought to be an adaptation to light attenuation with increasing habitat depth (Hallock and Schlager, 1986). In the Oligo- Miocene, Heterostegina inhabited high energy forereef environ- ments (BouDagher- Fadel et al., 2000), of 20 to 30 m depth (Banner and Hodgkinson, 1991), preferring to live on hard substrates (Reiss and Hottinger, 1984). Holocene Amphistegina has adapted to high energy conditions, however, it is also found in mud- free sands in areas of sea grass or coralline algae and in reefal areas down to depths of 35m (McKee et al, 1959). Dead tests of Amphistegina have been found at greater depths (Chaproniere, 1975), but its main depth range is reported as 5 to 20m (Murray, 1973). Photoautotrophic symbionts are the only food source for larger benthic foraminifera (Leutenegger, 1984; Krüger, 1994; Hohenegger, 2004) and provide the potential for calci- fication of large skeletons (Hallock et al., 1991). Eocene Nummulites has no counterpart in present- day protist groups (Jorry et al., 2006). The diameter of the largest Operculinella venosus microsphere is 6.4 mm, and 3.2 mm for the largest macrosphere (Hohenegger et al., 2000), whereas fossil Nummulites often reach several centimetres in diameter. Except for Cycloclypeus carpenteri, whose maximum observed diameter is 120 mm (Hohenegger 488 Evolution and Geological Significance of Larger Benthic Foraminifera et al., 2000), the largest size reported (Nemkov, 1962) is of a Mesopotamian specimen of Nummulites millecaput that reached 160 mm in diameter. Such gigantism is considered by Cowen (1983) as proof of an active algal symbiosis, which is also supported by the pres- ence of microstructures similar to those observed in present- day larger benthic foramin- ifera, which provide shelter for symbionts and allow respiration (Bartholdy, 2002). In deducing the palaeoecological distribution of the Orthophragminidae (such as Discocyclina and Asterocyclina), Ćosović et  al. (2004) regarded these elongate, thin, flat to biconvex forms as homeomorphs to the Holocene Cycloclypeus and Baculogypsinoides (see Chapter 7). The orthophragminids at the lower limit of the photic zone are very flat, and their lateral chamberlets are particularly low in shape. The Baculogypsinoides individuals, with 3– 4 strong spines, are similar to the genus Asterocyclina. They live on coral rubble and are extremely rare on sandy bottoms in Okinawa (Hohenegger, 2000; Hohenegger et al., 2000; Hohenegger and Yordanova, 2001; Yordanova and Hohenegger 2002) and the Spermonde Archipelago, Indonesia (Renema and Troelstra, 2001). Therefore, the homeomorph asterocyclinids with five to six rays might have lived on firm substrates in high energy environments, along with asterigerinids and amphisteginids, at depths of less than 30 m (Hallock, 1999; Ćosović et al., 2004). Smaller, globular nummulitid specimens are morphologically similar to the Holocene Operculinella venosus Fichtel and Moll, with thick lenticular tests, which inhabits sandy bottoms at moderate depths in the euphotic zone (40– 80 m). The division of the median chambers in Cycloclypeus has evolutionary parallels with the division of the cubiculae (or small chamberlets) of Miogypsina, where the illuminated chambers are also nests for diatoms (see BouDagher- Fadel et al., 2000). Each of the chambers cubiculae would act not only as a small convex lens for the focusing of sun- light, but could also act as a greenhouse for the containment and development of symbi- otic diatoms. The diatoms enable these forms to acquire nutrients without food-g athering pseudopodial activity (Röttger, 1971). Although many small rotaliine foraminifera (e.g. Rotalia, Rosalina, etc.) gather food particles to their umbilici for extrathalamous diges- tion, larger foraminifera with cubiculae, but no umbilici (e.g. Miogypsinella ubaghsi, see Chapter 7), must have had a different method of ingestion. In the Miogypsinidae several lineages evolved (BouDagher-F adel et al., 2002) which followed similar evolutionary patterns, all related to the utilization of radiant light in the promotion of populations of photosymbionts. In Miogypsinella the flange of median chamberlets is relatively small (not much broader than the coil of the nepiont); in Miogypsinoides the flange grew very much broader, and would have become heavily thickened laterally. These lateral walls, becoming very thick, would have reduced the amount of sunlight which could reach the median layered chamberlets, so splitting into many layers of convex cubiculae in Miogypsina became competitively advantageous. Although diatom endosymbionts may initially enter the foraminifera as ingested food, they establish themselves very quickly as a permanent resident population inside the cell (Richardson, 2001). It is probable that Miogypsinella had active, food-g athering pseudopodia, however, it is likely that Miogypsina had pseudopodia which were of little effect in food gathering and instead used its cubiculae as an arable farm. The cubiculae are not linked by apertures, instead perforations occur in the walls. These perforations would be too fine to allow the passage of diatoms, which cubicular apertures would have allowed to escape. The requirement for and use of sunlight by Miogypsina is shown by The Cenozoic Larger Benthic Foraminifera: The Paleogene 489 the common occurrence of Miogypsina spp. only in shallow water marine limestones, where fossil algae also occur. The irregular shape of many species of Miogypsina, often leading in section to apparent division of their flanges (e.g. Miogypsina bifida, Plate 7.12, fig.14), shows they could become concavo-c onvex and were not growing on a flat surface. Such a shape would be developed if the species were attached on a strongly curved substrate, such as the stems or leaves of seagrass (the substrate would be bio- degradable and seagrass is not seen preserved in thin sections). On such an irregular substrate surface it would be difficult for any organism to remain fully attached and yet migrate on that surface. Therefore we believe that sedentary, attached miogypsinids grew to accommodate the shape of the vegetable substrate to which they adhered. Only in strong ambient sunlight, which would benefit both the miogypsinids and their veg- etable substrate, could true Miogypsina flourish (BouDagher- Fadel et al., 2000). The lineages of the lepidocyclinids evolved gradually, and newly evolved taxa persisted to become contemporaneous with their descendants, but occupied differ- ent ecological niches. They were common throughout warm-w ater, shallow- marine environments of the tropics. It is almost certain that their latitudinal distribution was controlled by temperature, as the 150C isotherm of the colder months limits the mod- ern geographic distribution of Rotaliida that host algal endosymbionts (Langer and Hottinger, 2000; BouDagher-F adel and Price, 2010a). An example of evolutionary changes influenced by ecological parameters is given by forms such as L. (N.) fer- reroi (Plate 7.8, fig.  1; see BouDagher- Fadel et  al., 2001), in which the massive pil- lars between lateral chamberlets may be thought to be diagnostic of living in high energy marine environments (BouDagher-F adel et al., 2000). However, when pillars are growing on the lateral walls, as is the case in Discocyclina, orbitoids, lepidorbitoids, some lepidocyclinids, as well as in evolute nummulitids, they might act as true lenses, which focus light into the test for insolation of symbionts (Ferrancez-C aňadell and Serra- Kiel, 1992). On the other hand, this function is performed by thin lateral shell walls in some nummulitids (Hottinger, 1983), such as Operculina (Drooger, 1983) and Cycloclypeus (Reiss and Hottinger, 1984). The miliolide genus Austrotrillina has a distinctly alveolar wall, again presumably to harbour symbiotic algae. During the Oligocene and Early Miocene, Austrotrillina was associated with alveolinids, such as Borelis pygmaeus, in the very low hydrodynamic- energy back- reefs (BouDagher- Fadel et al., 2000; see Fig. 6.54). The biological advan- tage of possessing an alveolar wall with increasing complexity of chamber division is not only found in Paleogene taxa, it is a well-e stablished evolutionary trend of alveo- linid foraminifera from the Carboniferous onwards. It may be analogous in function to the increased complexity of the Austrotrillina wall, probably increasing the efficiency of symbiosis with algae and diatoms. Recent alveolinids occur in a wide range of habitats, from deep lagoons and to fore- reef settings, existing down to depths of about 80m. This, together with the fact that alveolinids are miliolines with a tolerance to salinity and temperature fluctuations, prob- ably makes the group less sensitive to smaller sea level changes. The Eocene alveolinids became extinct, however, at the onset of Late Eocene rapid sea level changes, which led to the disappearance of vast carbonate platform and lagoonal areas. Oligocene and Miocene alveolinids also had a wide range of habitats, but were particularly common 490 Evolution and Geological Significance of Larger Benthic Foraminifera in deeper lagoonal settings. Their response to sea level changes might therefore be simi- lar to that of other miliolides, were it not that extinctions during sea level falls are unlikely due to the group’s extensive ecological range. Obliteration of the lagoonal habitat during sea level falls may have caused a reduction in population sizes but trig- gered an increase in the rate of evolution. Extant species of soritids live in seagrass communities in tropical-s ubtropical, shal- low water habitats, and are characterized by the possession of intracellular symbionts, such as rhodophyte, chlorophyte, or dinophyte photosymbionts (Lee and Anderson, 1991; Gast and Caron, 1996; Hallock, 1999). Species that host chlorophyte symbionts are found in relatively shallower waters (0- 20m) than those that host diatom endo- symbionts, which for example are found in a wider range of depth habitats (0-1 30m) (Leutenegger, 1984). According to Richardson (2001), the acquisition of dinophyte endosymbionts in the soritids facilitated a change in habitat from an epifaunal, free- living mode of life to one of living attached to phytal substrata. Larger benthic foraminifera showed a diversification at a specific level, i.e. involv- ing a rapid increase in species diversity, shell size, and adult dimorphism, during the Paleocene- Eocene transition (Hottinger, 1998). Fossil Nummulites are known to have developed a large range of shapes induced by reproductive strategies (small sexual A- forms and large asexual B- forms, see Chapter  1) and by environmental factors (light intensity and hydrodynamic conditions), which significantly affect the size, shape and thickness of the tests (Hallock 1979, 1981; Hallock and Glenn, 1986; Hohenegger et al., 2000, Ćosović et al., 2004). The hydrodynamic behaviour of Nummulites, which depends on their size, shape and density, is a fundamental parameter controlling their transport. According to Beavington-P enney and Racey (2004), foraminifera that live in shallow water produce ‘robust’, ovate tests with thick walls (e.g. Orbitoclypeus) to prevent photoinhibition of symbiotic algae within the test in bright sunlight, and/ or to prevent test damage in turbulent water. However, there is a tendency towards flatter tests and thinner test walls with increasing water depth (e.g. Discocyclina). The change in the test shape reflects decreased light levels at greater depths. There is also a relationship between size, longevity and fecundity (Hallock 1985). It is known that larger foraminifera living in fairly turbulent waters become relatively large, with a thickness to diameter ratio of 0.6 to 0.7 (Ćosović et al., 2004). They produce up to ten times more offspring per reproduction than deeper dwelling species, with the ratio between juvenile and adult specimens also reflecting this phenomenon. Racey (2001) plotted the relationship between num- mulitid accumulations and hydraulic energy; shallowing causes a progressive con- centration of nummulitic tests as current energy increases, and as wave influences become dominant, imbrication gives way to chaotic stacking of the tests. As this combination of shape and foraminiferal accumulation is influenced by light inten- sity and hydrodynamic forces, the relationship of all these factors could be used to estimate facies type and depth depth in the Paleogene warm seas (Hottinger, 1997). The Acervulinoidea (Fig. 6.55) live fixed to a substrate or most commonly to corals. The distribution of the coral- encrusting foraminiferan associations is also related to water depth gradients. Flat specimens mostly encrust the lower coral surfaces, where low light levels generally reduce competition for space with coralline algae. In contrast, globose morphotypes are successful on the coral upper surfaces, where lateral spatial The Cenozoic Larger Benthic Foraminifera: The Paleogene 491 competition with algae is higher (Bosellini and Papazzoni, 2003). In lamellar- perforate foraminifera, the thinning of the chamber walls results from thinning of each lamella, reflecting a decrease in the biomineralisation process with increased depth. Lamella thickness and flattening of shell shapes of larger foraminifera are products of water turbulence (Hallock 1979, 1999; Hallock and Glenn 1986; Hottinger 1977, 1997, 2000; Pêcheux 1995). It is also suggested by Ćosović et al. (2004) that the presence of fixed PLANORBULINOIDEA ACERVULINOIDEA Epoch "Letter - Stage Stages" L Lower Te Td E Tc L Tb M Ta3 E Ta2 L M Ta1 E Figure 6.55. Range chart of the main species of the Planorbulinoidea and the Acervulinoidea. 66.0 56.0 33.90 23.03 Age (Ma) Paleocene Eocene Oligocene Danian Selandian Thanetian Ypresian Lutetian Bartonian Priabonian Rupelian Chattian PLANKTONIC ZONES Cymbalopora Fabiania Planorbulina Planorbulinella Linderina Gunteria Korobkovella Eoannularia Caudriella Biarritzina Halkyardia Eorupertia Neocarpenteria Maslinella Wadella Carpenteria Victoriella Planolinderina Sphaerogypsina Sporadotrema Gypsina Wilfordia Discogypsina 492 Evolution and Geological Significance of Larger Benthic Foraminifera (attached) foraminifera, such as Biarritzina and Carpenteria, is a good argument for clear water environments in the Eocene carbonate ramp environments, while the pres- ence of orthophragminid dominated microfacies imply environments generally below storm wave base in middle to outer ramp areas. The spatial distribution of Paleogene larger benthic foraminifera within the coralgal framework buildups of the Cenozoic carbonates is illustrated in Fig. 6.54. The dep- osition occurred entirely within the photic zone, in symbiosis with algae. The region was one of patch reefs in which the areas attacked by oceanic swell may be called “forereef”. The energy distribution produced characteristic carbonate lithofacies and allowed distinctive assemblages of foraminifera to flourish: • The reef itself, being shallowest, produced characteristic bioherms and calcirudites of corals and algae, cemented by sparitic matrix as the lime muds were flushed away by wave action. • In the forereef shelves, proximal biostromes of corals and algae were cemented by sparite and micrite carbonates. Distally from the reefs, coral debris diminished and micrite increased, cementing the calcarenitic biogenic debris. Below the fair- weather wave- base, micrite would accumulate, which contained scattered larger foraminifera swept in from the reef shelves. The proximal forereef shelf contained faunas dominated by planorbulinids, cycloclypids, lepidocyclinids, operculinids and heterosteginids. • These shallow water carbonate facies were surrounded distally by deeper water sedi- ment containing abundant planktonic foraminifera, which may constitute up to 35% volume. The planktonic foraminifera had flourished in the water column offshore and reflect relatively deeper water, and low energy. Here coral debris diminished and micrite increased, cementing the calcarenitic biogenic debris. • Alveolinid, miliolid and miogypsinid larger benthic foraminifera are abundant in the backreef, sheltered from oceanic wave energies, or lagoon environment (see BouDagher- Fadel et al., 2000, p. 358; BouDagher-F adel and Price, 2013). Here bio- stromes of corals and algae were cemented by sparite and micrite carbonates. 6.5 Palaeogeographic Distribution of the Paleogene Larger Foraminifera The Paleocene (“ancient recent life”) epoch marks the beginning of the Paleogene period and the Cenozoic eon. During this period, the continents continued to drift towards their present positions, with the development of the Alpine- Himalyan orog- eny, the opening of the circum- Antractic seaway, but with South and North America still remaining separated (Fig. 6.56). In the wake of the end Cretaceous crisis, when about 83% of the Maastrichtian larger benthic foraminifera (see Chapter 5) became extinct, the Early Paleocene was a period of recovery. The Paleocene exhibited a cooler climate than the Late Cretaceous. In the Early Paleocene, the absence of other reef- building, high- temperature tolerating organisms in the low latitudes, enabled larger foraminifera to occupy this vacant niche and rapidly evolve (Scheibner et al. 2005). The Danian and Selandian mark the start The Cenozoic Larger Benthic Foraminifera: The Paleogene 493 Figure 6.56. Palaeogeographic and tectonic reconstruction of the Late Eocene (by R. Blakey http://j an.ucc. nau.edu/ ~rcb7/ paleogeographic.html). of the recovery period for the larger foraminifera. At this stage, most of the Tethyan foraminifera (75%) were small, primitive, European- form rotaliides with low diversity (see Hottinger, 2013), and it was not until the Late Paleocene (Thanetian) that many new larger foraminifera made their first appearance. This evolutionary trend allows a closely spaced zonation of Paleogene carbonates to be defined. Alveolinids and orthophragminids, followed by nummulitic foraminifera with thick marginal cords, developed during this warm period of the Late Paleocene. They invaded the Tethyan margins during the Eocene, and become large and abundant in the forereef environ- ments of the Early to Middle Eocene, only to disappear during the Middle Oligocene. Parallel to this evolution was that of the alveolinids, which after appearing in the Late Paleocene of Tethys, became abundant, colonizing reefal and backreef environments. The Paleocene-E ocene boundary saw the extinction of 25% of larger benthic foraminiferal species (Fig. 6.57), triggered by the thermal maximum (PETM) event, during which the sea surface temperature rose by 5°C in the tropics (Zachos et al., 494 Evolution and Geological Significance of Larger Benthic Foraminifera Paleogene Extinctions 35 30 25 20 15 10 5 Total Tethyan Cosmopolitan 0 American Figure 6.57. The number of larger foraminifera genera becoming extinctions at the top of each Paleogene stage boundary. 2003) and high CO2 levels led to oceanic acidification (Zachos et al., 2005). The ori- gins of the PETM are controversial. The Paleocene-E ocene boundary is marked by a period of intense flood basalt magmatism accompanying the opening of the North Atlantic (Eldhom and Thomas, 1993). The so-c alled North Atlantic Volcanic Provence (NAVP) (Fig. 6.58) is dated at ~58-5 5 Ma, and gave rise to volcanic fields that covered 1.3 million km2 (Courtillot and Renne, 2003). It has been suggested that this may have generated metamorphic methane from sill intrusion into basin- filling carbon- rich sed- imentary rocks (Storey et al., 2007). On the other hand, the reported presence of an iridium anomaly at the P-E boundary could be indicative of meteorite collision with Earth (Dolenec et al., 2000, Schmitz et al., 1996), which might in turn have triggered a change in the climate by releasing large amount of CO2 from oxidized underlying marine sediments (Higgins and Schrag, 2006). Cramer and Kent (2005) argue that the very rapid onset of the PETM is best explained by such an impact mechanism, and refer to the after effects of this proposed impact as the “Bolide Summer”. However, whatever its origin, during this stressful event 30 to 50% of the deep water small benthic foraminifera suddenly became extinct. The breakdown of the stable oligo- trophic environment of the larger benthic foraminifera resulted in the disappearance of most of the extreme K-s trategists (Hottinger, 1983). The replacement of SBZ4 by SBZ5 faunas (see Chart 6.2), as indicated by the gradual disappearance of Paleocene taxa such Ranikothalia and Miscellanea and the rise of Nummulites and Alveolina, suggests that Danian Selandian Thanetian Ypresian Lutetian Bartonian Priabonian Rupelian Chattian The Cenozoic Larger Benthic Foraminifera: The Paleogene 495 Figure 6.58. The Paleocene- Eocene world showing the position of the North Atlantic Volcanic Provence (NAVP) and the Ethiopian Traps. such an interruption may have taken place in platform environments at low-l atitude con- tinental margins (Scheibner et al., 2006; BouDagher- Fadel et al., 2015; Li et al. 2016). The Early Eocene witnessed the appearance of many new larger foraminifera genera (85% of which were cosmopolitan). A gradual increase of lineages followed this initial explosion of new forms, and during the Lutetian the American province (which up to the Middle Eocene had few endemic genera) experienced the highest appearance of new genera (Fig. 6.59). During the Lutetian the differences between the American province and the Western/E astern Tethys province were considerable, and larger foraminiferal bioprovinces became pronounced. At this stage, the main province was the Tethyan (containing 35% of all genera, see Fig. 6.60), but this itself exhibited sub- provinces with some genera being restricted to for example Western Tethys or the Indo- Pacific province. Traditionally, in the Paleogene the larger benthic foraminifera were considered to define three major, distinct palaeogeographic realms; namely, the American, the Western Tethys (which includes the modern day regions of West Africa, the Mediterranean and Tibet), and the Indo- Pacific provinces. However, a fourth distinct palaeogeographic province, the South West African realm has also been identified (see BouDagher- Fadel and Price, 2010a, 2013, 2014, 2017). Both the discocyclinids and lepidocyclinids origi- nated in the American province (see BouDagher-F adel and Price, 2010a, 2017), and although the discocyclinids rapidly attained a world- wide distribution, most of the earlier species became extinct at the end of the Middle Eocene in the American prov- ince, but continued elsewhere, only eventually to become extinct in the Late Eocene 496 Evolution and Geological Significance of Larger Benthic Foraminifera Paleogene Appearances 40 35 30 25 20 15 10 Total 5 Tethyan Cosmopolitan 0 American Figure 6.59. The number of new larger foraminifera genera appearing in each of the Paleogene stages. Paleogene Genera 90 80 70 60 50 40 30 20 Total 10 Tethyan Cosmopolitan 0 American Figure 6.60. The total number of larger foraminifera genera in each Paleogene stage. Danian Danian Selandian Selandian Thanetian Thanetian Ypresian Ypresian Lutetian Lutetian Bartonian Bartonian Priabonian Priabonian Rupelian Rupelian Chattian Chattian The Cenozoic Larger Benthic Foraminifera: The Paleogene 497 (BouDagher- Fadel and Price, 2017). The lepidocyclinids appeared in the Lutetian and, with the exception of some forms which reached the West African shelf (Brun et al., 1982), were mainly confined to the American province in the Middle to Late Eocene (BouDagher- Fadel and Price, 2010a). Lepidocyclinids migrated eastward to the Mediterranean in the early Rupelian and reached the Indo- Pacific towards the end of the Rupelian. The earliest L. (Nephrolepidina) species in Tethys occurred in the late Chattian, and they were almost certainly the direct descendants of L. (Lepidocyclina) and L. (Nephrolepidina) of the Eocene and Oligocene of West Africa and America (BouDagher- Fadel and Price, 2010a). Lepidocyclinid migration from the American to the Tethyan province ended after sea-l evel rose following the early Oligocene sea- level minimum noted by Berggren and Prothero (1992), Miller et al. (2005), Katz et al. (2008), and Miller et al. (2011) (Fig. 6.53). During the Eocene, the nummulitoid forms found in the Americas and South African provinces are very small (with diameters no more than 2mm), and different from those of Western Tethys and the Indo- Pacific (BouDagher- Fadel and Price, 2014). In the Western Tethyan province forms similar to the American Pararotalia, Chordoperculinoides and Operculina first appeared in West Africa in the Thanetian (P4b); later than their first appearance in the Americas. The Western Tethyan nummulitids have no apparent indig- enous Tethyan ancestors, but BouDagher- Fadel and Price (2014) demonstrated that they were derived from American ancestors presumably by trans- Atlantic migration, a process also inferred (BouDagher- Fadel and Price 2010a; 2010b; 2013; 2017) to have Figure  6.61. The inferred migration routes of orthophragminids during the Paleogene, shown by black arrows, from the Americas (1), to the Western Tethys (2), and on to the Indo- Pacific (3), and to South Africa (4) (from BouDagher- Fadel and Price, 2013). 498 Evolution and Geological Significance of Larger Benthic Foraminifera occurred at a later geological epoch to explain the global dispersal of two other LBF groups, the discocyclinids, the lepidocyclinids and the miogypsinids; see Fig. 6.61). During the Middle Eocene, the Tethyan (both Western Tethyan and the Indo- Pacific) province was dominated by large Nummulites and alveolinids. The abundance of Nummulites and Assilina, together with Discocyclina and Spiroclypeus distinguished this region (see BouDagher-F adel et al., 2015). The nummulitoids evolved first in the Americas and then migrated eastward to Western Tethys (and then eventually on to the Indo- Pacific) and to SW Africa (BouDagher- Fadel and Price, 2014). As species became geographically isolated, they evolved parallel but distinct lineages. Eocene to Oligocene nummulitoids of Southern Africa evolved directly from American ancestors and were distinct from the Tethyan and Indo-P acific forms, but a wave of nummulitoid migration occurred in the Miocene from the Mediterranean into the SW African prov- ince (see Chapter 7). The assilines are unknown in the Americas. The extinction of Nummulites can be correlated with the end of the Rupelian and planktonic zones P21. Miogypsinids lived from Oligocene to Middle Miocene times, and form a largely coherent morphological and phylogenetical unit (BouDagher-F adel and Price, 2013). They show a distinct provincialism: the evolutionary histories of American, European and Indo- Pacific miogypsinids, although showing roughly the same trends, are quite different (see BouDagher- Fadel and Price, 2013). They also originated in the Americas from Neorotalia in the Early Oligocene (Rupelian and P18; see BouDagher-F adel and Price 2010c, 2013). During the Early Oligocene a series of sea-l evel regressions (Katz et al., 2008; Miller et al., 2011) reduced the effective width of the early Atlantic Ocean sufficiently to facilitate trans-o ceanic migration of Neorotalia from the American province to the North African coast and on into the Mediterranean (BouDagher- Fadel and Price, 2013). During this time, Mediterranean shallow water niches were still occupied by the Paleogene Nummulites. However, towards the end of the Early oligocene (around 31– 29 Ma), environmental stresses, perhaps associated with cool- ing and the large flood basalt event in Ethiopia and Yemen (see Courtillot and Renne 2003), contributed to the disappearance of the last Mediterranean Nummulites. The Mediterranean Neorotalia (e.g. N. tethyana, see Boudagher- Fadel and Price, 2013) had by this stage become distinct from its American counterparts, and the disappearance of the Nummulites provided an opportunity for new phylogenetic lineages of miogyp- sinids of miogypsinids to fill the warm reefs of the Mediterranean. As the morpholo- gies of American and Mediterranean miogypsinids are seen to be crucially different, it follows that their evolutionary development was independent but closely parallel. After the last major regression in the early Chattian, the rising sea level and the con- tinuing oceanic rifting effectively isolated the Mediterranean–W est African shelf from the American province (around 28 Ma), ending any flow of Neorotalia or miogyp- sinids from America to the Mediterranean. It should be noted that during this time there was also a major change in oceanic circulation that resulted from the reversal of the direction of flow through the Panama Seaway (von der Heydt and Dijkstra, 2006). This reversal of flow is reported as being due to the tectonically driven widening of the Drake Passage and the narrowing of the seaway between the Mediterranean and Indo- Pacific, and that this may have also mitigated against further trans- Atlantic mio- gypsinid migration to the Mediterranean within the Late Oligocene (see BouDagher- Fadel and Price, 2013). The Cenozoic Larger Benthic Foraminifera: The Paleogene 499 The Middle to Late Eocene boundary (37.2 Ma) witnessed other major changes in global climate and ocean circulation (Berggren and Prothero, 1992), which might have been responsible for the major extinction event of foraminifera seen at the end of the Bartonian, notably the disappearance of the large nummulitid species. The global temperature fell by an average of 2- 40C (see Fig.  6.53). The Eocene- Oligocene boundary temperature fall followed the onset of the Late Eocene rapid sea level changes, which led to the disappearance of vast carbonate platform and lagoonal environments and the final extinction of the Eocene alveolinids. The move into an “ice house” climate may have been triggered at this time by the open- ing of the Tasmanian gateway (Smith and Pickering, 2003). What ever the cause, this boundary saw a large number of extinctions, with the large Nummulites and Assilina disappearing from Tethys, and in the American province 50% of the larger foraminifera abruptly becoming extinct (see Fig.  6.57). This was accompanied a sudden crash in the taxonomic richness of the tropical phytoplankton species (Macleod, 2015). The selectivity of Late Eocene extinctions, with the demise of mainly discoidal morphotypes of well-e stablished calcareous larger foraminifera (40% globally), seems to constrain the possible causes of this Late Eocene event (Banerjee and Boyajian, 1997). Palaeoclimatic evidence shows a trend of cooling during late Middle Eocene and at the Eocene- Oligocene (E-O ) boundary (Berggren and Prothero, 1992). Accelerated global cooling, with a sharp temperature drop of >2 °C occurred near the E- O boundary (Montanari et al., 2007). These global climate changes are attrib- uted to the expansion of the Antarctic ice cap following its gradual isolation from other continental masses. However, multiple bolide impact events, possibly related to a comet shower lasting 2.2 Ma, may have played an important role in causing the deterioration of the global climate at the end of the Eocene epoch (Montanari et al., 2007). Indeed, the Eocene saw an unusually large number of significant impact events (see Fig. 6.62), the cumulative effect of which would have added significantly to the environmental stress during this epoch. These happened at Chesapeake Bay (C) 35.5 ± 0.3 Ma (crater diameter = 90km), Popigai (P) 35.7 ± 0.2 Ma (diameter = 100km), Mistastin (Mi) 36.4 ± 4 Ma (diameter = 28 km), 28 Ma, Haughton (H) 39 ± 23 Ma (diameter = 23 km), Logancha (La) 40 ± 20 Ma (diameter = 20 km), Logoisk (Lo) 42.3 ± 1.1 Ma (diameter = 15 km), Kamensk (K) 49.0 ± 0.2 Ma (diameter = 25 km), and Montagnais (Mo) 50.50 ± 0.76 Ma (diameter = 45 km). These impacts happened in parallel with major changes in global climate, beginning in the Middle Eocene and culminating in the major earliest Oligocene Oi- 1 isotopic event (Montanari et al., 2007). The largest of these impacts (Chesapeake Bay and Popigai) at ~35Ma, when associ- ated with the later large flood basalt event in Ethiopia and Yemen (Fig 6.58) around 30 Ma (Courtillot and Renne, 2003), might have contributed to the disappearance (globally 41 %, in Tethys 17%) of many of the Eocene survivors. However, despite these impacts, and as seen in Fig. 6.57, the extinction event is not as large as that of the Bartonian-P riabonian boundary. Although some groups, such as the alveolinids, were more severely affected than others and disappeared completely at the E-O boundary, others survived this event only to become extinct within the Early Oligocene. 500 Evolution and Geological Significance of Larger Benthic Foraminifera Figure 6.62. The end-P aleogene world, showing the position of larger impacts occurring at that time. In the Early Oligocene (from about 33.5 Ma), the Drake Passage opened and there was further significant climatic cooling and ice volume increase (Berggren and Prothero, 1992). Most larger foraminifera which had survived the E- O boundary became adapted to cooler environments, while others migrated to the warmer Tethys, such as the American lepidocyclinids and miogypsinids. At the end of the Oligocene there were very few notable extinctions. This stratigraphic boundary is probably mainly connected to plate tectonic events, such the development of the Alpine-H imalayan orogeny, which do not seem to trigger major extinctions, or to gradual changes in climate caused by the growing thermal isolation of Antartica as Australia drifted northwards (Berggren and Prothero, 1992). 1 2 3 4 5 6 7 8 9 13 10 11 12 14 15 22 16 17 19 18 20 21 23 Plate 6.1 Scale bars: Figs 1, 4, 8, 11-1 2, 18, 20-2 3 = 0.5mm; Figs 2, 5- 7, 9 = 0.25mm; Figs 3, 10, 13-1 7, 19 = 1mm. Fig. 1. Cubanina alavensis Palmer, figured by Loeblich and Tappan (1988), Late Oligocene, Cuba. Fig.  2. Jarvisella karamatensis Brönnimann, holotype, figured by Loeblich and Tappan (1988), Miocene, Trinidad. Fig. 3. Liebusella soldanii (Jones and Parker), figured by Loeblich and Tappan (1988), Holocene, Caribbean. Fig.  4. Matanzia bermudezi Palmer, paratype, figured by Loeblich and Tappan (1988), Early Oligocene, Cuba. Figs 5-7 . Spiropsammia primula Seiglie and Baker, Congo Fan, West Africa, Kender coll., UCL. Fig.  8. Pavonitina styriaca Schubert, figured by Seiglie and Baker (1983), Cabinda. Fig.  9. Pavopsammina flabellum Seiglie and Baker, paratype, figured by Seiglie and Baker (1983), Cameroon. Fig. 10. Zotheculifida lirata (Cushman and Jarvis), figured by Loeblich and Tappan (1988), Late Oligocene, Trinidad. Fig. 11. Cyclammina sp., Eocene, Trinidad, UCL coll. Fig. 12. Haddonia torresiensis Chapman, lectotype, Holocene, western south Pacific, NHM 97.11.20.1. Figs 13-1 6. Saudia discoidea Henson, para- types, Middle Eocene (Lutetian), Ansab, Iraq, 13) NHM P36020; 14) NHM P36018; 15) NHM P36019; 16) NHM 36021. Fig. 17. Thomasella labyrinthica Grimsdale, figured by Sirel (1988), Early Eocene, Turkey. Fig. 18. Barattolites trentinarensis Vecchio and Hottinger, figured by Vecchio and Hottinger (2007), Ypresian to Lower Lutetian, Trentinara Formation, Italy. Figs 19-2 1. Coleiconus christianaensis Robinson, Middle Eocene, Upper Chapelton Formation, Jamaica, 19) holotype, NHM P52805; 20- 21) paratypes, 20) NHM P52808; 21) NHM P52809. Figs 22-2 3. Coskinolina sp., Late Eocene, India, UCL coll. Plate 6.2 Scale bars:  Figs 1, 2, 5- 6, 10-1 2, 14  =  0.5mm; Figs 3- 4, 7-9 , 13, 15, 16- 19  =  1mm. Fig.  1. Coskinolina sp., Ypresian, Laki Formation, Pakistan, UCL coll. Fig. 2. Coskinolina cf. douvillei (Davies), Middle Eocene, Upper Chapelton Formation, Jamaica, UCL coll. Fig. 3. Coskinon sp., Paleocene, Meting Limestone, Pakistan, UCL coll. Fig.  4. Anatoliella ozalpiensis Sirel, holotype, figured by Sirel (1988), Thanetian, Turkey. Fig. 5. Cushmania Americana (Cushman), Middle Eocene, Oman, NHM P35802. Fig. 6. Daviesiconus sp., Early Eocene, lower Laki Formation, Pakistan, UCL coll. Fig. 7 Dictyoconus sp., Early to Middle Eocene, Cuba, NHM P40043. Figs 8- 9. Dictyoconus indicus Davies, 8) Early Eocene, axial sec- tion, India, NHM P28105; 18) Lutetian, Sulman, S.W. Iraq, NHM 35828. Fig. 10. Coskinolina sp., White Limestone, Manchester, Jamaica, UCL coll. Figs 11- 12. Verseyella jamaicensis (Cole), Early Eocene, lower Chapelton Formation, Jamaica, NHM P52823- 24. Figs 13. Coskinolina balsilliei Davies, Lutetian, Sulman, S.W. Iraq, NHM P35781. Fig.  14. Fallotella sp., Middle Paleocene, Laki-K hirthar, Pakistan, UCL coll. Fig. 15, 19. Alveolina elliptica var. flosculina Silvestri, Middle Eocene, Qatar, 15) NHM P40266; 19) NHM P40256. Fig.  16. Alveolina aramaea Hottinger, Early Eocene, Dunghan Hill, Pakistan, NHM P52544. Fig. 17. Alveolina katicae White, Eocene, Oman, NHM coll. White and Racey, Wr77. Fig. 18. Glomalveolina delicatissima (Smout), holotype, Middle Eocene, Qatar, NHM P40265. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 1 3 2 4 A B 7 5 6 7 8 9 Plate 6.3 Scale bars:  Figs 1- 4, 6  =  1mm; Figs 5, 7- 9  =  0.5mm. Figs 1-2 . Alveolina elliptica (Sowerby) and Alveolina elliptica nuttalli Davies, Eocene, Afghanistan, NHM Skinner coll. P7404. Fig. 3. Alveolina minervensis Hottinger, Late Paleocene to Early Eocene, Montagne Noire, Aquitaine, South France, NHM Hottinger coll. Fig. 4. Alveolina globosa (Leymerie), Ypresian, Eocene, Laki Limestone (Laki Formation), Sakesar Peak in the Salt Range, Pakistan, NHM Davies coll. Fig. 5. A) Coskinolina sp., B) Glomalveolina sp., Ypresian, Lower Laki Formation, Pakistan, UCL coll. Fig. 6. Alveolina vredenburgi Davies 1937 (= Alveolina cucumiformis Hottinger), topotype, axial section, Late Paleocene, Aquitaine, France,UCL coll. Figs 7-9 . Austrotrillina asmariensis Adams, Oligocene, Kirkuk Well K.93, Iraq, NHM P47581- 3. Plate 6.4 Scale bars:  1-1 1 = 1mm. Fig.  1. Alveolina vredenburgi Davies 1937 (= Alveolina cucumiformis Hottinger), topotype, equatorial section, Late Paleocene, Aquitaine, France, UCL coll. Figs 2, 4. Alveolina leupoldi Hottinger, Early Eocene, Coustouge, France, 2) solid specimens embedded in rock, 4) thin section from the same rock, UCL coll. Figs 3, 5-7 . Alveolina globosa (Leymerie), Early Eocene, Coustouge, France, 3) thin section, 5) solid specimen embedded in the same rock, 6) Early Eocene, Khirthar, Pakistan; 7) Early Eocene, Meting Limestone, Laki group, Pakistan, UCL coll. Figs 8, Alveolina oblonga d’Orbigny, Dunghan, Siah Koh, NHM Davies coll. Fig. 9. Alveolina subpyrenaica Leymerie, Early Eocene, Zagros Limestone, Iran, NHM coll. Fig. 10. Alveolina subpyrenaica Leymerie, Lutetian, India, UCL coll. Fig. 11. Alveolina elliptica nuttalli Davies, Early Eocene, Meting Limestone, Laki group, Pakistan, UCL coll. 1 2 3 5 4 7 6 8 9 10 11 1 2 3 4 A A B C B 6 5 7 8 9 10 11 12 13 14 Plate 6.5 Scale bars: Figs 1-4 , 6, 12-1 4 = 1mm; Figs 7, 9-1 1, = 0.5mm; Figs 5, 8 = 0.25mm. Fig. 1. Alveolina subpyrenaica Leymerie, Early Eocene, Zagros Limestone, Iran, NHM coll. Fig. 2. Alveolina elliptica nuttalli Davies, Lutetian, India, UCL coll. Fig. 3. Thin section photomicrograph of Alveolina elliptica (Sowerby) and Alveolina cf. stipes Hottinger, Late Eocene, Khirthar group, Pakistan, UCL coll. Fig. 4. Alveolina leupoldi, Hottinger, Early Eocene, Coustouge, France, UCL coll. Fig. 5 A) Borelis pygmaeus Hanzawa, B) rodophytes fragments, C) Heterostegina (Vlerkina) borneensis van der Vlerk, Oligocene, Borneo, 65/9 Loc.205, UCL coll. Fig. 6. Borelis sp., Late Oligocene, Ras Chekka, Lebanon, UCL coll. Fig. 7. Pellatispira sp., Borelis sp., Late Eocene, Syria, UCL coll. Fig. 8. Borelis haueri (d’Orbigny), Oligocene, Iran, UCL Banner coll. Fig. 9. Bullalveolina bulloides Reichel, figured by Hottinger (2006), Early Oligocene, Spain. Fig. 10. Globoreticulina iranica Rahaghi, figured by Hottinger (2006), Middle Eocene, Shiraz, Iran. Fig. 11. Nummulites fossulata de Cizancourt, Middle Eocene, figured by BouDagher-F adel et al. (2015), Shenkeza section, Tibet. Fig. 12. Assilina leymeriei Archiac and Haime, Nummulites atacicus Leymerie, Discocyclina sp., Alveolina sp., Late Eocene, Upper Khirthar, Pakistan, UCL coll. Fig. 13. Assilina sublaminosa Gill, Middle Eocene, figured by BouDagher- Fadel et al. (2015), Shenkeza section, Tibet. Figs 14. Alveolina palermitana Hottinger, Middle Eocene, Middle Khirthar, Pakistan, UCL coll. Plate. 6.6 Scale bars: Figs 1-2 , 6- 8, 11 = 1mm; Figs 3-5 , 9-1 0 = 0.5mm. Fig. 1. Thin section photomicrograph of Alveolina elliptica nutalli Davies, central, flosculinized parts; outer whorls missing, Orbitolites complanatus Lamarck, small miliolids, Operculina sp., Early- Middle Eocene, Laki- Khirthar, UCL coll. Fig. 2. Alveolina moussoulensis Hottinger, Early Eocene, figured by BouDagher- Fadel et al. (2015), Qumiba section, Tibet. Figs 3- 5. Rhabdorites urensis (Henson), paratypes, Lutetian, Iraq, NHM P35986. Fig. 6. Linderina burgesi (Schlumberger), Kohat formation, Pakistan UCL coll. Fig. 7. Rhabdorites sp. registered as Neorhipidionina macfadyeni Henson, syntypes, late Lutetian, Iraq, NHM P36000. Fig. 8. Neorhipidionina macfadyeni Henson, syntypes, late Lutetian, Iraq, NHM P36000. Fig. 9. Neorhipidionina williamsoni (Henson), Lutetian, Iraq, NHM P36012. Fig. 10. Daviesina langhami Smout, Paleocene, figured by BouDagher-F adel et al. (2015), Zongpubei section, Tibet. Fig. 11. Yaberinella jamaicensis Vaughan, Eocene, Yellow Limestone, Jamaica, Davies coll., sample J 505 M. 3 1 2 4 6 7 10 5 8 9 11 2 1 4 3 7 5 6 8 9 10 11 12 Plate 6.7 Scale bars: Figs 1- 12 = 1mm. Figs 1- 3. Opertorbitolites sp., Early Eocene, Alveolina Corbarica Zone, Eastern Aquitaine, France, NHM Hottinger coll. Figs 4. Opertorbitolites lehmanni (Montanari), reg- istered as “Opertorbitolites sp.”, Eocene, Oman, White-R acey coll., NHM P52865. Fig. 5. Opertorbitolites cf. douvillei Nuttall, Eocene, Oman, White coll., NHM P52866. Figs 6-7 . Orbitolites complanatus Lamarck, Middle Eocene, 6) Bois- Gouët, France, NHM coll.; 7) figured by BouDagher- Fadel et al. (2015), Zongpubei section,Tibet. Fig. 8. Opertorbitolites douvillei (Nuttall), Early Eocene, Alveolina Corbarica Zone, eastern Aquitaine, France, NHM coll. Fig. 9. Thin section with Orbitolites biplanus Lehmann and Alveolina leupoldi, Hottinger, Ilerdian (Lower Eocene), Coustouge (Corbières), France, UCL coll. Figs 10- 12 Pseudophragmina floridana (Cushman), Eocene, Georgia, UCL coll. Plate 6.8 Scale bars: Figs 1-3  = 1mm; Figs 4- 7 = 0.25mm. Figs 1- 2. Orbitolites complanatus Lamarck, Middle Eocene, 1)  Libya; 2)  Coustouge, France UCL coll. Fig.  3. Thin section photomicrograph of A) Alveolina sp., B) Orbitolites omplanatus Lamarck, Middle Eocene, 1) Libya; 2) Coustouge, France UCL coll. Figs 4- 6. Praerhapydionina delicata Henson, Oligocene, Buff calcarenites, Jamaica, NHM P52829- 21. Fig. 7. Rhabdorites malatyaensis (Sirel), figured by Hottinger (2007), Middle Eocene, Turkey. 1 4 5 2 A B 6 C 3 7 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 Plate 6.9 Scale bars:  Figs 1- 4, 12- 17  =  0.5mm; Figs 5- 11  =  1mm. Fig.  1. Glomalveolina dachelensis Schwager, figured by Loeblich and Tappan (1988), Paleocene, Egypt. Fig.2 Malatyna drobneae Sirel and Acar, type figured by Sirel and Acar (1993), Lutetian, Malaya. Fig. 3. Praebullalveolina afyonica Sirel and Acar, holotype, figured by Sirel and Acar (1982), Eocene, Turkey. Fig.  4. Fabularia hanzawai Robinson, cotype, Eocene, Saint Andrew Claremont Formation, Jamaica, NHM P52840. Figs 5- 7. Fabularia discolithus Defrance, 5) Eocene, calcaire grosier de Rennes, France, Brady coll., NHM P41603; 6-7 ) Lutetian middle- calcaire grossier, Chaumont-e n- Vexin, Paris Basin, NHM P33071. Fig.  8. Aberisphaera gambanica Wan, figured by BouDagher-F adel et al. (2015) from the Paleocene, Shenkeza section, Tibet. Fig. 9. Lacazina sp., Eocene, Spain, UCL coll. Figs 10-1 1. Lacazinella wichmanni (Schlumberger), Late Eocene, Indonesia, UCL coll. Fig. 12. Pseudofabularia matleyi (Vaughan), figured by Loeblich and Tappan (1988), Middle Eocene, Chapelton Formation, Jamaica. Figs 13- 14. Wilfordia sarawakensis Adams, paratype, Eocene, Sarawak, Malaysia, NHM P46464. Figs 15- 17. Austrotrillina paucialveolata Grimsdale, syntypes, Oligocene, Kirkuk Well 14, Iraq, NHM P40689. Plate 6.10 Scale bars: Figs 1- 4, 6, 8, 10- 11, 16 = 1mm; Figs 5, 7, 9, 12- 15 = 0.5mm. Figs 1- 2. Austrotrillina paucialveolata Grimsdale, syntypes, Oligocene, Kirkuk Well 14, Iraq, NHM P40681. Fig. 3. Karsella hot- tingeri Sirel, holotype, figured by Sirel (1997), Thanetian, Turkey. Fig. 4. Archiacina armorica (d’Rrchiac), figured by Loeblich and Tappan (1988), Oligocene, France. Fig.  5. Dendritina cf. rangi d’Orbigny, Early Miocene (Aquitanian), Kirkuk Well 22, Iraq, NHM P39432. Fig. 6. Haymanella paleocenica Sirel, figured by Sirel (1999), Thanetian, Turkey. Fig. 7. Elazigella altineri Sirel, type figured by Sirel (1999), Thanetian, Turkey. Fig. 8. Archaias sp. (pillars are present, but registered as Heterillina hensoni Grimsdale), syntype, Oligocene, Kirkuk Well 14, Iraq, NHM P40679. Fig. 9. Heterillina hensoni Grimsdale, syntype, Oligocene, Kirkuk Well 14, Iraq, NHM P40679. Fig. 10. Hottingerina lukasi Drobne, figured by Drobne (1975), Middle Paleocene, Yugoslavia. Fig. 11. Archaias aduncus (Fichtel and Moll), Oligocene, Shiranish Islam, Iraq, NHM P39651. Fig. 12. Archaias sp. (pillars are present, but registered as Peneroplis sp.), Oligocene, Buff calcaren- ites, Jamaica, NHM P52828. Fig. 13. Cyclorbiculinoides jamaicensis Robinson, holotype, figured by Robinson (1974), Eocene, Jamaica. Fig. 14. Neorhipidionina spiralis Hottinger, figured by Hottinger (2007), Middle Eocene, Iran. Fig. 15. Neotaberina neaniconica Hottinger, figured by Hottinger (2007), Middle Eocene, Iran. Fig. 16. Yaberinella hottingeri Robinson, Eocene, Punta Gorda-1 6270’- 80’, Nicaragua, NHM P52822. 1 2 3 4 5 6 7 8 9 11 12 10 13 14 15 16 2 3 4 5 6 1 7 8 9 10 11 12 Plate 6.11 Scale bars: Figs 1- 6, 9-1 2 = 1mm; Figs 7- 8 = 0.5mm. Figs 1-2 . Somalina hottingeri White, Eocene, Seeb Limestone Formation, Wadi Fatah, Oman, NHM P52861- 2. Figs 3- 6. Twaraina seigliei Robinson, Middle Eocene, Twara- 1, Nicaragua Rise, NHM P52814-1 8. Fig. 7. Meghalayana indica Matsumaru and Sarma, holotype, Late Eocene, Meghalaya State, NE India, Matsumaru’s coll., Saitama Univ, 8865. Fig. 8. Protogypsina indica Matsumaru and Sarma, holotype, Paleocene, Jaintia Hills, Meghalaya State, NE India, Matsumaru’s coll., Saitama Univ, 8866. Figs. 9. Yaberinella jamaicensis Vaughan, Eocene, Yellow Limestone, Jamaica, NHM Davies coll., sample J.505 M. Figs 10–​11. Yaberinella jamaicensis Vaughan registered in the NHM as from Lutetian, Oman, 10) axial section, NHM P52270; 11) A- form bilocular protoconch, NHM. P52266. These are undoubtedly Yaberinella, however, this genus is Caribbean. Their recovery in Oman is man- made:  either by a confusion of labels or by the occurrence of ballast stones transported in slavery ships on their return from Jamaica. Hottinger personal communication:  “I have been at the point indi- cated near Muscat: No trace nor even a possibility of their occurrence there considering the local geology.” Fig. 12. Raoia indica Matsumaru and Sarma, holotype, Paleocene, Jaintia Hills, Meghalaya State, NE India, Matsumaru’s coll., Saitama Univ, 8867. Plate 6.12 Scale bars:  Figs 1-7   =  1mm. Fig.  1. Nummulites intermedius (d’Archiac), Operculina sp., Discocyclina sp., Early Eocene, Lower Nari, Pakistan, UCL coll. Fig. 2. Nummulites atacicus Leymerie, Early Eocene, Meting- Laki, Pakistan UCL coll. Fig.  3. Alveolina sp., Orbitolites complanatus Lamarck, Early Eocene, Meting- Laki, Pakistan, UCL coll. Fig. 4. Archaias kirkukensis Henson, paratype, Late Oligocene to Early Miocene (Chattian to Aquitanian), Kirkuk Well no. 57, Iraq, NHM P39645. Fig. 5. Distichoplax biserialis (Dietrich), Eocene, Java, UCL coll. Fig. 6. Codiacean algae, Sorites sp., Oligocene, France, UCL coll. Fig. 7. Thin section photomicrograph of Wilfordia sarawakensis Adams, Eocene, Sarawak, Malaysia, same section as of the paratypes (Plate 6.9, figs 12- 14). 1 2 3 4 5 6 7 4 1 2 3 1 6 7 8 11 5 13 9 12 10 14 15 Plate 6.13 Scale bars:  Figs 1- 15  =  1mm. Figs 1- 2. Biplanispira mirabilis (Umbgrove), Late Eocene, Indonesia,UCL coll. Fig. 3. Ranikothalia nuttalli kohatica (Davies), latest Paleocene- earliest Eocene, lower member of the Jafnayn Formation, Al Khawd, Oman, BMNHP 52442. Figs 4- 7. Miscellanea miscella (d’Archiac and Haime), S.W. of Hilaia, South India, UCL coll. Figs 8- 10. Miscellanea meandrina (Carter), Paleocene, Qatar; 7) A- form, axial section, NHM P40232; 8) axial section, NHM P40234; 9) equatorial sec- tion, NHM P40230; 10) paratype, B- form, equatorial section, NHM P40239. Fig. 12. Miscellanea miscella var. dukhani Smout, paratype, Paleocene, Qatar, NHM P40251. Figs 13- 14. Miscellanea stampi (Davies), Paleocene, Upper Ranikot Beds, the Samana Range Pakistan, NHM P41615. Fig. 15. Miscellanea miscella (d’Archiac and Haime), Paleocene, Indonesia, UCL coll. 1 3 4 2 5 6 7 10 8 11 9 15 12 13 14 16 Plate 6.14 Scale bars: Fig. 1, 7, 11, 13, 15 - 16 = 1mm; Fig. 2- 6, 8-1 0, 12, 14 = 0.5mm. Figs 1- 5. Pellatispira sp., Late Eocene, 1) Pakistan; 2- 5) Indonesia, UCL coll. Fig. 6. Pellatispira fulgeria Whipple, Late Eocene, Sumatra, UCL coll., AS33. Fig. 7. Vacuolispira inflata (Umbgrove), Late Eocene, Borneo, UCL coll. Figs 8- 10. Assilina cuvillieri Schaub, B forms, Wadi Bani Khaled WK21, 8,10) equatorial sections, NHM P52513, NHM P52516, 9) axial section, NHM P52516. Figs 11, 16. Heterostegina (Vlerkina) borneensis van der Vlerk, Late Oligocene, Borneo, UCL coll. Fig. 12. Grzybowskia multifida Bieda, Eocene, Carpathians, NHM coll. Fig. 13. Heterostegina (Heterostegina) sp., Eocene, Brazil, UCL coll. Figs 14- 15. Heterostegina (Vlerkina) borneensis van der Vlerk, paratypes, Te- Miocene, Borneo, NHM P45037- 8. 1 2 8 3 14 5 7 9 4 10 6 13 11 12 Plate 6.15 Scale bars:  Figs 1- 14  =  1mm. Fig.  1. Planostegina sp., Operculina sp., Oligocene, Borneo, UCL coll. Fig.  2. Heterostegina (Vlerkina) borneensis van der Vlerk, Late Oligocene, Borneo, UCL coll. Fig. 3. Tansinhokella tatauensis Banner and Hodgkinson, paratype, Eocene Limestone, Tatau Formation, Sarawak, Borneo, NHM P52296. Fig. 4. Tansinhokella sp., Cycloclypeus sp., Late Oligocene, Borneo, UCL coll. Fig. 5. Tansinhokella yabei (van der Vlerk), latest Oligocene, Soembal, Borneo, NHM P45042. Fig. 6. Spiroclypeus leupoldi van der Vlerk, Lepidocyclina sp., fragments of rodophyte algae, Oligocene, Java, UCL coll. Figs 7. Spiroclypeus umbonata Yabe and Hanzawa, Oligocene, Java, UCL coll. Figs 8-9 . Heterostegina (Hetrostegina) sp., Oligocene, Indonesia, UCL coll. Fig.  10. Silvestriella tetraedra Gümbel, Eocene, Gassino, Torino, Italy, NHM P44956. Figs 11- 13. Spiroclypeus vermicularis Tan Sin Hok, Late Eocene, East Borneo, UCL coll. Fig. 14. Nummulitoides margaretae Haynes and Nwabufo-E ne, paratype, Late Paleocene (Thanetian), El Fogaha, Libya, NHM P52253. 1 2 3 4 5 6 7 8 9 10 11 12 13 Plate 6.16 Scale bars: Figs 1, 3- 4, 7- 13 = 1mm; Figs 2, 5- 6 = 0.5mm. Fig. 1. Lepidocyclina (Lepidocyclina) ocalana Cushman, Middle Eocene, Gilchrist Co., Florida, NHM P51961. Fig.  2. Pseudolepidina trimera Barker and Grimsdale, cotype, base of Middle Eocene, Vera Cruz, Mexico, NHM P51986. Figs 3, 12. Discocyclina sp., Eocene, Fontcouverte, France, UCL coll. Fig.  4. L. (Nephrolepidina) veracruziana (Vaughan and Cole)  =  Triplalepidina veracruziana Vaughan and Cole, topotype, Late Eocene, Arroyo Terrero, near Palma Sola Mexico, NHM P37905. Figs 5- 6. Pellatispirella antillea Hanzawa, Middle Eocene, Soldado Rock, Trinidad NHM P33349. Figs 7- 8. Eulepidina papuaensis (Chapman), Late Oligocene, Borneo, UCL coll. Fig. 9. Nemkovella mcmilliana BouDagher- Fadel and Price, holotype, Early Ypresian, figured by BouDagher- Fadel and Price (2017), South Africa, UCL MF491. Fig. 10. Biplanispira sp., Eocene, Tatau Formation, Sarawak, Borneo, UCL coll. Fig.  11. Nummulites somaliensis Nuttall and Brighton, Middle Eocene, Qatar, NHM P40244. Fig.  13. Tansinhokella tatauensis Banner and Hodgkinson, paratype, Eocene,Tatau Formation, Sarawak, Borneo, NHM P49525. Plate 6.17 Scale bars: Figs 1- 4, 6- 10 = 1mm; Fig. 5 = 0.5mm. Fig. 1. Cycloclypeus sp., Late Oligocene, with a planktonic foraminifera test embedded in the broken part of the test, Java, UCL coll. Figs 2-4 . Assilina granulosa var. chumbiensis Gill, Early Eocene, Lower Bhadrar Beds, Pakistan, NHM P41522; 3) enlargement of Fig. 5. Assilina subdaviesi Gill, Early Eocene, Lower Bhadrar Beds, Pakistan, NHM P41543. Figs 6-7 . Assilina sp., Late Paleocene, Pakistan, Ranikot group, solid specimens, UCL coll. Fig. 8. Assilina sp., Eocene, axial sections, Coustouge, France, UCL coll. Fig. 9. Orbitosiphon praepunjabensis Adams, holotype, Late Paleocene, Khairabad Limestone, Dhak Pass, Salt Range, Pakistan, NHM P51968. Fig. 10. Actinosiphon semmesi Vaughan, Chicontepec Formation, El Cristo Well, Veracruz, Mexico, UCL coll. 1 2 3 4 5 6 4 9 7 8 10 2 1 3 4 5 6 7 8 11 10 9 12 Plate 6.18 Scale bars:  Figs 1-1 2  =  1mm. Fig.  1. Assilina mamillata (d’Archiac), Early Eocene, Yozgat, Turkey, UCL coll. Figs 2- 4. Assilina granulata (d’ Archiac), Middle Eocene, 2- 3) Barail Formation, Bangladesh; 4) solid specimen, UCL coll. Figs 5- 6. Operculina sp., Oligocene, Kalimantan, UCL coll. Fig. 7. Thin section photomicrographs of Planostegina sp., fragments of Eulepidina sp., Amphistegina sp., Loc. 130, Borneo, UCL coll. Fig. 8. Nummulites masiraensis Carter, Eocene, India, UCL coll. Fig. 9. Operculina douvillei Dorıcieux, Early Lutetian, NHM Davies coll. Fig. 10. Nummulites fichteli Michelotti, Oligocene, Biarritz, France, UCL coll. Fig. 11. Operculina aegyptiaca Hamam, paratype, megalospheric form, latest Early Eocene, Gebel Gurnah, Luxor, Egypt, NHM P49827. Fig.  12. Operculina subgranosa Grimsdale, Eocene, France, NHM coll. Plate 6.19 Scale bars: Figs 1- 9 =1mm. Figs 1- 3. Nummulites fichteli- intermedius (d’Archiac) (= Nummulites clypeus), Lower Nari Formation, Oligocene, Pakistan, UCL coll. Fig.  4. Nummulites fichteli Michelotti, Oligocene, Nummulitic rock, cliffs at Biaritz south of France, NHM P49522. Figs 5-8 . Nummulites gize- hensis (Forskal), Late Lutetian, 5- 7) Spain; 8) France, UCL coll. Fig. 9. Assilina sp., Eocene, France, solid specimen, UCL coll. 2 1 4 3 5 6 7 8 9 1 2 4 3 4 5 5 6 7 8 11 12 13 9 10 15 14 16 Plate 6.20 Scale bars: Figs 1- 16 =1mm. Figs 1- 2. Nummulites mamilla Fichtel and Moll, Early Eocene, Laki, Pakistan, 1) megalospheric form; 2) microspheric form, UCL coll. Fig. 3. Nummulites perforatus (de Montfort), microspheric form, Lutetian, San Giovanni Ilarione, Italy, NHM Davies coll. Fig. 4. Nummulites intermedius (d’Archiac), Eocene, India, NHM P30148. Fig. 5. Nummulites beneharnensis de la Harpe, B- form, middle Lutetian, Wadi Rusayl, Oman, NHM P52275. Fig. 6. Assilina leymeriei (d’Archiac and Haime), Early Eocene, Lower Bhadrar Beds, Pakistan, NHM P41509. Figs 7- 8. Palaeonummulites kugleri (Vaughan and Cole), Oligocene, Falling Waters State Park, Chipley, Florida, Suwannee Limestone, USA, UCL MF3237) equatorial section; 8) axial section. Figs 9- 11. Assilina daviesi de Cizancourt, Early Eocene, Lower Bhadrar Beds (Salt Range), Pakistan, 9) NHM P41524; 10-1 1) NHM 41527- 8. Figs 12- 13. Operculinoides ocalanus (Cushman), Oligocene, Brazil, UCL coll. Fig. 14. Assilina sp. and Planocamerinoides sp., Middle Eocene, Kopili Formation, UCL coll. Figs 15- 16. Chordoperculinoides sahnii (Davies), India, registered as Ranikothalia sahnii Davies, Paleocene, French West Africa, NHM P40350, 15) equatorial section showing an initial nummulitic (involute) spiral and operculine final stage; 16) axial section. Plate 6.21 Scale bars: Figs 1-1 6 = 1mm. Fig. 1. Nummulites lamarcki d’Archiac and Haime, megalospheric form of Nummulites laevigatus (Bruguière), Middle Eocene, England, UCL coll. Fig.  2. Nummulites sp., Eocene, S.E. Coast of Arabia, UCL coll. Fig. 3. Nummulites gizehensis (Forskal), late Lutetian, Libya, UCL coll. Fig. 4. Nummulitc Limestone, Eocene, Libya, UCL coll. Figs 5-6 . Nummulites sp., Eocene, Gerona, Spain; 6)  enlargement of Fig. 5, UCL coll. Fig. 7. Nummulites globulus Leymerie, Eocene, Qatar, NHM 40258. Figs 8-9 . Nummulites vascus Joly and Leymerie, Oligocene, Iran, UCL coll. Fig.  10. Nummulites aturicus Joly and Leymerie, Middle Eocene, UCL coll. Figs 11-1 3. Nummulites sp., Middle Eocene, France, 11) SEM photograph of a solid specimen showing the proloculus; 12) thin section; 13) enlargement of 12 to show the marginal cord, UCL coll. Figs 14- 15. Nummulites fichteli Michelotti, Oligocene, from Nummulitic rock cliffs at Biaritz, France, NHM P49521. Fig. 16. Nummulites cf. striatus (Bruguière), late Middle Eocene, Priabonian, India, UCL coll. 1 2 3 4 5 6 77 8 9 10 10 11 12 13 14 15 16 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 Plate 6.22 Scale bars: Figs 1- 6, 9- 10, 13- 15 = 1mm; Figs 7- 8, 11- 12 = 0.5mm. Figs 1- 2. Ranikothalia nut- talli (Davies), topotypes, Late Paleocene, Upper Ranikot Beds, Thal, Pakistan, NHM P41614. Figs 3- 5. Ranikothalia sindensis (Davies), Late Paleocene, Upper Ranikot, Beds, Pakistan, 3) a fragment showing the marginal cord, 4) a thin section with an assemblage of Ranikothalia, Assilina, fragments of Discocyclina and rodophyte algae; 5) Late Paleocene, Punjab Salt range upper part Khairabad Limestone, NHM coll. Fig. 6. Thin section photomicrograph of Actinocyclina radians (d’Archiac), Discocyclina sp., Middle Eocene, Kopili Formation, Bangladesh, UCL coll. Fig. 7. Asterophragmina pagoda (Rao), figured by Loeblich and Tappan (1988), Late Eocene, Burma. Fig. 8. Discocyclina ranikotensis Davies, Late Paleocene, Upper Ranikot Bed, Pakistan, UCL coll. Figs 9- 10. Discocyclina dispansa Sowerby, Eocene, Goojerat, Western India, NHM P539, 9) axial section; 10) equatorial section. Figs 11-1 2. Discocyclina sheppardi Barker, Paleocene, Soldado Rock, Trinidad, NHM P33350-1 , 11)  axial section. Figs 13-1 4. Discocyclina peruviana (Cushman), Terebratula Bed, Peru, 13) axial section; 14) equatorial section, UCL coll. Fig. 15. Thin section photomicrograph of Nummulites perforatus (de Montfort), Discocyclina dispansa Sowerby, Globigerina sp., UCL coll. Plate 6.23 Scale bars:  Figs 1- 2, 4- 11= 1mm; Figs 3, 12- 13  =  0.5mm. Figs 1- 2. Thin section photomi- crograph of Discocyclina dispansa Sowerby, Biplanispira sp., Asterocyclina sp., Middle Eocene, Kopili Formation, Bangladesh, UCL coll. Fig.  3. Polylepidina sp., latest Middle Eocene, Jamaica, UCL coll. Figs 4- 11. Discocyclina sp., microspheric equatorial section, Eocene, France,, UCL Grimsdale Coll. GS/ 50, 4) enlargement of fig. 5; 6- 10) solid specimens;7) enlargement of embryonic chambers; enlargement of equatorial chambers. Fig. 12. Operculina sp., Eocene, Philippines, UCL coll. Fig. 13. Orbitolites sp., Eocene, Libya, UCL coll. 2 3 1 4 5 6 7 8 9 10 11 12 13 1 3 2 4 5 6 7 8 9 10 11 Plate 6.24 Scale bars:  Figs 1, 3, 5, 8 =0.5mm; Figs 2, 4, 6- 7, 9- 11  =  1mm. Fig.  1. Discocyclina sella (d’Archiac), Early Eocene, Laki, India, UCL coll. Figs 2-5 . Hexagonocyclina cristensis (Vaughan), 2-3 ) topotype, Early Eocene, Mexico, NHM P32633; 3) enlargement of the embryonic apparatus of fig. 2; 4-5 ) Early Eocene, well at El Cristo, Mexico, UCL coll. 5) enlargement of the embryonic apparatus of fig. 4. Fig. 6. Discocyclina californica (Schenk), topotype, Vaquelos Formation, California, UCL coll. Figs 7- 8) Discocyclina sp., equatorial and axial sections, Eocene, France, UCL coll; 8) enlargement of the embryonic apparatus of fig. 7. Figs 9-1 0) Asterocyclina stella (Gümbel), Late Eocene, 9) Karia, Turkey, NHM P37910; 10) Eocene, France, UCL coll. Fig. 11. Thin section photomicrograph of Discocyclina sp., Asterocyclina stel- lata (d’Archiac), Middle Eocene, Kopili formation, UCL coll. 1 2 3 4 5 6 7 8 9 10 11 12 Plate 6.25 Scale bars: Figs 1-1 2 = 1mm. Figs. 1- 3. Lepidocyclina (Lepidocyclina) canellei Lemoine and Douvillé, Oligocene, Brazil, UCL coll. Figs. 4- 6. Lepidocyclina (Lepidocyclina) pustulosa (Douvillé), (4–​5) Oligocene, Brazil; 6. Pliolepidina tobleri Douvillé, synonymous with L. (L.) pustulosa (Douvillé), Eocene, Masparrito Member, Venezuela, UCL coll. Fig. 7. Discocyclina sp., Eocene, Venezuela, UCL coll. Fig. 8. Thin section photomicrograph of Eulepidina sp., Eulepidina dilatate (Michelotti), Oligocene, Java, UCL coll. Fig. 9. L. (Nephrolepidina) chaperi Lemoine, microspheric form, Late Oligocene, Brazil, UCL coll. Fig. 10. Eulepidina ephippioides (Jones and Chapman), Oligocene, Borneo, UCL coll. Figs. 11–​12. Laffitteina vanbel- leni Grimsdale, Eocene, Kourdane, Syria, NHM P40676- 7. 1 3 4 2 5 6 7 8 9 B A C 10 11 12 Plate 6.26 Scale bars: Fig. 1, 5, 9- 12 = 1mm; Figs 2- 4, 6- 8= 0.5mm Fig. 1. Eulepidina sp., Victoriella sp., Oligocene, Borneo, UCL coll. Figs 2-4 . Eorupertia sp., Late Eocene, 2) Brazil, UCL coll. 3-4 ) France, UCL coll. Fig. 5. Biarritzina sp., Late Eocene, Brazil, UCL coll. Fig. 6- 7. Discogypsina discus (Goës), 6) Eocene, France, UCL coll; 7) Batu, North Borneo, NHM NB 9030. Fig. 8. Sphaerogypsina sp., Oligocene, Kalimantan, UCL coll., 62-4 66B. Fig. 9. Pseudophragmina floridana (Cushman), Eocene, Georgia, UCL coll. Fig. 10. Thin section photomicro- graph of A) Gypsina sp., B) fragment of Eulepidina sp., C) Amphistegina sp, UCL coll. Figs 11- 12. Eulinderina sp., Middle Eocene, Yecuatla, Veracruz, Mexico, NHM P51959, 11) axial section; 12) equatorial section. 1 2 3 4 5 8 6 7 8 11 9 12 10 13 14 Plate 6.27 Scale bars:  Figs 1- 2, 5, 12- 14  =  1mm; Fig.  3- 4  =  0.5mm; Figs 6- 11  =  0.25mm. Fig.  1. Neodiscocyclina anconensis (Barker), figured by Vaughan (1945), Eocene, Barbados. Fig. 2. Orbitoclypeus nummuliticus (Gümbel), Middle Eocene, France, UCL coll. Fig. 3. Eoannularia eocenica Cole and Bermúdez, Middle Eocene, figured by Cole and Bermúdez (1944), Pinar del Rio, Province. Fig. 4. Linderina buranensis Nuttall and Brighton, Eocene, France, UCL coll. Fig. 5 Linderina floridensis Cole, Late Eocene, Pakistan, NHM P48503. Fig.  6. Linderina brugesi Schlumberger, Middle Eocene, Qatar, NHM P40266. Fig.  7. Planorbulinella solida Belford, Late Oligocene, Borneo, UCL coll. Figs 8-1 1. Fabiania cassis Silvestri, Late Eocene, 8)  India, solid specimen; 9-1 1) UCL coll.; fig. 9. with Lacazina sp. Fig. 12. Halkyardia sp., Late Eocene, France, UCL coll. Fig. 13. Daviesina langhami Smout, microspheric (note double periphery, regis- tered as Daviesina sp.), Kohat Limestone, Pakistan, NHM coll. Fig. 14. Daviesina langhami Smout, Kohat Limestone, Pakistan, NHM coll., G.136, R.145. Plate 6.28 Scale bars:  Figs 1-4 , 9- 11, 20-2 1 =0.5mm; Figs 5  =  0.25mm; Figs 6-8 , 12-1 9, 22-2 6=1mm. Figs 1- 2. Laffitteina vanbelleni Grimsdale, Eocene, Kourdane, Syria, NHM P40676- 7. Figs 3-4 . Eorupertia incrassata var. laevis Grimsdale, syntype, Middle Eocene, Ain Zalah Well no. 1, Iraq, NHM P40695. Figs 5-6 . Dictyoconus sp., registered as Dictyoconoides cooki (Carter), Eocene, Egypt, NHM P36656. Figs 7- 8. Dictyoconoides kohaticus Davies, Early Eocene, Kohat Shales, India, 7) NHM P22632; 8) NHM P22633. Figs 9- 11. Lockhartia sp., Eocene, France, UCL coll. Fig. 12. Lockhartia haimei (Davies), Paleocene, Qatar, NHM P40156. Figs 13-1 5. Lokhartia diversa Smout, paratypes, Paleocene, Qatar, NHM P40192 (2) Figs 16- 17. Lockhartia conditi (Nuttall), Paleocene, 16)  Qatar, Arabia, NHM P40206; 17)  Paleocene, India, NHM P38261. Figs 18- 19. Palaeonummulites pristina (Brady), syntypes, Calcaire de Namur, Belgium. NHM P35503-4 . Figs 20- 21. Cuvillierina sp., Eocene, Bangladesh, UCL coll. Fig. 23. Dictyokathina simplex Smout, Eocene, Qatar Arabia, NHM P40222. Fig. 22. Lockhartia cf. newboldi (d’Ârchiac et Haime), Early Eocene, Meting-L aki, Pakistan, UCL coll. Fig. 24. Kathina major Smout, paratype, Paleocene, Qatar, NHM P40210. Fig. 25. Kathina delseota Smout, paratype, Paleocene, Qatar Arabia, NHM P40214. Fig. 26. Daviesina khati- yahi Smout, paratype, Paleocene, Qatar, NHM 40242. 1 2 3 4 5 6 7 8 9 10 11 16 13 14 12 15 17 18 20 19 22 23 24 25 264 21 1 3 4 2 6 75 8 10 12 13 14 9 11 15 16 17 18 19 Plate 6.29 Scale bars: Figs 1- 19= 1mm. Fig. 1. Elphidium sp., Late Oligocene, Syria, UCL coll. Figs 2- 4. Eulinderina sp., Middle Eocene, Yecuatla, Veracruz, Mexico, 2) axial section showing thickened lateral walls, NHM P51959; 3) equatorial section with thin walls without pustules, NHM P51958; 4) equatorial section with thick walls with coarse pustules, NHM P51960. Fig. 5. Helicolepidina spiralis (Tobler), equatorial sec- tion, microspheric form, Eocene, El Alto, NW Peru, NHM P302478. Figs 6-1 1. Helicostegina gyralis Barker and Grimsdale, Middle Eocene, 6 -7 ) equatorial sections, topotypes, Yecuatla, Veracruz, Mexico, NHM 51950- 6; 8- 9) axial sections, topotypes, B-f orms, NHM P33356; 10-1 1) axial sections, A-f orm, Sabaneta, Vercruz, Mexico, topotypes, NHM P51952. Figs 12- 13. L. (Nephrolepidina) praemarginata (Douvillé 1908), Late Oligocene, uppermost part of the Mesolouri section, Greece, Wielandt-S chuster coll., 12) FNS 2001z0155/0 010, 13) FNS 2001z0155/0 006. Figs 14-1 6. Helicosteginopsis soldadensis (Grimsdale), topotype, Late Eocene, Soldado Rock Trinidad, NHM P51953-5 , 14) A- form, axial section; 15-1 6) equatorial sections. Figs 17- 18. Eulepidina ephippioides (Jones and Chapman), Oligocene, Kirkuk Well no.19, Iraq, 17) NHM P40667; 18)  axial section, NHM P40668. Fig.  19. Eulepidina sp., Oligocene, Dutch New Guinea, NHM P22790. Plate 6.30 Scale bars: Figs 1, 2-3 , 10, 13, 18- 19 = 1mm.; Figs 4- 9, 11- 12, 14- 17 = 0.5mm. Fig. 1. Sakesaria dukhani Smout, paratype, Paleocene, Qatar, NHM P40203. Fig.  2. Sakesaria ornata Smout, paratype, Paleocene, Qatar, NHM P40205. Fig.  3. Miogypsinella cf. borodinensis Hanzawa, Chattian, Sukau Road Quarry, North East Borneo, UCL coll. Fig. 4. Americogypsina braziliana BouDagher-F adel and Price, para- type, Early Oligocene, figured by BouDagher-F adel and Price (2010b), offshore Brazil, UCL MF57. Fig. 5. Americogypsina americana BouDagher- Fadel and Price, holotype, Early Oligocene, figured by BouDagher- Fadel and Price (2010b), offshore Brazil, UCL MF63. Fig.6. Paleomiogypsina boninensis Matsumaru, Chattian, North East Borneo, UCL coll. Fig.  7. Rotalia trochidiformis Lamarck, Kairabad Limestone, Pakistan, UCL coll. Figs 8- 10. Miogypsinella sp., Late Chattian, Java, UCL coll. Fig. 11. Miogypsinoides for- mosensis Yabe and Hanzawa, Late Oligocene, Corsica, UCL coll. Fig. 12. Miolepidocyclina mexicana Nuttall, Late Oligocene, from BouDagher- Fadel and Price (2010b), offshore Brazil. Fig. 13. Miogypsina triangulata BouDagher- Fadel and Price, holotype, Late Oligocene, figured by BouDagher-F adel and Price (2010b), Brazil, UCL MF142. Fig.14. Miolepidocyclina panamensis (Cushman), Late Oligocene, from BouDagher- Fadel and Price (2010b), Brazil. Fig. 15. Miogypsina gunteri Cole, Late Oligocene, from BouDagher-F adel and Price (2010b), Brazil. Fig. 16. Miolepidocyclina braziliana BouDagher- Fadel and Price, holotype, Late Oligocene, figured by BouDagher- Fadel and Price (2010b), Brazil, UCL MF124. Fig. 17. Neorotalia sp. 1, BouDagher- Fadel and Price, Early Oligocene, from BouDagher- Fadel and Price (2010b), offshore Brazil, UCL MF54. Fig. 18. Americogypsina koutsoukosi BouDagher-F adel and Price, paratype, Early Oligocene, figured by BouDagher- Fadel and Price (2010b), offshore Brazil, UCL MF69. 1 2 3 4 5 7 6 8 9 10 11 12 13 14 17 18 15 16 543 Chapter 7 The Cenozoic Larger Benthic Foraminifera: The Neogene 7.1 Introduction As seen in Chapter 6, the Oligocene-M iocene boundary was not such a significant event for larger benthic foraminifera as the sharply defined Cretaceous-P aleocene bound- ary. Indeed, almost 80% of species survived the boundary into the Miocene. Although most of the Miocene superfamilies are extant, provincialism is prominent at both the generic and specific levels. Following the initial transoceanic migrations of larger ben- thic foraminifera between the American and the Tethyan provinces, facilitated by the series of global sea-l evel regressions, this migration stopped after rising sea-l evel in the early Oligocene separating the American from the other larger benthic provinces (see BouDagher- Fadel and Price, 2010a, b, c, 2013, 2017). As species became geographi- cally isolated, colonizing new areas environmentally similar to each other, they thrived and evolved similar but distinct parallel lineages during the Neogene. Despite showing different evolutionary lineages at the species level, the main line of evolution follows the same patterns as in the Oligocene, and larger foraminifera become very impor- tant biostratigraphical markers globally at this time. The Tethyan seaway between the proto- Mediterranean and the proto-I ndian Ocean became narrower during the Early Miocene, which restricted further migration between the two provinces and resulted in further provincialism. During the Neogene, lepidocyclinids and miogypsinids completely disappeared from America in the late Early Miocene, and they disappeared from the proto- Mediterranean in the Serravallian. Deep water textulariides made their first appearance in America, while new genera of alveolinoids appeared in the Indo- Pacific. The development of the Indo- Pacific as a separate province continued in the Late Miocene. Cycloclypeus continues to range up to present, while there was a considerable proliferation of the calcarinids in the Pliocene. Most of the superfamilies continued to survive with glob- ally spread representatives, except the American deep water Pavonitoidea, which disap- peared completely in the early Pliocene. 7.2 Morphology and Taxonomy of Neogene Larger Benthic Foraminifera In this section, the main superfamilies (see Fig. 7.1) and families of the following three Neogene orders are discussed, namely the: • Miliolida • Rotaliida • Textulariida. 544 Evolution and Geological Significance of Larger Benthic Foraminifera Figure  7.1. The evolution of the Neogene orders (thick lines) and superfamilies (thin lines) of larger foraminifera. ORDER MILIOLIDA Delage and Hérouard, 1896 The miliolides have tests that are porcelaneous and imperforate made of high M g- calcite with fine randomly oriented crystals. They range from the Carboniferous to the Holocene. Superfamily ALVEOLINOIDEA Ehrenberg, 1839 The test is enrolled along an elongate axis, initially being planispiral or streptospiral, or milioline with chambers added in varying planes. Cretaceous to Holocene. Family Alveolinidae Ehrenberg, 1839, emend. Hottinger at al., 1989 The test is free, fusiform, and coiled along an elongate axis (see full description Chapter 6). Early Cretaceous (Aptian) to Holocene. • Alveolinella Douvillé, 1907 (Type species: Alveolina quoyi d’Orbigny, 1826). The test is elongate fusiform, with several rows of chamberlets in axial section. Pre-s eptal The Cenozoic Larger Benthic Foraminifera: The Neogene 545 passages connect adjacent chamberlets, with smaller secondary pre-s eptal passages in later whorls. The final chamber has numerous rows of apertures. Late Middle Miocene (Serravallian) to Holocene (Fig. 7.2; Plate 7.1, fig. 20; Plate 7.2, figs 1- 8). • Flosculinella Schubert, 1910 (Type species: Alveolinella bontangensis Rutten, 1912). The early part of the test is streptospiral and similar to Borelis de Montfort, 1808, but with double rows of chamberlets on the floor of each chamber, one row being smaller than the other. Early Miocene (Burdigalian) to Middle Miocene (mid Serravallian) (Plate 7.2, figs 9-1 2). A C e Third row Septula First row Second row of foramen/alveoles off oramen/ of foramen/alveoles alveoles f B D E F Septal furrows Chambers Figure 7.2 Alveolinella quoyi (d’Orbigny), Port Moresby, Coral Sea, New Guinea: A-B , axial thin sections showing elongate fusiform test, with several rows of chamberlets; C, solid specimen by Banner (1971) showing a parasite/s ymbiont completely enclosed by the whorls, Planorbulinopsis parasitica Banner (e) embedded in the test; D, a solid specimen showing a bored area (f); E- F, schematic drawing showing the single pre- septal passage in E) Flosculinella and F) doubled pre- septal passage of Alveolinella. Scale bars: A, C, D = 0.25mm, B = 1mm. 546 Evolution and Geological Significance of Larger Benthic Foraminifera Superfamily CORNUSPIROIDEA Schultze, 1854 The test is free or attached composed of a globular proloculus, followed by a tubu- lar enrolled chamber. The coiling is planispiral or trochospiral, evolute or involute, and may become irregular. The aperture is simple, at the end of the tube. Lower Carboniferous to Holocene. Family Discospirinidae Wiesner, 1931 The form is discoid with a globular proloculus, followed by a planispirally enrolled test. The later chambers are annular and may be subdivided into small chamberlets. Middle Miocene to Holocene. • Discospirina Munier- Chalmas, 1902 (Type species: Orbitolites tenuissimus Carpenter, 1870). The test is large, fragile, thin and flattened, with a peneropliform early stage, followed by annular chambers subdivided by numerous internal septa that fall short of the anterior wall of each chamber. Middle Miocene to Holocene (Plate 7.3, fig. 10). Superfamily MILIOLOIDEA Ehrenberg, 1839 The test is coiled in varying planes or uncoiled, cylindrical or compressed with partial partitions. Upper Triassic to Holocene. Family Rivieroinidae Saidova, 1981 The test is planispiral, ovate in outline with chambers subdivided by oblique sutures. The aperture is a single curved slit, but may be cribrate. Middle Eocene to Holocene. • Riveroina Bermúdez, 1939 (Type species: Riveroina caribaea Bermúdez, 1939. The test has flattened sides. Chambers are one- half coil in length, subdivided by oblique septula, extending completely across the chamber lumen. The aperture is an arched slit at the end of the final chamber. Holocene. • Pseudohauerinella McCulloch, 1981 (Type species:  Pseudohauerina dissidens McCulloch, 1977). The test has a quinqueloculine early stage and adult planispiral chambers with incomplete subepidermal partitions. The aperture is terminal and cribrate. Holocene. Superfamily SORITOIDEA Ehrenberg, 1839 Chambers are planispiral, uncoiling, flabelliform or cyclical, and may be subdivided by partitions or pillars. Late Permian to Holocene. Family Peneroplidae Schultze, 1854 The test has a closely coiled early stage becoming uncoiled in later stages. Chamber interiors are simple. The aperture is a single rounded and slit-l ike, or multiple. Late Cretaceous to Holocene. The Cenozoic Larger Benthic Foraminifera: The Neogene 547 • Laevipeneroplis Šulc, 1936 (Type species: Peneroplis karreri Wiesner, 1923). The test is compressed and flaring with chambers becoming progressively broader and more curved, but increasing very little in height. The interiors of the chambers are undi- vided. The aperture is multiple near the base of the apertural face, becoming two rows of openings at the end of the apertural face. Miocene to Holocene. Family Soritidae, Ehrenberg, 1839 The test is biconvex, involute-p lanispiral to an uncoiled evolute, flaring, annular dis- coid with partial or complete partitions. Apertures are multiple. Late Cretaceous (Cenomanian) to Holocene. • Androsina Lévy, 1977 (Type species Androsina lucasi, Lévy 1977). The test is tightly coiled, with chambers increasing rapidly in size, strongly curved, but do not become completely annular. Chambers are subdivided by flattened pillars in the median plane. Double annular passages occur in lateral positions. The apertural face has two to four rows of pores. Pleistocene to Holocene. • Androsinopsis Hottinger, 2001 (Type species: Androsinopsis radians Hottinger, 2001). This genus is similar to Androsina but becomes circular at an early stage of ontogeny with heavy equatorial pillars in the annular adult stage. Late Miocene. • Annulosorites Hottinger, 2001. This genus is similar to Sorites (see Chapter 6), but with an involute spiral nepiont. The megalospheric protoconch is large, involute and planispiral with a rapidly expanding flexostyle, ending with a frontal wall bear- ing multiple apertures. The following deuteroconch has multiple apertures and is subdivided by two septula. The microspheric generation is not known. Whereas in Sorites the opposing subepidermal partitions form a continuous septulum in the same chamber, in Annulosorites, this connection is lacking and the two opposing subepidermal partitions form two separate septula. From one chamber to the next, the septula are alternating in a radial disposition. Apertures are large and rounded, forming a single row in the equatorial plane of the shell, in alternating position with the septula. In late, adult growth- stages, the apertures may become separated into two rows. Late Miocene. • Archaias de Montfort, 1808 (Type species:  Archaias spirans de Montfort, 1808 = Nautilus angulatus Fichtel and Moll, 1798), as defined by its type species Archaias angulatus (Fichtel and Moll), extensively emended by Rögl and Hansen, 1984. The test is compressed, planispiral and involute, and may be partially evo- lute in the last whorls, with a thickened middle part and radial endoskeletal ele- ments. Multiple apertures are flanked by irregular free and interseptal pillars. The subepidermal partitions are incomplete, and tests lack exoskeletal structures and no marginal subdivision of the chambers. Middle Eocene to Holocene (Fig. 7.3; see Chapter 6). • Cycloputeolina Seiglie and Grove, 1977 (Type species: Peneroplis pertusus (Forskal) var. discoideus Flint, 1899). Final chambers become circular and subdivided with vertical partitions. The aperture is one or two rows of openings, each bordered by a lip. Miocene to Holocene. 548 Evolution and Geological Significance of Larger Benthic Foraminifera • Fusarchaias Reichel, 1952 (Type species: Fusarchaias bermudezi Reichel, 1952). The test is fusiform, planispiral with numerous chambers and endoskeleton of interseptal pillars. Oligocene to Miocene. • Marginopora Quoy and Gaimard, 1830 (Type species: Marginopora vertebralis Quoyi and Gaimard, 1830). The test is large and biconcave. The embryonic apparatus of the megalospheric test consists of a large deuteroconch that embraces a small pro- toconch, including its wide-o pen flexostyle with an almost cylindrical frontal wall bearing numerous pores. The microspheric test has an early planispiral and pen- eropline stage followed by annular, concentric chambers, with thickened and folded margins. Initially with two layers of annular chamberlets, later low chambers are inserted between them. The annular chambers are subdivided by incomplete vertical septula and partitions. Oblique stolons connect the lateral chamberlets to the cham- bers above and below. The aperture is multiple over the peripheral wall. Miocene to Holocene (Plate 7.2, figs 16- 17; Plate 7.3, figs 1- 5). • Miarchaias Hottinger, 2001 (Type species: Miarchaias meander Hottinger, 2001). The test has a pillared, radial endoskeleton and a radial exoskeleton consisting of short radial partitions (beams). The megalospheric test has a narrow flexostyle. It differs from Archaias in having marginal apertures and complete subepidermal partitions in alignment from one chamber to the next one. Late Miocene (Fig. 7.3). Septa Pillars Apertures Septa Pillars A B Beams Apertures Septula 500 μm C D Figure 7.3. A-B , Archaias sp.; C, D, Miosorites americanus (Cushman), C) Miarchaias sp. with an early peneropline evolute stage and annular chambers, divided by simple radial septula. All figured specimens are from Bahamas (courtesy of G.J. Fischer). The Cenozoic Larger Benthic Foraminifera: The Neogene 549 • Miosorites Seiglie and Rivera, 1976 (Type species:  Orbitolites americana Cushman, 1918). The test is annular and evolute. The microspheric test has an early peneropline evolute stage. The megalospheric protoconch consists of a large, flaring, involute flexo- style and of a semilunar, subdivided, slightly involute deuteroconch (Hottinger, 2001). The following few chambers are reniform. The chambers become annular in the later stage. Annular chambers are divided by simple radial partitions, interpreted as septula, alternating in radial position from one chamber to the next. Oval apertures occur in marginal positions, with crosswise- oblique stolon axes relative to the radius of the test. The genus Miosorites differs from Amphisorus (see Chapter 6) by its narrower apertural face, confined in an equatorial depression, and by its involute embryonic apparatus, where the flexostyle envelops large, lateral surfaces of the megalosphere (see Hottinger, 2001, Fig. 8). No internal pillars are present. (Fig. 7.3) Burdigalian to?Pliocene. • Parasorites Seiglie and Rivera, 1977 (Type species: Praesorites orbitolitoides Hofker 1930). The test is annular and evolute. The microspheric test has an early peneropline stage. The megalospheric test has a subglobular proloculus with a long flexostyle, fol- lowed by an evolute, rapidly flaring peneropline and an annular later stage. Annular chambers are subdivided by simple radial lateral partitions (interpreted by Hottinger, 2001, as exoskeletal beams), that alternate in radial positions in subsequent chambers. The aperture is one to five rows of rounded openings. Late Miocene to Holocene. • Pseudotaberina Eames, 1977, emended, Banner and Highton, 1989 (Type spe- cies: Orbitolites malabarica Carter 1853). The test becomes cyclical in the latest growth of the microspheric form. Megalospheric forms have a large proloculus. Later cham- bers are not embracing and may become cyclical and evolute. Chambers show “stal- agmitic” and “stalactitic” projections/p illars that discontinuously fuse across the chambers away from the lateral walls, producing chamber-s ubdividing structures, dis- tinct from the separated pillars of Archaias (Fig. 7.3; see Chapter 6). The discontinuity of the internal structures allows a distinction to be made between Pseudotaberina and the Cretaceous Larrazetia (see Chapter  5), while Cyclorbiculina possesses true sub- epidermal partitions. Unlike the apertures of Archaias, which are situated in regular, parallel rows, those of Pseudotaberina are scattered over the apertural face and each of them is surround by a projected lip. Early Miocene (Fig. 7.4; Plate 7.4, figs 1- 6, 12). • Sorites Ehrenberg, 1839 (Type species: Nautilus orbiculus Forsskal, 1775). The test is a large discoid, with an early peneropline stage. Annular chambers are divided into numerous curved to rectangular small chamberlets, which are connected to each other and to those in adjacent chambers by stolons. The aperture is a single paired row. Oligocene to Holocene. (Fig. 7.5; Plate 7.3, figs 6-8; Plate 7.8, fig. 3; see Chapter 6). Family Keramosphaeridae Brady, 1884 The test is globular with concentric chambers connected by stolons in the same series as well as those of successive series. Early Cretaceous (Berriasian) to Late Cretaceous (Maastrichtian), and Miocene to Holocene. • Kanakaia Hanzawa, 1957 (Type species:  Kanakaia marianensis Hanzawa, 1957). The test is large, formed of encrusting layers of chambers. Adjacent chambers are 550 Evolution and Geological Significance of Larger Benthic Foraminifera B Peneropline Megalospheric embryont Aperturesstage A Pillars C D Figure 7.4. Pseudotaberina malabarica (Carter), type figures deposited in the NHM: A, solid specimen of a microspheric specimen; B, oblique axial section showing the disposition of the pillars; C, oblique equatorial sec- tion showing the alternating disposition of the foramina and the pillars; D, enlargement of the chambers show- ing the “stalagmitic” and “stalactitic” pillars that discontinuously fuse across the chambers. Scale bars = 1mm. Early peneropline stage Curved ch amberlets Figure 7.5. Sorites sp., a microspheric form showing a peneroplid stage followed by annular series of small chamberlets. The Cenozoic Larger Benthic Foraminifera: The Neogene 551 connected by horizontal stolons, and oblique stolons connect those of successive layers. Early Miocene (Aquitanian). • Keramosphaera Brady, 1882 (Type species:  Keramosphaera murrayi Brady, 1882). The test is globular, with irregular chamberlets, added in concentric unaligned spherical series. The adjacent chamberlets are connected by stolons. Holocene. ORDER ROTALIIDA Delage and Hérouard, 1896 The test is multilocular, with a calcareous wall made of perforate, hyaline lamellar cal- cite. The aperture is simple or with an internal tooth- plate. Triassic to Holocene. Superfamily ASTERIGERINOIDEA d’Orbigny, 1839 The test is trochospiral to planispiral, with a closed umbilicus. Chambers have inter- nal partitions. The aperture is umbilical, and may extend up the apertural face with complex chamberlets at the centre of umbilical side. Late Cretaceous (Santonian) to Holocene. Family Lepidocyclinidae Scheffen, 1932 The test is discoidal, involute, and biconvex with a broad centrum, which grades into a narrow flange. Adauxiliary chambers may be present. The primary spire per- sists into the equatorial layer, or with annular rings of chamberlets that follow the embryont immediately. Stacks of “lateral chamberlets” (cubiculae) occur on each side of the median chamberlets. Pillars may be present between adjacent vertical stacks of cubiculae or scattered in the central region. The chamber walls are per- forated by stolons, but there is no canal system. Middle Eocene to Late Miocene (Early Pliocene?). Subfamily Lepidocyclininae Scheffen, 1932 Representatives of this subfamily have a bilocular or multilocular embryonal stage, surrounded by a thickened wall and adauxiliary chambers. Microspheric tests have an early planispiral coil, while megalospheric tests have a globular protoconch, enclosed or followed by a larger reniform deuteroconch. Post- embryonic chambers evolve from cyclical, arcuate to hexagonal in shape, usually with two or more apertures. The lateral chambers are well differentiated from the equatorial layer and in the advanced forms they are arranged in tiers on either side of the equatorial layer. Surface ornaments and development of pillars are of specific importance. Middle Eocene to Late Miocene (Early Pliocene?), (see Chapter 6). Superfamily NUMMULITOIDEA de Blainville, 1827 The test is planispiral or cyclic, lenticular, multicamerate, with a septal flap and cana- liculated septa. A  spiral marginal cord and spiral canal system are present in early forms, and are modified in advanced forms or replaced by intraseptal canals. Paleocene to Holocene. 552 Evolution and Geological Significance of Larger Benthic Foraminifera Family Cycloclypeidae Galloway, 1933 emend. BouFagher- Fadel, 2002 This family is distinguished by the development of concentric annular chambers, that are wholly evolute, with each chamber divided into numerous chamberlets in the median plane, and with each chamberlet separated from adjacent chamberlets by canaliculated, straight walls. No marginal cord exists, except in the early stages of the microspheric generation. Eocene to Holocene. • Cycloclypeus Carpenter, 1856 (Type species:  Cycloclypeus carpenteri Brady, 1881). A nummulitoid with a nepionic morphology like Heterostegina, but with a final growth stage with cyclic chambers. No alar prolongations occur. Early Oligocene to Holocene (Fig. 7.6; Plate 7.5, figs 15- 19; Plate 7.6, figs 1- 11; Plate 7.7, fig. 9; see Chapter 6). • Katacycloclypeus Tan Sin Hok, 1932 (Type species: Cycloclypeus (Katacycloclypeus) martini Tan Sin Hok, 1932). Tan Sin Hok (1932), in his description of Katacycloclypeus, assigned it as a new subgenus of Cycloclypeus. However, there is no direct evidence of intergradation, either in the modelling of the test or in the embryonic structure between Cycloclypeus and Katacycloclypeus. The latter has a trilocular embryont and a thin test with a central umbo, surrounded by several annular inflations of the solid lateral walls. The stratigraphic range is also quite dif- ferent. Katacycloclypeus is confined to the upper Tf1 and Tf2 Letter Stages of the Middle Miocene of the Indo- Pacific, while Cycloclypeus ranges from the Oligocene to the Holocene throughout all the tropics. Therefore, the two forms should be con- sidered to be generically different. Middle Miocene (late Langhian to Serravallian) (Plate 7.6, figs 12- 13; Plate 7.7, figs 1- 8; Plate 7.8, fig. 7). D Pr A B Figure 7.6. A, thin section of Cycloclypeus indopacificus Tan Sin Hok, Middle Miocene, Nias, Sumatra, UCL coll., scale bar  =  0.25mm; B, SEM of Cycloclypeus carpenteri (Brady), Holocene, from Hottinger (2006) Bikini, Pacific (abbreviations: d: deuteroconch; f 1: foramen of protoconch; f 2: foramen of deutero- conch; isc: intraseptal canal system; pr: proloculus; s: septum; sl: septulum; st: stolon (Y- shaped)). The Cenozoic Larger Benthic Foraminifera: The Neogene 553 Family Nummulitidae de Blainville, 1827 The test is planispiral, involute or evolute, with septal, marginal and vertical canals. Paleocene to Holocene. • Heterocyclina Hottinger, 1977 (Type species:  Heterostegina luberculata Mobius, 1880). The test is discoidal, planispiral, and evolute with whorls becoming annular in the adult. Chambers are divided by septula into rectangular chamberlets. Suture canals are unbranched and the stolon system is L- shaped. Holocene. • Bozorgniella Rahaghi, 1973 (Type species:  Bozorgniella qumiensis Rahaghi. 1973). The planispiral, involute test has two and a half whorls. Sutures are straight to slightly curved near the end, but externally obscured by pustules. Early Miocene (Aquitanian). • Planoperculina Hottinger, 1977, emended Banner and Hodgkinson, 1991 (Type species: Operculina heterostegnoides Hofker, 1933). The test is wholly evolute, with incomplete chamber partitions. Holocene. • Radiocycloclypeus Tan Sin Hok, 1932 (Type species:  Cycloclypeus neglectus Martin var. stellatus Tan Sin Hok, 1932). The test is stellate with irregular rays. The embryonic apparatus is enclosed by a thick wall and consists of a proto- conch surrounded by a large deuteroconch. It is followed by about six thin- walled embryonic chambers that gradually increase in length and finally become annu- lar in the later part of the test. Chambers are divided into rectangular chamber- lets, which alternate in position. Early Miocene (Burdigalian) to Middle Miocene (Serravallian). • Operculinella Yabe, 1918 (Type species:  Amphistegina cumingii Carpenter,1860). The last true Nummulites spp. became extinct at the top of the Td “Letter Stage” with Nummulites fichteli Michelotti 1841 from the upper Early Oligocene of Italy. Contrary to the opinions of S. Cole (in Loeblich and Tappan, 1964) and Loeblich and Tappan (1988), Nummulites can be distinguished from Operculinella. Eames et al. (1962) illustrated a simple Nummulites vascus Joly and Leymerie (their plate 1, figures A, B) to compare with Operculinella cumingii (Carpenter) (Palaeonummulites nomen oblitum). The strong dimorphism seen between microspheric and megalo- spheric forms of Oligocene specimens of Nummulites is never seen in Operculinella (where the microspheric and megalospheric generations are externally identical). The presence of trabeculae in Nummulites and their absence from Operculinella is note- worthy, but, most importantly, the diameter of the megalospheric protoconch of Nummulites (in both simple and complex forms) is much greater than the diameter of the proloculus of Operculinella. The megalospheric loosely coiled Operculinella (e.g. Operculinella cumingii) persists to the Holocene but the large protoconch of true Nummulites does not occur beyond the Early Oligocene. Oligocene to Holocene (Fig. 7.7). Superfamily PLANORBULINOIDEA Schwager, 1877 The test is trochospiral in early stages, but later may be uncoiled and rectilinear, or bise- rial or may have many chambers in the whorl. Intra- to extra-u mbilical apertures occur, 554 Evolution and Geological Significance of Larger Benthic Foraminifera A B D C E F Figure  7.7. Comparison between:  A-B , Nummulites vascus Joy and Leymerie, Oligocene, Cyrenaica, NHM P44493; C- D, Operculinella cumingii (Carpenter), Holocene, Port Moresby, Papua, NHM coll.; E- F, Palaeonummulites kugleri (Vaughan and Cole), Oligocene, Falling Waters State Park, Chipley, Florida, UCL MF323. Scale bar = 2mm. and additional equatorial apertures may be present. Early Cretaceous (Berriasian) to Holocene. Family Planorbulinidae Schwager, 1877 The test is free or attached. The early stage is trochospiral, but later becoming dis- coid, cylindrical or conical. The aperture is single or multiple. Eocene to Holocene. • Planorbulinopsis Banner, 1971 (Type species:  Planorbulinopsis parasitica Banner, 1971). The test is attached, with the early part trochospiral, the spiral side evo- lute, and the umbilical side involute. The umbilicus is open and deep. Later cham- bers alternate with those of the preceding whorl. The walls are coarsely perforate. Holocene (Fig. 7.2). The Cenozoic Larger Benthic Foraminifera: The Neogene 555 Family Victoriellidae Chapman and Crespin, 1930 The test is attached or may be free in the juvenile stage, with a trochospiral early stage, later becoming an irregular mass of chambers. Cretaceous (Santonian) to Holocene. Subfamily Carpenteriinae Saidova, 1981 The test is attached, trochospiral throughout, and planoconvex with a large aperture, open in the umbilicus. Paleocene to Holocene. • Carpenteria Gray, 1858 (Type species: Carpenteria balaniformis Gray, 1858). The test has a carinate periphery, and is planoconvex with a flat spiral side and distinct rims or keels, a strongly convex, distinctly perforate umbilical side surrounded by thick pillars. Late Eocene to Holocene (Plate 7.9, figs 3, 6- 7; Plate 7.10, figs 1- 4)). Subfamily Rupertininae Loeblich and Tappan, 1961 The test is attached with a distinct flattened disk, but coiling grows out away from the site of attachment. Late Cretaceous to Holocene. • Rupertina Loeblich and Tappan, 1961 (Type species; Rupertia stabilis Wallich. 1877). The test grows upright around a central column, and is trochospiral in the early stage, later becoming more loosely coiled. Miocene to Holocene (Plate 7.9, figs 8-9 ). Superfamily ROTALIOIDEA Ehrenberg, 1839 The test is involute to evolute, initially trochospiral or planispiral, commonly with many chambers in numerous whorls. As new chambers are added, septal flaps attach to previous the apertural face and enclose radial canals, fissures, umbilical cavities, and intraseptal and subsutural canals. The wall is made of perforate hyaline calcite, that is generally optically radial in structure. Primary apertures are single or multiple. Small openings into the canal system may occur along the sutures. Late Cretaceous (Coniacian) to Holocene. Family Calcarinidae Schwager, 1876 The test is enrolled with protruding canaliculated spines. Free living, but they live mostly adhered to algae by a kind of plate secreted at the end of one, or more rarely two, spines (Röttger and Krüger, 1990). Rows of areal foramina are found mostly near the base of the septa. Late Cretaceous (Maastrichtian) to Holocene. • Baculogypsina Sacco, 1893 (Type species: Orbitolina concava Lamarck var. sphae- rulata Parker and Jones, 1860). The test is biconvex, with 5 to 7 canalicular spines roughly in a single plane, radiating from the spiral juvenarium by originating as extensions of an intraseptal interlocular space (Hottinger, 2006), and continuing to enlarge with growth. Following the spiral juvenarium, elongated supplemental chamberlets overgrowing the base of the spine, are aligned in a chessboard pattern and connected to the spine canals. Pleistocene to Holocene (Plate 7.10, figs 5- 7). 556 Evolution and Geological Significance of Larger Benthic Foraminifera • Baculogypsinoides Yabe and Hanzawa, 1930 (Type species:  Baculogypsinoides spi- nosus Yabe and Hanzawa, 1939). The test is globular with protruding canalicular spines, but unlike Baculogypsina they are not in a single plane, but instead they give a triangular to tetrahedral appearance to the test. Holocene. • Calcarina d’Orbigny, 1826 (Type species: Nautilus spengleri Gmelin, 1791). The test is biconvex, trochospirally coiled throughout, with many blunt or bifurcating radial spines. There is an enveloping spiral canal system consisting of a layer formed by numerous canals that run through the spines. There are ten to twenty chambers in the final whorl, each containing passages to the canal system which is connected to the outside via many openings on the test surface. Pustules and spinules cover the whole test with the umbilicus filled by pillars. Pliocene to Holocene (Plate 7.5, figs 11- 14; Plate 7.10, fig. 8). • Quasirotalia Hanzawa, 1967 (Type species: Quasirotalia guamensis Hanzawa, 1967). The test is planoconvex with a flat spiral side. Layers of chambers are added to the periphery and the umbilical side. Pillars fill the umbilicus. Coarse pores fill the thick calcareous walls. Pliocene (Fig. 7.8). • Schlumbergerella Hanzawa, 1952 (Type species:  Baculogypsina floresiana Schlumberger, 1896). The test is globular with slightly projecting spines or tuber- cles. In the microspheric generation there is a planispiral coil of about two whorls, with spines. Megalospheric forms have an embryonic apparatus composed of three chambers with tetragonal spines, formed from the outer chamber walls. There are numerous dome- like lateral chamberlets, communicating with each other through stolons in the lateral walls and by large pores with chambers of the same radial wall, produce a globular test. Numerous pillars are formed between the radiating rows of chamberlets. Unlike Baculogypsinoides, the corners of the tetrahedral test in this genus do not support prominent spines and the canal system is much reduced, as none of the pillars of the latter have an internal canal system, but one which only occurs on the outer edge of the spines. The walls are coarsely perforated. Pleistocene to Holocene. Family Chapmaninidae Thalman, 1938 The test is conical, with a trochospiral initial part, followed by a uniserial part and a tubular apertural system. Chambers may be annular in the adult part of the test. Septa are invaginated into tube pillars. Late Paleocene to Late Miocene (Tortonian). • Tenisonina Quilty, 1980 (Type species: Tenisonina tasmaniae Quilty, 1980). The test is planoconvex, an early trochospire is followed by annular chambers, divided into curved chamberlets, but the final chambers are undivided. Early Miocene. Family Miogypsinidae Vaughan, 1929 The test is flattened to biconvex. The microspheric form has a trochospiral or planispi- ral early spire, while the megalospheric form has a bilocular embryonal stage followed by a fan of median chamberlets. Middle Oligocene to Middle Miocene. The Cenozoic Larger Benthic Foraminifera: The Neogene 557 Po Po Pi Pi R Pi Po Pi Po Figure 7.8. Photomicrograph of thin sections of Quasirotalia guamensis Hanzawa, Sulawesi, SEA coll., showing a plano-c onvex test with a flat spiral side, pillars (Pi) filling the umbilicus and coarse pores (Po) fill- ing the thick calcareous walls and rodophyte algae (R). Scale bars = 0.5mm. • Heterosteginoides Cushman, 1918 (Type species: Heterosteginoides Cushman, 1918). The test is similar to Miolepidocyclina, with the nepiont centrally placed, however, the nepionic spiral is much longer, overriding in its later part a variable number of equatorial chambers. Early Miocene (Aquitanian to early Burdigalian). • Lepidosemicyclina Rutten, 1911 (Type species:  Orbitoides (Lepidosemicyclina) the- cideaeformis Rutten, 1911). A  roughly circular test, with an embryonic apparatus made of a spherical protoconch and a reniform deuteroconch, that has a tendency to become enlarged in most advanced forms. Two sets of planispiral periembryonic chambers surround the embryo, the larger primary spiral and three unequal second- ary spirals. The equatorial chamberlets are at first ogival, then rhombic and finally distinctly hexagonal. Early Miocene (Burdigalian) (Plate 7.9. fig.  15; Plate 7.11, fig. 12; Plate 7.12, fig. 11). • Miogypsinodella BouDagher- Fadel et  al., 2000 (Type species:  Miogypsina (Miogypsina) primitiva Tan, 1936). The embryont coil is similar to that of Miogypsinoides (see Chapter 6), it is virtually planispiral, but there is only one whorl around the megalospheric proloculus, and a septal canal system is present. However, the lateral walls have gaps between the lamellae, which begin to split apart and form 558 Evolution and Geological Significance of Larger Benthic Foraminifera the beginnings of lateral chamberlets. This splitting results in thick-w alled irregular chamberlets, unlike the regularly formed, stacked chamberlets of Miogypsina (see Chapter  6). Late Oligocene (late Chattian) to Middle Miocene (Langhian) (Plate 7.11, figs 1-4 ). • Miolepidocyclina Silvestri, 1907 (Type species: Orbitoides (Lepidocyclina) burdigalen- sis Gümbel, 1868). The embryonic apparatus, consisting of large protoconch and deuteroconch, is surrounded by a thick wall. The megalospheric nepiont is similar to that of Miogypsina, with no coil around the proloculus but 2- bidirectional coils around the proloculus. However, the nepiont is centrally placed, instead of being at the edge of the test, as in Miogypsina. Early Oligocene (Rupelian, P20) to Early Miocene (Burdigalian) (Plate 7.11, figs 14-1 5; Plate 7.13, fig. 21). • Tania Matsumaru, 1990 (Type species: Tania inokoshiensis Matsumaru, 1990). The embryonic apparatus has a globular protoconch and reniform deuteroconch in megalospheric generations, with two unequal sets of spiral nepionic chambers sit- uated along the outer side of deuteroconch. Tania differs from Miogypsinoides in having well developed lateral chambers. It differs from Miogypsina by the arrange- ment of the embryonic chambers in the apical portion and by the development of hexagonal to spatulate chambers. It is distinguished from Lepidosemicyclina by the arrangement of embryonic chambers and from Miolepidocyclina and Miogypsinita (see Chapter  6) in having an apical embryonic apparatus. Early Miocene (Aquitanian). Superfamily ACERVULINOIDEA Schultze, 1854 The test is trochospiral to discoidal, commonly with raspberry- like (or framboidal) early chambers, but with an encrusting later part consisting of numerous irregularly formed chambers. Paleocene to Holocene. Family Acervulinidae Schultze, 1854 The early stage is followed by spreading chambers with one or more layers. Mural pores act as apertures. Paleocene to Holocene. • Acervulina Schultze, 1854 (Type species: Acervulina inhaerens Schulze, 1854). Early chambers are coiled, later they are vermiform and irregularly arranged. The aperture is made up of coarse perforations. Miocene to Holocene. • Alanlordia Banner and Samuel, 1995 (Type species: Alanlordia niasensis Banner and Samuel, 1995).The test is biconvex, the proloculus is followed by a single, nearly planispiral whorl of triangular chambers each successively linked by a single basal, septal aperture, but in which multiple, cribrate, pore-l ike apertures develop in their distal, outermost walls. The initial whorl is succeeded both dorsally and ventrally by layers of small chambers which are added in radial rows to form successive layers of small chambers. These chambers communicate with succeeding chambers in the next layers by cribrate, pore- like, small apertures. Radial pillars may develop in the median plane, both ventrally and dorsally. It is similar to Wilfordia (see Chapter 6) but the latter has no true initial spire. The closest homeomorph for Alanlordia (especially The Cenozoic Larger Benthic Foraminifera: The Neogene 559 A. niasensis) is Vanderbeekia (see Chapter 5), which has very similar pillaring and dorsal thickening, however, Vanderbeekia (like its close relatives Sirtina, Irunites and Neurnannites) appears to have had a distinct medial layer of thick- walled chambers, unlike the very short, single initial coil of Alanlordia. Middle Miocene (Serravallian, Tf2) to Late Pliocene (Plate 7.4, figs 7- 11). • Borodinia Hanzawa, 1940 (Type species: Borodinia septetrionalis Hanzawa, 1940). The test has more than one encrusting layer, with chambers alternating in position in successive layers. Early Miocene (Aquitanian) (Plate 7.9, figs 1- 2). • Ladoronia Hanzawa, 1957 (Type species:  Acervulina (Ladoronia) vermicularis Hanzawa. 1957). Early chambers are clustered in a framboidal arrangement, fol- lowed by irregular elongate to vermiform chambers. Chambers of successive lay- ers are connected by fine pores, but are neither aligned nor alternating in position. Miocene. • Planogypsina Bermúdez, 1952 (Type species: Gypsina vesicularis var. squamiformis Chapman, 1901). The test is very thin, having globular early chambers followed by a single layer of irregular elongate to vermiform chambers. The aperture is made up of septal pores. Miocene to Holocene. Family Homotrematidae Cushman, 1927 The test is attached, with a trochospiral early stage, later chambers growing in a mas- sive branching structure. Eocene to Holocene. • Homotrema Hickson, 1911 (Type species:  Polytrema cylindrica Carter, 1880). A homotremid form, with a four-c hambered embryo and conical projections or erect branches. Early chambers occur in clustered arrangement, but later in numerous layers have large irregular passages. The aperture is made up of larger perforations. Miocene to Holocene (Plate 7.9, figs 4- 5). • Miniacina Galloway, 1933 (Type species:  Millepora miniacea Pallas, 1766). Megalospheric forms have a three-c hambered juvenile stage, while microspheric forms have a trochospiral early stage. Following the free early stage, the attach- ment surface is narrow, and from this surface arise vertical irregularly branch- ing structures with pillar- pore chambers (calyces) surrounding the central core. Adult forms have one to multiple rounded apertures with a bordering lip at the end of the branches. Early Miocene (Aquitanian) to Holocene (Plate 7.9, figs 10- 13). ORDER TEXTULARIINA Delage and Hérouard, 1896 The tests of these agglutinated foraminifera are made of foreign particles bound by organic cement. They range from lower Cambrian to Holocene. Superfamily ATAXOPHRAGMIOIDEA Schwager, 1877 Members of this family have a multilocular, trochospiral test to biserial or uniserial in later stages. Middle Triassic to Holocene. 560 Evolution and Geological Significance of Larger Benthic Foraminifera Family Alveovalvulinidae Seiglie, Fleisher and Baker, 1986 The test is trochospiral to triserial and uniserial. Chambers have alveoli conjoined with vertical radial partitions or “tubiliform connections”. Apertures are interiomarginal or terminal. • Alveovalvulina Brönnimann, 1951 (Type species: Alyeovalvulina suteri Brönnimann, 1951). The test is trochospiral to triserial, with an interiomarginal aperture. Late Early Miocene to Early Pliocene. • Guppyella Brönnimann, 1951 (Type species:  Goesella miocenica Cushman, 1936). The later stage of the test is uniserial, and the aperture is terminal and circular. Late Early Miocene to Holocene. • Jarvisella Brönnimann, 1953 (Type species:  Jaryisella karamatensis Brönnimann, 1953). A trochospiral to triserial test that has chamber interiors subdivided by verti- cal folding of the outer wall, forming double- walled septula. Late Early Miocene to Middle Miocene. Family Textulariellidae Grönhagen and Luterbacher, 1966 Tests have an early trochospirally enrolled stage, which later is reduced to triserial or uniserial. The chambers are overlapping and the wall is agglutinated with alveoles. Miocene. • Cuneolinella Cushman and Bermúdez, 1941 (Type species:  Cuneolinella lewisi Cushman and Bermúdez, 1941). The test is flattened, with a later stage that is bise- rial and compressed. Chambers increase rapidly in breadth so that the test becomes flabelliform. The aperture is a multiple row of openings. Middle Miocene. • Textulariella Cushman, 1927 (Type species:  Textularia barrettii Jones and Parker, 1876). The later stage of the test is biserial, with the interior of the chambers having numerous vertical partitions that are anastomising inward to form tiny alveoles. The aperture is a low arch. Miocene to Holocene. Superfamily LITUOLOIDEA de Blainville, 1825 Members of this superfamily have a multilocular, rectilinear and uniserial test. The early stage has plani- (strepto- ) or trochospiral coiling. The peripheries of the chambers have radial partitions, but centrally there are either no or scattered, separated pillars. The aperture is simple or multiple cribrate. Early Jurassic (Sinemurian) to Holocene. Family Cyclamminidae Marie, 1941 The test is involute with alveolar walls. The aperture is near the septal face. Jurassic to Holocene. • Cyclammina Brady, 1879 (Type species:  Cyclammina cancellata Brady, 1879). The test is planispiral, and flattened. The walls are thick, with an alveolar subepider- mal meshwork of a thickness exceeding that of the chamber lumen. The primary The Cenozoic Larger Benthic Foraminifera: The Neogene 561 aperture is a basal sutural slit, and the supplementary apertures are areal in the aper- ture face and septa and bordered by a lip. Paleocene to Holocene (Fig. 7.9). • Reticulophragmium Maync, 1952 (Type species:  Alveolophragmium venezuelanum Maync, 1952). The wall is thick with an alveolar hypodermis. The septa are solid and the aperture is situated in the basal suture with a lip on the upper side only. Paleocene to Holocene (Fig. 7.9). Superfamily PAVONITINOIDEA Loeblich and Tappan, 1961 The early stage of the test is coiled, triserial to biserial or uniserial. The interiors of the chambers are partially divided by numerous vertical partitions (beams) or septula, that project downwards from the roof and rarely may have a few connecting horizontal partitions (rafters). Late Cretaceous and Oligocene to Pliocene. Supplementary apertures A B C Primary aperture Lip Alveolar hypodermis D E Imperforate septa Figure 7.9. A- C, Cyclammina cancellata Brady, topotypes from Challenger Station 168, northwest flank of the Hikurangi Trench, NHM 1964.12.9.2- 4, showing the alveolar hypodermis and the lipped aperture and areal cribrate aperture; D-E , Reticulophragmium orbicularis (Brady), topotypes from Challenger Station 323, south flank of the Rio Grande Rise, NHM P1964.12.9.17-1 8, showing basal aperture, imperforate septa and alveolar hypodermis. Scale bars = 0.5mm. 562 Evolution and Geological Significance of Larger Benthic Foraminifera Family Pavonitinidae Loeblich and Tappan, 1961 The test is palmate and triangular in section. The aperture is terminal single or multi- ple. Oligocene to Pliocene. Subfamily Pavonitininae Loeblich and Tappan, 1961 The test is triserial, biserial or uniserial. Chambers are undivided by secondary septula. Oligocene to Miocene. • Pseudotriplasia Malecki, 1954 (Type species:  Pseudotriplasia elongata Małecki, 1954). The test is triangular in thin section with concave sides, uniserial throughout. The aperture is cribrate. Miocene. Superfamily TEXTULARIOIDEA Ehrenberg, 1838 The test is trochospiral, biserial or triserial in early stages, but later may be uniserial or biserial. Walls are agglutinated and canaliculated. Early Jurassic (Sinemurian) to Holocene. Family Textulariidae Ehrenberg, 1839 The early stage is biserial, but may be reduced to uniserial. The aperture is single or multiple. Paleocene to Holocene. Subfamily Tawitawiinae Loeblich and Tappan, 1961 The test is biserial throughout, compressed and palmate. Chambers are subdi- vided by short septula. Walls are thin and canaliculate. Apertures are multiple. Holocene. • Tawitawia Loeblich, 1952 (Type species: Textularia immensa Cushman, 1913). The chambers are strongly overlapping in the plane of biseriality. Holocene. 7.3 Biostratigraphy and Phylogenetic Evolution During the Neogene different faunal assemblages dominated different dispo- sitional environments, however, the most dominant fossils in all formations are warm water, shallow marine groups including larger benthic foraminifera, her- matypic corals and coralline algae. Tracing the stratigraphic distribution of the larger benthic foraminifera allows the understanding of the impact of climate, tectonic activity and volcanism on long- term (i.e. millions years) evolution of these shallow- water carbonate platforms (Courgeon et al., 2016, 2017; Gold et al., 2017a, 2017b). Larger foraminifera and planktonic foraminifera overlap in occurrence in many localities allowing direct comparison of larger foraminifera “letter stages” biozones with oceanic planktonic scales (BouDagher- Fadel, 2002; BouDagher- Fadel, 2013, 2015; Sharaf et al., 2013). The Cenozoic Larger Benthic Foraminifera: The Neogene 563 The most important superfamilies that dominated Tethyan facies are the Nummulitoidea and the Soritoidea (see Fig. 7.10, and Charts 7.1- 7.3), with the latter dominating the assemblages from Late Miocene to present day. However, in order to fully understand the Neogene, all three of the main groups need to be studied, namely: • the porcelaneous miliolides, • the calcareous rotaliides and • the agglutinated textulariides. Also, as in the Cenozoic, bio-p rovincialism was strongly expressed in the Neogene, and the “Letter Stage” biostratigraphic sub-d ivisions are particularly important in the study of SE Asian assemblages (see Chat 7.1). Figure 7.10. A schematic representation of the range and diversity of the main superfamilies in the Neogene (see also Charts 7.2 and 7.3 for details). 564 Evolution and Geological Significance of Larger Benthic Foraminifera 7.3.1 The Letter Stages for the Neogene of SE Asia As introduced in Chapter  6, the “Letter Stages” subdivision of the Indo- Pacific Paleocene and Neogene (Leupold and van der Vlerk, 1931; Adams, 1970; Chapronière, 1984; BouDagher-F adel and Banner, 1999; BouDagher- Fadel, 2008) is based on larger foraminifera, some of which have also been used as range fossils in western hemisphere stratigraphy (Barker and Grimsdale, 1936; Caudri, 1996). The details of these stages and their type fossils are given in Chart 7.1. Many of the Neogene species are short lived and biostratigraphers have relied on them to date the Tethyan carbonates. The distinction between Oligocene and Early Miocene parts of the Te “stage” is drawn on the occurrence of Miogypsinella borodi- nensis (Plate 7.14, fig. 3) in the former, and true Miogypsina and Miogypsina tani (Plate 7.8, fig. 1) in the latter. The base of the Tf1 stage is marked by the disappearance of Austrotrillina striata (Plate 7.1, figs 4- 7), Eulepidina spp. and Heterostegina (Vlerkina) borneensis (Plate 7.8, figs 8-9 ). The latter may be restricted in Melanesia to the lower part of the Te interval, but evidence from Borneo has showed that it ranges through- out the Te interval in the eastern region as a whole (BouDagher-F adel et al., 2001). Austrotrillina howchini (Plate 7.1, figs 1-3 , 10, 12)  replaces A.  striata and marks the beginning of Tf1 stage (see Chart 7.1). The top of the lower Tf1 at the base of the Middle Miocene is defined by the extinc- tion of Lepidosemicyclina sp. (Plate 7.11, fig.  12), while the top of the Tf1 stage is defined by the disappearance of Austrotrillina with the extinction of the youngest spe- cies, A. howchini (Plate 7.1, figs 1- 3, 10, 12), and the top of the Tf2 stage is marked by the disappearance of Katacycloclypeus (Plate 7.6, figs 12- 13; Plate 7.7, figs 1-5 ) and Planorbulinella solida (Plate 7.5, figs 2-5 ) (see Chart 7.1). However, BouDagher- Fadel and Lokier (2006) emended the correlation of BouDagher- Fadel and Banner (1999) by extending the age of Katacycloclypeus (K. annulatus, Plate 7.16, figs 12- 13; Plate 7.7, figs 1- 5) down into the Tf1 stage (Langhian) and that of Flosculinella (Plate 7.2, figs 9- 12) up into the Tf2 stage (Serravallian). The Letter Stage Tf3 (or Upper Tf stage of some authors) is remarkable only by what is does not contain. Faunas of this age are depleted in larger foraminifera with only rare Lepidocyclina, sometimes Cycloclypeus and, in shallower settings with quieter environments, dwelling on algal substrate, sand, dead coral or seagrass, Alveolinella quoyi (Plate 7.2, figs 2-4 ), Marginopora (Plate 7.2, figs 16-1 7; Plate 7.3, figs 1-5 ), Operculina (Plate 7.13, fig. 6; Plate 7.13, fig. 2), Amphistegina (Plate 7.14, fig. 1) and a few other long ranging species. Alveolinella quoyi arose from its ancestral A. praequoyi (Plate 7.1, fig. 20; Plate 7.2, figs 6- 7), referred to by previous authors as A. fennemai (Plate 7.2, fig. 8), at the base of the Tf3, which can be correlated with the horizon at which Katacycloclypeus goes extinct, and the base of Zone N13 (Late Serravallian). The same horizon is marked by the first appearance of Marginopora vertebralis (Plate 7.2, figs 16-1 7), which ranges up into the Holocene. The top of the Tf Letter Stage was defined by BouDagher-F adel and Banner (1999) on the extinc- tion of Lepidocyclina. The Late Miocene Tg does not contain any distinct larger benthic foraminiferal assemblages. The Letter Stages Tg and Th represent shallow marine biohermal car- bonates that changed from mixed coral and coralline algal boundstones with larger The Cenozoic Larger Benthic Foraminifera: The Neogene 565 foraminifera grainstones to more dominant coral reefs with a marked increase in Halimeda green algae. The latter is often preserved in recognisable forms, or as an increase in micrite and fine bioclastic products from the early breakdown of its arago- nitic platelets. Larger foraminiferal grainstones are rare after this event, as deeper pho- tic Cycloclypeus facies or minor Operculina/A mphistegina or Alveolinella calcarenites. Most well and field sections encountering limestones of this later Miocene or Pliocene age in Southeast Asia do not sample rocks composed of larger foraminifera tests. 7.3.2 The Miliolides of the Neogene The miliolides of the Neogene fall into four superfamilies: • the Alveolinoidea • the Milioloidea • the Soritoidea • the Cornuspiroidea Most of the Paleogene Aveolinoidea disappeared at the Eocene-O ligocene bound- ary and the only survivor of the terminal Middle Eocene extinction of the alveolinids was the simple yet cosmopolitan form Borelis, which in turn survived the Oligocene- Miocene boundary and continues to thrive to the present day. Over this period, alveo- linids underwent a very slow evolution, with an incremental increase in the length of the test from pole to pole, changing from globular, ovoid to elongated spindle- shaped, and with an equally incremental increase in the number of secondary chamberlets (“mansardes”) from zero in Borelis through one in Flosculinella to several rows of chamberlets in Alveolinella. Parallel lineages have each developed with the acquisition of an additional row of chamberlets in the Borelis- Flosculinella descent. This must have occurred at least twice during Te- Tf interval (Early to Middle Miocene). An example of one of these lineages is seen with the appearance of Flosculinella reicheli within the Early Miocene (upper Te) and of F. bontangensis (Plate 7.2, figs 9- 11) in the early Middle Miocene (just at the top of the Te stage). The latter gave rise to Alveolinella praequoyi (Plate 7.1, fig. 20) in the Serravallian, at the base of Tf2 stage. A. praequoyi has early whorls akin to F. bontangensis, but in the latter the chamberlets of each whorl are covered by at least 2 layers of smaller chamberlets. In Alveolinella quoyi (Plate 7.2, figs 2- 4), in the Tf3 stage, all of the whorls have a multiple layer of chamberlets. Flosculinella bontangensis grades into Alveolinella forms about the same time as the disappearance of Austrotrillina (see Chart 7.1). This evolution parallels to that described by BouDagher- Fadel and Lord (2000) for Lepidocyclina sensu lato. The genus Austrotrillina of the Milioloidea is essentially a Quinqueloculina (a simple small milioline), with a distinctly alveolar wall, presumably to harbour symbiotic algae. The evolution in the Indo-P acific realm of A. paucialveolata (see Chapter 6) - striata (Plate 7.1, figs 4- 7) - asmariensis (Plate 7.1, figs 8-9 ) - howchini (Plate 7.1, figs 10, 12) is a straightforward lineage with a gradually increasing number and complexity of alveo- lae. A. asmariensis is distinguished by its closely spaced, narrow alveoles, which were present in a single series and did not bifurcate peripherally as they would have done in 566 Evolution and Geological Significance of Larger Benthic Foraminifera A. howchini. In many places transitional forms, in which the alveoles in the later whorls are simple and undivided (as in A.  asmariensis) while the first whorls have thicker, bifurcating and more complex alveoli (as in A. howchini), co- exist with specimens typ- ical of A. asmariensis with narrow alveoles. The gradual evolution from the primitive form A. asmariensis into more advanced form A. howchini occurred in the lower Tf1 stage and only the advanced form A. howchini persisted into the upper Tf1 stage, where it disappears completely near the top (see Chart 7.1). The top of the lower part of lower Tf (Tf1) stage is defined by the extinction of the youngest species, Austrotrillina howchini, while the extinction of Flosculinella bontan- gensis occurred within the lower part of Tf2 where it occurs parallel to the appear- ance of Alveolinella praequoyi. BouDagher- Fadel and Banner (1999) summarized the ranges of individual species, and the series Borelis pygmaeus - Flosculinella bontan- gensis - Alveolinella praequoyi - A. quoyi seems to be the central evolutionary lineage. Borelis pygmaeus persists to the top of the Te stage (BouDagher- Fadel and Banner, 1999)  where it is suddenly replaced by Flosculinella bontangensis, which disappears within the early part of the Tf2 stage. On the other hand, it evolves in the Tf2 into A. praequoyi, which is then replaced by Alveolinella quoyi, in the Tf3 stage. The imperforate Soritoidea (Fig.  7.11) are characterized by peneropliform- planispiral, flabelliform and annular tests. They evolved by enlarging their apertural face and multiplying their apertures with increasing test size during ontogeny and phylogeny (Hottinger, 2001). This resulted in a discoidal shape with the aperture con- stituting the total shell margin (Hottinger, 2000), e.g Marginopora, or in an involute- fusiform shell in which the apertural face is enlarged at the poles by “polar torsion” (Leppig, 1992), e.g. Archaias (see Figs 7.3, 7.11). They are divided into two subfamilies (Loeblich and Tappan, 1987), the Soritinae and Archaiasinae; the soritids have been separated from the archaiasines by the presence of lateral, subepidermal partitions (Fig. 7.3; Henson, 1950) or intradermal plates (Seiglie et al., 1976), in contrast to a pil- lared endoskeleton in the archaiasines (Hottinger, 2001). However according to Levy (1977) some forms, such as Cyclorbiculina (Plate 7.3, figs 11- 12), have both elements combined in the same shell, and it is therefore difficult to classify in such a system. For this reason, Hottinger (2001) recommended using the pattern of the stolon axes, radial or crosswise oblique, as the general, overriding character dividing the two groups. In this book, no assignations to subfamilies of the Soritidae are given at all because the morphological divisions of the soritines the archaiasines remain unclear. Many of the cosmopolitan Soritoidea (75%) survived the Oligocene- Miocene boundary. Very few new forms made their appearance in the Miocene of Tethys. The Early Miocene (Te5 and lower Tf1) included the complex cyclical form with subdivided chambers, Pseudotaberina (Fig. 7.4; Plate 7.4, figs 1-6 , 12), and in the latest Miocene (Tf3) its analogue Marginopora vertebralis made its first appearance. M.  vertebralis evolved from Amphisorus martini in the Serravallian at the base of Planktonic Zone N13 (see Chart 7.1) by acquiring transverse medial partitions (Lee et al., 2004). While A. martini ranges from latest Oligocene to the Serravallian, M. vertebralis and different species of Amphisorus range up into the Holocene. In the Middle Miocene of the Caribbean, the Soritoidea evolved into new forms, attaining the largest known size for member of this superfamily (Fig.  7.12). They developed a lineage resembling the Late Cretaceous meandropsinids (Ciry, 1964) by newgenrtpdf The Cenozoic Larger Benthic Foraminifera: The Neogene 567 Holo. N22 GELA. N20- Th PIAC. N21 N19 N18 N17 Tg N16 N15 N14 N13 Tf3 N12 Tf2 N11 N10 Upper N9 MiddleTf N8 1 Lower Tf1 N7 N6 N5 Upper Te (Te5) N4 P22 Lower Te (Te1-4) Figure 7.11. The main genera of the Soritoidea of the Neogene. LATE EARLY MIOCENE MIDDLE MIOCENE LATE MIOCENE PLIOCENE Pleist. Epoch OLIGOCENE CHATTIAN AQUIT. BURDIG. LANGHIAN SERRAVALLIAN TORTONIAN MESSINIAN ZANCL. Stage Planktonic Zone Peneroplis Archaias Cyclorbiculina Sorites Amphisorus Fusarchaias Marginopora "Letter Stages" 568 Evolution and Geological Significance of Larger Benthic Foraminifera Archaias af af Marginopora E Annular concentric with thickened folded margins F M. Miocene - Holocene Planispiral involute Middle Eocene to Holocene D Puteolina Cyclorbiculina C Planispiral evolute almost circular af Oligocene - Holocene Planispiral uniserial stage flaring B Peneroplis A Eocene - Holocene Late Cretaceous - Holocene Planispiral followed by uniserial stage Spiroloculina Figure 7.12. The development of the soritids from (A) planispiral spiroliniform to (B) planispiral- evolute and flaring, (C) peneropliform, planispiral-e volute, approaching- annular stage and to annular-c oncentric, with concave side views and with thickened margins, as in Marginopora vertebralis. (af = apertural face). (E) and (F) from Hottinger (2006). producing alar prolongations overgrowing each other on the lateral surface of the large discoidal shell (Hottinger, 2001). Most of the new forms did not survive the Late Miocene and they are not the direct ancestors of the species living today in the Caribbean. These porcelaneous soritids are more closely related to the Early Miocene forms of the Neotethys than to the Holocene Caribbean endemists (Hottinger, 2001). 7.3.3 The Rotaliides of the Neogene The biostratigraphically important rotaliides of the Neogene form five superfamilies: • the Nummulitoidea • the Planorbulinoidea • the Rotalioidea The Cenozoic Larger Benthic Foraminifera: The Neogene 569 • the Acervulinoidea • the Asterigerinoidea. The Nummulitoidea were major reef-f orming organisms from the Middle Paleogene to Early Neogene. Their morphology, phylogeny and palaeogeographic evolution have been recently revised by BouDagher- Fadel and Price (2014), who described for the first time new Mediterranean- derived species of Planostegina in the Early Miocene (Burdigalian) of SW Africa (Fig. 7.13). Most of the Nummulitoidea that survived of the Oligocene- Miocene boundary are extant forms, except the wholly involute Heterostegina (Vlerkina). The latter evolved in the Late Eocene from Heterostegina sensu stricto by losing the evolute coiling and acquiring chamberlets. It reached a maximum in abundance and diversity in the Late Oligocene (Banner and Hodgkinson, 1991), but diminished in the Lower Miocene (BouDagher- Fadel and Banner, 1999; BouDagher- Fadel et al., 2000), persisting through- out the Middle and Late Miocene deposits of the Far East (BouDagher- Fadel, 2002). 5 6 4 7 3 P D 8 2 1 l L Figure 7.13. Morphometric measurements in the equatorial section of megalospheric Planostegina africana BouDagher- Fadel and Price, from BouDagher- Fadel and Price (2014); P, proloculus and D, deuteroconch; Pre- planosteginid chambers (X) are 8; l, the maximum diameter of the shell in the first whorl, L, the maxi- mum diameter of the first one and subsequent half whorl. 570 Evolution and Geological Significance of Larger Benthic Foraminifera Tansinhokella also evolved from an involute ancestor in the Eocene by developing meandriform alar prolongations, which overlapped each other, and became wide- spread in Tethys and the Pacific during Late Eocene, Oligocene and Early Miocene. Spiroclypeus and Tansinhokella ranged up from the Td stage to the top of the Te stage. Classical morphology- based taxonomy divides the Nummulitidae into two subfamilies, the Nummulitinae and Heterostegininae, according to the presence or absence of secondary septa. The evolutionary importance of this morphological fea- ture was tested by Holzmann et al. (2003) by sequencing fragments of the SSU and LSU rRNA gene of many nummulitid species. According to their results, species characterized by septate chambers (Heterostegina depressa, Planostegina operculin- oides, and Cycloclypeus carpenteri) either group with species lacking septate cham- bers (Operculina ammonoides, Nummulites venosus) or branch separately. They also concluded that chamber subdivisions developed several times independently in the evolutionary history of the Nummulitidae, providing another example of parallel evolution in Foraminifera. The most primitive species of Cycloclypeus are very close morphologically to Heterostegina (see Chapter 6). Tan Sin Hok, in his monograph (1932) on Indonesian forms, documented the apparent gradual reduction of nepionic chambers in Cycloclypeus. Cycloclypeus eidae (Plate 7.5, figs 15- 19) evolved in the Late Oligocene (see Chapter 6) from C. opernoorthi, reducing the diameter of the nepionic part of the test to about one third and the number of nepionic chambers in microspheric forms from 34 to 21 (O’Herne, 1972 and van der Vlerk). It persisted to the Early Miocene where it was followed by a growth phase reduction from 21 to 18 in C. pos- teidae in Early Miocene times, and 18 to 9 in C. indo- pacificus in Early to Middle Miocene times. This group led to the younger forms which Tan Sin Hok placed within the species C. carpenteri (Plate 7.6, figs 1- 4), with 2 to 5 nepionic chambers (see Chapter 7). In the Middle Miocene of the Far East C. carpenteri, with its large embryont, is common along with stellate types (e.g. Radiocycloclypeus). However, Tan Sin Hok (1932) recognised that this progression is not gradual and he proposed a model of minor saltations with an overall smooth transition. These saltations were not found by later workers (Drooger, 1955, 1993; MacGillavry, 1962). On the other hand, Tan Sin Hok recognised two major groups in his material on the basis of ornamentation of the test. C. carpenteri has pustules, while C. guembelianus has a smooth test. These features are hard to see in thin sections and they only have ecophenotypic importance, but they have led to an interesting line of research in recent years (Laagland, 1990). European Cycloclypeus has been studied in detail by many authors (Cosijn, 1938; Matteucci and Schiavinotto, 1985; Laagland, 1990; Drooger, 1993) and the successive morphometric species C. droogeri and C. medi- terraneus were proposed by Matteucci and Schiavinotto (1985). More recently Laagland (1990) and Drooger (1993) remarked that C. eidae might have originated in the Mediterranean from C.  mediterraneus. On the other hand, the European C. eidae has a thinner test and distinct pustulate ornamentation which, according to Drooger and Roelofsen (1982) might have lived at greater depth than C. mediter- raneus. Altogether, the Cyclolypeus lineages appear to follow the same pattern of nepionic acceleration from Oligocene to Holocene. The Cenozoic Larger Benthic Foraminifera: The Neogene 571 Katacycloclypeus made its first appearance in the Middle Miocene (upper Tf1 and Tf2) of Tethys. In the upper Tf1 stage, a lineage of Cycloclypeus evolved into forms with widely separated annular inflations and a trilocular embryont (Katacycloclypeus annu- latus), which in turn gave rise in the Tf2 stage to a form with broad closely spaced annu- lar inflations and a trilocular embryont (K. martini). The co- existence of Cycloclypeus and Katacycloclypeus spp. in the Far East (BouDagher- Fadel and Lokier, 2006) dem- onstrates clearly parallel evolutionary sequences. Primitive specimens of K.  martini showing the transition between both species occur together with K. annulatus in the same sample. These specimens link the two species and demonstrate a gradual evolu- tion from K. annulatus into K. martini. The evolutionary sequences mean that these evolved taxa must occur one after the other in the same sequence if one is sampling the Indo- Pacific province. Katacycloclypeus is confined to the Tf1-T f2 stages, upper Early Miocene to Middle Miocene of the Indo- Pacific, while Cycloclypeus ranges from the Eocene to Holocene throughout all the tropics. Most of the Planorbulinoidea, which survived the Oligocene- Miocene boundary are extant forms, except Neoplanorbulinella and Planolinderina, both of which disappeared within the Early Miocene of Tethys. On the other hand, the forms which appeared in the Early Miocene were the short lived American form Heterosteginoides and the extant form Rupertina. Although many cosmopolitan forms of the Planorbulinoidea are still living, no new forms seem to have appeared in the Neogene, except the Holocene Planorbulinopsis in the Indo- Pacific. Many of the Rotalioidea became extinct during the Miocene. Only the robust forms of the Rotalioidea survived to the present day, developing spines in the Pliocene to Holocene to form the calcarinids. The spiral of the calcarinid chambers is covered by a layer of canals (with the exception of the genus Baculogypsina), which connect the chamber lumina with the ambient sea- water and, as in the nummulitids, is necessary for pseudopodia formation, excretion, attachment to the substratum, protection and prob- ably also growth and reproduction. (Röttger and Kriiger, 1990). These canal systems which characterise the families Calcarinidae and Nummulitidae originate from spaces between chambers (so- called interlocular spaces), and from spaces between the many calcite lamellae deposited successively on the test surface during ontogeny (so-c alled secondary lamellae). The canal system of calcarinids differs morphologically from that of nummulitids, as it is an enveloping canal system consisting of a layer formed by innumerable canals (Röttger and Kriiger, 1990). It covers and often conceals the spi- ral of chambers (Hottinger and Leutenegger, 1980). Morphological differences in the same species, such as Calcarina gaudichaudii (see Chapter 1), extends to the number of spines and stages of life cycles (Röttger et al., 1990). The relationship between the calcarinids and nummulitids was proposed by earlier workers such as Glaessner (1945), Pokorny (1958), Haynes (1981), and Tappan and Loeblich (1988), all of whom claim that both groups descended from a common rotaliid ancestor, and that their separation from each other and from early Rotaliidae took place in the Late Cretaceous. This was confirmed later on by Holzmann et al. (2003) who tested the phylogenetic relationships of five extant nummulitid genera by sequencing fragments of the SSU and LSU rRNA gene. According to their results, there is a close relationship between nummulitids, and the Rotaliidae and Calcarinidae. According to these authors, the Nummulitidae branch is a sister group to the Calcarinidae and Rotaliidae, and evolved in a parallel lineage. 572 Evolution and Geological Significance of Larger Benthic Foraminifera In the Early Miocene, and prior to the calcarinids, the Miogypsinidae of the Rotalioidea made their first appearance, evolving from a rotaliid ancestor (see Chapter 6) within a very short time span in the Late Oligocene. In the Burdigalian (the uppermost part of the upper Te stage), Miogypsinodella evolved from Miogypsinoides (BouDagher- Fadel et al., 2000; BouDagher- Fadel and Price, 2013) by developing splits between the lamellae in the otherwise solid lateral walls (see Chapter 6). The splits in the laminae occur throughout the thickened lateral walls in the more advanced forms (e.g. in Miogypsinodella primitiva Plate 7.11, figs 2- 4), but do not form stacks of cubiculae/ chamberlets, as seen in Miogypsina spp. (with cubiculae and quadriserial embryonts). Miogypsina evolved at the base of the upper Te stage, Aquitanian (BouDagher-F adel and Banner, 1999). While microspheric forms of Miogypsina retain the uniserial embryonic coils of their Miogypsinella ancestors, the megalospheric nepionts possess a deuteroconch larger than either the protoconch or first auxiliary chambers, and the latter form biserial whorls surrounding the proloculus. The cubiculae were arranged in oblique stacks and the biserial embryont evolved to achieve bilateral symmetry, (e.g. M. indonesiensis, Plate 7.11, fig.16). The cubiculae of Miogypsina are not in vertical stacks (Fig. 7.14), as they are in Lepidocyclina, but in obliquely developed columns. The general trend of evolution in this group is towards shorter nepionic spirals and larger embryons in successively younger species. Another descendant from Miogypsina, Miolepidocyclina has been distinguished on the more central location in the median C B A Figure 7.14. SEM images of Miogypsina borneensis Tan Sin hok from BouDagher-F adel and Price (2013) showing; (a) embryonic apparatus, and equatorial chambers; and (b, c) the internal surface of equatorial chamber wall bearing ‘eggholders’. The Cenozoic Larger Benthic Foraminifera: The Neogene 573 Figure 7.15. The phylogenetic evolution of the Mediterranean Miogypsina spp. and Miolepidocyclina spp., from BouDagher- Fadel and Price (2013). plane of the nepiont. Forms with chambers with hexagonal shapes, Lepidosemicyclina, are found in the Burdigalian of the Far East. Miogypsinids originated in the American province (Rupelian and P18, see BouDagher-F adel and Price, 2010b, 2013), becoming extinct in the Early Miocene (Burdigalian and the lower part of N8) in the American and South African provinces, in the earliest Middle Miocene (Langhian and the upper part of N8) in the Mediterranean, but their final global extinction occurred in the latest Middle Miocene (Serravallian and N13) in the Indo-P acific (Figs 15- 17, see BouDagher- Fadel and Price, 2013). The biometric method (see Chapter 6) appeared to be a useful tool in the study of the evolution of a single lineage of Miogypsina and of miogypsinid genera in general (see Chart 7.1). The “X scale” when applied to Miogypsina gave rise to the subdivision in basraensis 12.5 – gunteri – 9 – tani. At X values between 7 and 6 and γ values close to zero, “V values” become significant and the morphometric limits of the species are as follow: globulina - 45 – intermedia – 70 – cushmani – 90 – antillea (Drooger, 1993). However, as most specimens studied by scholars from random thin sections, biometric measurements on Miogypsinidae are rarely possible. In this case, combining the broad results gained by equatorial sections of the megalospheric nepiont (as published by newgenrtpdf 574 Evolution and Geological Significance of Larger Benthic Foraminifera Figure 7.16. The phylogenetic evolution of the Indo- Pacific Neorotalia, Paleomiogypsina, Miogypsinella and Miogypsinoides spp., from BouDagher-F adel and Price (2013). Mediterranean Indo-Pacific N13 Tf3 N12 Tf2 N11 Upper N10 Tf1 N9 MiddleTf1 N8 N7 Lower Tf1 N6 N5 Upper Te (Te5) N4 Te4 P22 Te3 Te2 P21 Te1 P20 Td P19 Tc P18 E p o c h S t a g e P l a n k t o n i c Z o n e t e t h y a n a N e o r o t a l i a b o n i n e n s i s P a l e o m i o g y p s i n a c y p r e a M i o g y p s i n e l l a c o m p l a n a t u s f o r m o s e n s i s M i o g y p s i n o i d e s b a n t a m e n s i s d e h a a r t i p r i m i t i v a M i o g y p s i n o d e l l a t e t h y a n a N e o r o t a l i a b o n i n e n s i s P a l e o m i o g y p s i n a b o r n e a b o r o d i n e n s i s M i o g y p s i n e l l a u b a g h s i c o m p l a n a t u s f o r m o s e n s i s M i o g y p s i n o i d e s b a n t a m e n s i s d e h a a r t i i n d i c a p r i m i t i v a M i o g y p s i n o d e l l a " L e t t e r S t a g e s " The Cenozoic Larger Benthic Foraminifera: The Neogene 575 OLIGOCENE EARLY MIOCENE MIDDLE MIOCENE AGE STAGE PLANKTONIC ZONES Neorotalia Paleomiogypsina Miogypsinella Miogypsinoides Miogypsinodella Miogypsina Americogypsina Miolepidocyclina Figure 7.17. Phylogenetic chart showing the evolutionary lineages of the American Miogypsinidae, from BouDagher-F adel and Price (2010b). Chaproniere 1984, and Van Vessem, 1978) with those obtained by vertical sections of the whole test (as followed by Cole 1957, 1963) is useful (see BouDagher- Fadel et al., 2000; BouDagher- Fadel and Price, 2013). Many new cosmopolitan Acervulinoidea joined the Paleogene survivors of this superfamily and continued to live in the present day, for example the coloured, reefal modern Homotrema and Miniacina. These extant forms appear to have arisen from the planorbulinids in the Early Miocene and their coloration (red in the case of Homotrema rubrum, Plate 7.9, Figs 4- 5) is due to the remnants of photosynthetic pigments of some type of green alga (Strathearn, 1986). Forms with more than one encrusting layer, such as Borodinia, seem to appear only in the Aquitanian of the Indo- Pacific. The acervuli- nid, Alanlordia banyakensis (Plate 7.4, figs 7-1 1), is first known from the Middle Miocene (Serravallian, Tf2), continues into the Pliocene where it strengthens the umbilical and spiral pillars with lamellar thickenings (A. niasensis, Plate 7.4, figs 9- 11). Alanlordia resembles the Paleogene Wilfordia, and from which it may have descended (Banner and Samuel, 1995). It is also a gross homeomorph of the Maastrichtian Vanderbeekia of the Middle East (see Chapter 5). In the Asterigerinoidea, Lepidocyclina sensu lato, an extinct three-l ayered Rotaliida, with a bilocular or multilocular embryonal stage, surrounded by a thickened wall and adauxiliary chambers. Microspheric tests have an early planispiral coil, while mega- lospheric tests have a globular protoconch, enclosed or followed by a larger reniform (M) (Ml) newgenrtpdf 576 Evolution and Geological Significance of Larger Benthic Foraminifera Figure 7.18. Chart showing the evolutionary lineages of lepidocyclinids throughout the bioprovinces. The Cenozoic Larger Benthic Foraminifera: The Neogene 577 Early Miocene Early Oligocene D Straight wall D P P A B Curved wall D D P P C D Figure 7.19. Thin sections of four specimens showing the two dominant types of lepidocyclinids offshore Brazil. A) Lepidocyclina (Lepidocyclina) canellei Lemoine and Douville with four equal spires, note P slightly narrower than D but slightly longer, the wall is thicker compare to B) L. (L.) yurnagunensis Cushman. C) Lepidocyclina (Nephrolepidina) bikiniensis Cole with the protoconch occupying 57% of the embryonic appa- ratus compared to 49% in D) L. (N.) braziliana BouDagher- Fadel and Price. In the latter, the arrangement of chambers is chaotic outside the periembryonic ring made of numerous short spires which are eventually replaced by the cyclic growth of hexagonal chambers. (after BouDagher- Fadel and Price, 2010a). deuteroconch (Figs 7.18-7.22). Post-embryonic chambers evolve from cyclical, arcu- ate to hexagonal in shape, usually with two or more apertures (Fig. 7.19-7.20). The lateral chambers are well differentiated from the equatorial layer and in the advanced forms they are arranged in tiers on either side of the equatorial layer (Fig. 7.23) (see BouDagher-Fadel and Price, 2010). They first evolved in the American province from an amphisteginid ancestor (see Chapter 5). Following the first explosion of new forms of L. (Lepidocyclina) in the Middle to Late Eocene, the gradual evolution of Lepidocyclina spp. seems to have been slow with the whole stratigraphic range from the Oligocene to Miocene showing little variation. From the Oligocene to the middle Miocene, many lineages of Lepidocyclina can be found in Mediterranean, with a grad- ual stratigraphic succession of species from those with primitive to those with more advanced nepionts, and from forms with thin-w alled cubiculae/ chamberlets to those with more numerous cubiculae in the swollen parts of the test (Fig. 7.18). Lepidocyclina sensu lato first appeared in the Chattian (P21b in the Mediterranean and P22 in the 578 Evolution and Geological Significance of Larger Benthic Foraminifera Indo- Pacific province). In the Mediterranean, the oldest forms, such as L. (L.) prae- marginata, which appeared in the Rupelian (see Fig. 7.18), are forms with an embryon intermediate in shape between L.  (Lepidocyclina) spp. and L. (Nephrolepidina) spp. They are similar in shape to their ancestral form in the American province, L. (L.) yur- nagunensis (Fig. 7.19), but with a slightly more curved wall separating the two embry- onic chambers. On the other hand, the L. (L.) spp. of the Indo- Pacific province had an embryonic apparatus with a straight wall dividing the two embryonic chambers, being similar to that of their ancestors in the American province but the wall surrounding the protoconch and deuteroconch is slightly thinner and the test is smaller (see BouDagher- Fadel and Price, 2010a). While the American isolepidine embryo persisted to the early Miocene, those of Tethyan lineage disappeared completely in the late Oligocene. Figure 7.20. Trend of average grade of lepidocyclinid protoconch enclosure (images not to scale). The Cenozoic Larger Benthic Foraminifera: The Neogene 579 i s i ii i A B s A i s Pa i Ps i Pa i C i s D Figure  7.21. A, C) Plan diagrams to illustrate the nepionic stages of the equatorial chambers of the megalospheric primitive nepiont A) and the advanced “trybliolepidine” in Nephrolepidina, modified from Tan (1939), Renz and Küpper (1947) and Eames et  al. (1962). B) Lepidocyclina (Nephrolepidina) aquita- niae Silvestri. D). L. (N.) angulosa Provale. Abbreviations: Protoconch (Pr); Deuteroconch (D); adauxiliary chambers (A); primary auxiliary chambers (pa), communicating directly with the nucleoconch and giving rise to the interauxiliary (i), symmetrical auxiliary (s), and protoconchal symmetrical auxiliary (pa) cham- bers. The arrows indicate the multispiral growth. The lepidocyclinid lineage (Chart 7.1) also demonstrates the evolutionary development of L. (Nephrolepidina), from L. (Lepidocyclina) at around the Late Oligocene- Miocene boundary in the Far East. Phylogenies such as this one must have occurred in other lin- eages of the Lepidocyclinidae. The L. (Nephrolepidina) lineage evolved many times. The earliest L. (N.) species in Tethys occurred in the Chattian, but they were rare (BouDagher- Fadel and Banner, 1999). Following their first appearance in the Mediterranean, there was a gradual stratigraphic evolution of species from those with the protoconch slightly enclosed by the deuteroconch (e.g., L. (N.) praetournoueri; see Fig. 7.19) to those with more advanced embryonic apparatus with the protoconch more enclosed within the embryonic apparatus (see Fig.  7.20). The European representatives of L. (Nephrolepidina) never reached the advanced phylogenetic evolutionary stages seen in the Indo- Pacific, where the protoconch became quadrate and completely enclosed by the deuteroconch (Fig. 7.21). In the latest Early Miocene (Late Burdigalian) a lineage with irregular embryonic apparatus evolved from L. (N.) transiens into more extreme forms with multiloculor embryont, Multilepidina (included in the synonyms of Nephrolepidina by Loeblich and Tappan, 1988), which survived until the end of the Serravallian (top Tf3). Species of L. (Nephrolepidina) witnessed through the Miocene an evolutionary process involv- ing the increase of the radial symmetry of the embryonic apparatus and the gradual enclosure of the protoconch by the deuteroconch (Figs 7.18, 7.21). Chart 7.1 shows an example of a lineage grading gradually towards forms with quadrate embryont, e.g. L. 580 Evolution and Geological Significance of Larger Benthic Foraminifera (N.) rutteni (Fig. 7.22) L. (N.) rutteni quadrata, which some authors call Lepidocyclina (Tribliolepidina) or just Tribliolepidina (see Fig. 7.21). The diversity of forms shown in assemblages of contemporary species, e.g. as reported by Cole (1957) from the Oligocene and Miocene of Saipan, suggests that evolution must have proceeded along several different parallel (see Plates 7.17- 7.20). It is thought that the earliest L. (Nephrolepidina) species in Europe occurred in the latest Early Miocene (Rupelian), reaching Tethys (from Iran to the Indo- Pacific) in the Chattian but they were rare. In Tf3, L. gigantea (Plate 7.8, fig. 6) became so large in Vietnam, Papua- New Guinea and East Africa that these microspheric lepidocyclinas characterize the so- called “Oyster Beds” in Vietnam. Their megalospheric partners are unknown but must have been small and undistinguished. Similar evolutionary patterns occurred in many lineages of Lepidocyclina s.l., forming taxa with no confidently known megalospheric partners (see BouDagher- Fadel and Wilson, 2000; BouDagher- Fadel and Price, 2010a). Many lineages of Lepidocyclina can be traced from the Eocene to the Early Miocene of the American province, and from the Oligocene to the Middle Miocene of Tethys, with a gradual stratigraphic succession of species from those with primitive to those with more advanced nepionts, and from forms with thin-w alled cubiculae to those with more numerous cubiculae in the swollen parts of the test. The change from biconvexity to biclavate morphology in axial section retained the development of massive pillars, which ultimately became thinner in species such as L. (N.) ferreroi (Pate 7.17, figs 15- 16). All Lepidocyclinidae disappear completely towards the end of the Early Miocene in the American province (BouDagher- Fadel and Price, 2010a) and in the southeastern Tethys in the Late Miocene (Hallock and others, 2006; BouDagher- Fadel and Price, 2013) or possibly in the Early Pliocene (Betzler, 1997) (see Fig. 7.18). Nephrolepidines evolved rapidly showing many evolutionary characters with short ranges. They reached their diversification peak during the early Miocene, often dominating whole assemblages of shallow warm carbonate platforms. The European Figure 7.22. Lepidocyclina gigantea Martin, Serravallian, Tf3, Japan, UCL coll. The Cenozoic Larger Benthic Foraminifera: The Neogene 581 representatives of Lepidocyclina (Nephrolepidina) form a well-d ocumented lineage, with three evolutionary stages: L. (N.) praemarginata, L. (N.) morgani and L. (N.) tournoueri (Fig. 7.23). The L. (Nephrolepidina) development in the Indo- Pacific was distinctly dif- ferent from the European lineage, and can be summarized as starting from L. (L.) isolep- idinoides, via L. (N.) sumatrensis, L. (N.) angulosa and L. (N.) martini to L. (N.) rutteni, based on the principle of nepionic acceleration (see Chart 7.1; Fig. 7.18). This is prob- ably an over simplification, as many more morphotypes can be distinguished, some of which may deserve species rank (BouDagher-F adel and Lord, 2000; BouDagher-F adel and Price, 2010a). As in Miogypsina s.l., the Indo- Pacific Lepidocyclinidae show an apparent reversal of nepionic acceleration (from L. (N.) sumatrensis to L. (N.) angulosa; see Chart 7.1, Figs 7.18, 7.21) which could possibly be explained by an immigration or influx of less evolved species from elsewhere in the sub-p rovince (BouDagher-F adel and Lord, 2000; BouDagher- Fadel and Price, 2010b, 2013). They could sometimes be explained by local environmental factors, or by invoking bottleneck events in which only less specialized and more primitive surviving forms provide the basis for further development (BouDagher- Fadel, 2008). However, local extinctions and re-c olonization by forms which evolved elsewhere and reached different evolutionary stages might also explain these relapses. 7.3.4 The Textulariides of the Neogene The Oligocene saw the disappearance of the shallow water Orbitolinoidea with the extinction of Dictyoconus. However, by the Neogene all of the complicated textulariides had changed their habitats completely and had moved to deeper water environments (see below). They are only studied briefly in this book as they do not colonise the same shallow warm environment as other larger benthic foraminifera, however, they do draw attention to the way larger foraminifera can adapt and change environment in order to survive. The three most important superfamilies of the Neogene textulariides are: • the Ataxophragmioidea • the Pavonitinoidea • the Textularioidea. The textulariides diversified in the Neogene and many forms carried on from the Oligocene to the present day. However, a major diversification occurred in the Miocene with new forms appearing in the Early to Middle Miocene. These forms were character- ised by inner structures consisting of alveoli conjoined through tubiform connections (Seiglie et al., 1986) and they include the Alveovalvulinidae, the Globotextularidae (see Chapter 6) the Chrysalidinidae (see Chapter 6) and the Pavonitinidae, the Textulariellidae and the Cyclamminidae (see Chapter  6). The latter are relict from the Cretaceous and continued to survive until the present day. The families Alveovalvulinidae and Liebusellidae evolved from a trochospiral agglutinated foraminifera with no inner structures. The flexible characteristic of the alveovalvulinid test contrasts, however, with the strong, thick-w alled liebusellid test, suggesting that they evolved separately (Seiglie et al., 1986). newgenrtpdf 582 Evolution and Geological Significance of Larger Benthic Foraminifera A B Figure 7.23. A) Equatorial layers of Lepidocyclina (Nephrolepidina) tournoueri Lemoine and Douvillé showing simple paired intercameral foram- ina and associated grooves; B) Eulepidina sp. showing more numerous intercameral foramina and more sharply incised grooves(modified from Eames et al., 1962). The Cenozoic Larger Benthic Foraminifera: The Neogene 583 7.4 Palaeoecology of the Neogene Larger Foraminifera Warmer global climates than those in the preceding Oligocene, and of today, prevailed during the Miocene, which lasted for 18 million years. During this time, the modern pattern of ocean circulation was established and the mixing of warmer tropical water and cold polar water was greatly reduced. This led to well- defined climatic belts that stretch from the Poles to Equator. In the Early Miocene, the larger benthic foraminifera were common throughout the warm- waters of the tropics, with many forms that were cosmopolitan and are still extant. Their latitudinal distribution was mainly controlled by temperature, thus the 15oC isotherm (for the colder months) limits their present day geographic distribution (Langer and Hottinger, 2000) and probably also did so in Miocene (Hughes, 2014). The distribution was also dependent on the light requirements of their endosymbionts and substrate (Murray, 2006). Their distribution in time followed parallel lineages showing gradual evolution, in which newly evolved taxa co-e xisted with their ancestors, but occupied different ecological niches, thus allowing palaeoenvironmental specialization into separate distinct biocenoses. In the assemblages characteristic of the Letter Stages Te to Tf1 in the Indo-P acific the miliolides were common and included forms such as the alveolinids Austrotrillina, Borelis, which avoided very shallow water and lived in sheltered habitats, such as coral rubble, in moderate energy areas (Murray, 2006; BouDagher- Fadel et al., 2000). Modern alveolinids live in a wide range of carbonate habitats, both in areas with very low hydrodynamic energy, such as deep lagoons and in forereef settings, down to depth of about 80m. This, together with the fact that alveolinids are miliolines with a toler- ance to salinity and temperature fluctuations, probably makes the group less sensitive to small sea level changes. Oligocene and Miocene alveolinids also had a wide range of habitats, but were particularly common in deeper lagoonal settings. Their response to sea level changes during the Cenozoic might therefore be similar to that of other mili- olines, were it not that extinctions during sea level falls are less likely due to the group’s wider ecological niche. Obliteration of the lagoonal habitat during sea level falls may, however, have caused reductions in population sizes and an increase in the rate of evo- lution. The appearance of Flosculinella reicheli within the Early Miocene (late Te) and of Flosculinella bontangensis in the late Early Miocene (just after the top of Te) may have been the result of smaller population size coupled with an already existing ten- dency to increase the number of rows of chamberlets. Whatever the biological advan- tage of increased complexity in chamber subdivision may have been, it is a feature of alveolinid foraminifera from the Carboniferous onward, and is a well- established evolutionary trend. It may be analogous in function to the increased complexity of the Austrotrillina wall, probably increasing the efficiency of symbiosis with algae and dia- toms. As with the fusulinines of the Permian (see Chapter 2), cyanobacterial symbionts are reported in living Marginopora (Lee et al., 1997). The sequential development of new characteristics in the lepidocyclinids and mio- gypsinids, such as the gradual development of the lateral chamberlets, may have been driven by environmental stress. Thus, the division of the solid lateral walls into the small illuminated chamberlets of Miogypsina and Lepidocyclina, were then used in 584 Evolution and Geological Significance of Larger Benthic Foraminifera favourable warm, light- filled shelf environments as nests for diatoms (see BouDagher- Fadel 2008; BouDagher- Fadel and Price, 2010a, 2013). Each of the small chamberlet would have acted not only as a small convex lens to focus sunlight, but they could also act as a greenhouse for the containment and development of symbiotic diatoms (see Fig. 7.14). The diatoms enabled these forms to acquire nutrients without pseudopodial activity in food- gathering (Röttger, 1971). Although many small rotaliide foraminifera (e.g. Rotalia, Rosalina, etc.) gather food particles to their umbilici for extrathalamous digestion, larger foraminifera with chamberlets but no umbilici (e.g. Miogypsinella and Miogypsina) must have had a different method of nutrition. In the Miogypsinidae sev- eral lineages evolved (BouDagher- Fadel et al. 2000)  that followed similar evolution- ary patterns, all related to the utilization of radiant light to enable the promotion of populations of photosymbionts. In Miogypsinella the flange of median chamberlets is relatively small (not much broader than the coil of the nepiont); in Miogypsinoides the flange grew very much broader and became heavily thickened laterally. As in the lepi- docyclinids (Fig. 7.23), the lateral walls that became very thick, would have reduced the amount of sunlight which could reach the median layered chamberlets, so they began to subdivide, as in Miogypsinodella, with the divisions becoming many layers of convex chamberlets in Miogypsina. Although diatom endosymbionts may initially enter larger foraminifera as ingested food, they establish themselves very quickly as a permanent resident population inside the cell (Richardson, 2001). This has evolutionary parallels with the division of the median chambers of Cycloclypeus (Fig. 7.6), and Heterostegina (Plate 7.15, figs 4- 6) (Banner and Hodgkinson, 1991), where the illuminated chambers are also nests for diatoms. It is probable that Miogypsinella had active, food- gathering pseudopodia, however, it is possible that Miogypsina had pseudopodia which were of little effect in food gathering and instead used its chamberlets as an arable farm. The chamberlets are not linked by apertures, instead perforations are present in the walls. These perforations would be too fine to allow the passage of diatoms, while cubicular apertures would have allowed to escape. The requirement for, and use of sunlight by Miogypsina is shown by its occurrence only in shallow water marine limestones, asso- ciated with fossil algae. The irregular shape of many species of Miogypsina suggest that their morphology is determined by the shape of the substrate on which they grew, such as the stems or leaves of seagrass (the substrate would be biodegradable and seagrass is not seen pre- served in thin sections). On such an irregular substrate surface it would be difficult for any organism to remain fully attached and yet migrate on that surface. It could be con- cluded, that sedentary miogypsinids grew to accommodate the shape of the vegetable substrate to which they adhered. Only in strong ambient sunlight, which would benefit both the miogypsinids and their vegetable substrate, could true Miogypsina flourish (BouDagher- Fadel et al. 2000). Abundant Soritoidea inhabited tropical waters of the Neogene. The extant repre- sentatives live today in the shallowest oligotrophic, tropical-s ubtropical, shallow- water habitats where the light intensity is high enough to produce photoinhibition in the symbionts of the foraminifera. They chose, therefore, dark substrates, most frequently on seagrass leaves (Hottinger, 1997, 2001). They are characterized by the posses- sion of rhodophyte, chlorophyte, or dinophyte photosymbionts (Richardson, 2001). The Cenozoic Larger Benthic Foraminifera: The Neogene 585 According to Richardson, endosymbiosis has played a major role in the evolution and diversification of the clade Soritoidea. In particular, the acquisition of dinophyte endo- symbionts appears to be a key innovation that facilitated a change in habitat from an epifaunal, free- living mode of life to one of living attached to phytal and non- phytal substrata. Banner (1971) described from New Guinea Alveolinella quoii with a small benthic foraminifera, Planorbulinopsis parasitica, embedded in its tests and completely enclosed by its whorls, that he concluded that it actually fed on the cytoplasm or uti- lised the test as a substrate. Keramosphaera’s habitat is an exception among the soritids. It lives in deep water (at about 3600m) from the South Atlantic to the Indian Ocean, in the Pacific off Japan, and in the North Sea (Loeblich and Tappan, 1988). The test morphology of the most common encrusting foraminifera, Miniacina and Homotrema, is mainly controlled by water energy. The forms with thick globu- lar tests dominate low-e nergy, more protected habitats than those of flat-e ncrusting tests. The latter typify exposed substrates under high- energy conditions (Brasier 1975; Ghose 1977; Reiss and Hottinger 1984; Fagerstrom 1987; Martindale 1992; Bosellini and Papazzoni, 2003). However, the coral encrusting, flat morphotypes seem to thrive on the corals’ lower surfaces, where less competition with other corals occurs, allow- ing the foraminifera to spread laterally, while the globular forms seem to prefer the upper surface, where the competition with algae pushes them to expand their test vertically (Bosellini and Papazzoni, 2003). According to present- day data, Miniacina and Homotrema seem to prefer cryptic microenvironments, often on the underside of dead corals (Vasseur 1974; Brasier 1975; Ghose 1977; Fagerstrom 1987; Martindale 1992), although Homotrema rubra from Bermuda has been observed on exposed reef surfaces too (Elliott et al., 1996). Miniacina is an important component of the polygenic- micritic crusts that strongly bind Porites coral branches of the Late Miocene Mediterranean reefs (Bosellini et al., 2002; Bosellini and Papazzoni, 2003). The succes- sion of encrusting foraminiferan assemblages in the Miocene is therefore interpreted as controlled mainly by light, competition with coralline algae, hydrodynamic energy, and coral growth fabric. Other encrusting foraminifera such as, Alanlordia, are confined to the carbonate- depositing, clean, middle to inner shelf, marine palaeoenvironments of the Neogene of the tropical Far East. In that area it is a potentially biostratigraphically valuable taxon. The Nummulitidae are common throughout the Cenozoic and most of them are extant. They are in fact the largest extant calcareous foraminifera, and the morphology of some species (e.g. C. carpenteri) seems to be constrained by biomechanical factors that enable them to achieve relatively gigantic sizes (Yan Song et  al., 1994). Living nummulitids are widely distributed in modern tropical and subtropical shallow- water seas and achieve their highest diversity in the subtropical and tropical West Pacific. Although they house photosymbionts, they prefer calm water conditions to highly illuminated areas, where their flat thin tests could be easily damaged by the hydrody- namic regime (Hohenegger et al., 2000). The relationship between their distribution and the environmental gradient “coenocline” has been investigated by some authors (e.g. Hohenegger, 2000). Test morphology depends on the depth gradient (Pêcheux, 1995) and test flattening and thinning of the walls promote photosynthesis in deeper euphotic zones (Hallock and Hansen, 1979; Hottinger, 1997). 586 Evolution and Geological Significance of Larger Benthic Foraminifera In the upper Tf1 stage of the Indo-P acific province a lineage of Cycloclypeus evolved into forms with widely separated annular inflations, Katacycloclypeus. The division of the median chambers of Cycloclypeus and Katacycloclypeus has evolutionary parallels with the division of the cubiculae (or “lateral chamberlets”) of Miogypsina, where the illuminated chambers are also believed to be nests for diatoms (see BouDagher-F adel et al., 2000). Each of the chambers cubiculae would act not only as a small convex lens for the focusing of sunlight, but could also act as a greenhouse for the containment and development of symbiotic diatoms. The diatoms enable these forms to be competitive in the supply of nutrients without pseudopodial activity in food- gathering (Röttger, 1971). Miocene shoals of the Philippines were dominated by Miogypsina (Hallock and Glenn, 1985, 1986). At the top of the Tf2 stage, and within the Serravallian, Katacycloclupeus became extinct together with the miogypsinids. The annuli of Katacycloclypeus were no longer needed to focus the sparse sunlight at greater water depth into preferred areas of the equatorial layer, and therefore it seems likely that the top of Tf2 coincides with a sharp sea level drop in the Middle Miocene. It could be argued that the Katacycloclypeus morphotypes had ample time to migrate down slope to maintain their preferred water depth, but depth distribution was dictated by temperature rather than water depth, and the deeper water may have become too cold due to the general cooling associated with the sea level fall. Living Operculina and Heterostegina occur over the same depth (20-1 30m), but the former is found to inhabit soft sediment between growths of the alga Halophila, while Hetrostegina is found to prefer living on hard substrates (Reiss and Hottinger, 1984). Hetrostegines, in developing partitioned chambers, were able to use their endosymbiotic algae more efficiently (Banner and Hodgkinson, 1991). Yordanova and Hohenegger (2004) demonstrated significant morphoclines in Operculina and Planoperculina spe- cies, and they used the different morphological characters for depth gradient estima- tion using regression analyses. They demonstrated that thick forms of Operculina with tight coiling live in shallow water (depth of 20 to 40m), while in the deeper parts of the euphotic zone (120m) thin forms with weakly coiled spirals predominate. Forms, such Elphidium, are suspension feeding foraminifera and they form “spiders web” between the stipes of coralline algae and utilize their pseudopodia to capture food from the water column (Langer and Lipps, 2003). The symbiont-b earing calcarinids, such as Calcarina and Baculogypsina, are pro- lific contributors to reef sediment in shallow- water habitats and their distribution appears to be related to dispersal rather than habitat restriction (Lobegeier, 2002). They are common epiphytes on both macroalgae and seagrasses (Lobegeier, 2002). The thick calcareous layer of the calcarinids protects them against the light intensity and UV- radiation in sub- littoral shallow-w aters (Röttger and Krüger, 1990). Their thick tests and special adhesion mechanisms (e.g. their spines) allow them to adapt to the shallowest regions, with high solar irradiation and vigorous water movement (Hohenegger, 2000). In modern Indo-P acific reefs, Calcarina, Baculogypsina and Amphistegina lobifera are the dominant robust larger foraminifera. The spinose Calcarina spp. and Baculogypsina populate environments with vegetation such as Jania mats, which are found immediately The Cenozoic Larger Benthic Foraminifera: The Neogene 587 behind or within the red algal ridges (Hallock and Glenn, 1986). Algal-c oated rubble leeward of the algal crest is often populated by Amphistegina, Calcarina and soritids, as well as abundant smaller rotaliines, peneroplids, and miliolines. Coarser, higher- energy sand flats are often devoid of living foraminifera, though the sand itself may be predominantly composed of robust foraminiferal tests. However, no living symbiont- bearing forms are found below 110m (Hallock, 1984). In the Gulf of Aqaba, these forms and Heterocyclina tuberculata are the deepest dwelling algal symbiont-b earing foraminifera, with depth limits of almost 130m (Reiss and Hottinger, 1984). A relatively large number of agglutinated foraminifera with alveolar walls, such as Lituola (see Chapter 5), which made its first appearance in the Triassic, continued to exist in the Neogene up to present day. Attached foraminifera, such as Haddonia (see Chapter 6), which made its first appearance in the Early Paleogene (Danian), is common at the present day but is confined to the fringing reef assemblages (Langer and Lipps, 2003). On the other hand, the Lower to Middle Miocene saw the diversification of the agglutinated pavonitinids, which had elongate triserial to biserial tests with complicated inner structures. Unlike the rest of the larger benthic foraminiferal assemblages, these assemblages occupy chiefly bathyal to abyssal sediments of tropical waters and consti- tute 40% of the agglutinated taxa in Neogene tropical waters (Seiglie et al., 1986). Neogene carbonate lithofacies represent a variety of palaeoenvironments rang- ing from shallow reef to relatively deep, open marine conditions. Individual out- crops often demonstrate a transition between these environments, thought to relate to changes in relative sea level, and levels of carbonate production and clastic input, which depending on location includes occasional volcaniclastic grains. Reef skeletal rich mound-shaped buildups are referred to as bioherms, while carbonate layered/ tabular build-ups are instead known as biostromes. The presence of well- preserved larger benthic foraminifera within the coralgal framework buildups of the Neogene carbonates suggests that deposition occurred entirely within the photic zone. Among the many different morphologies of carbonate platform, the most widely recognised are carbonate ramps, which are gently sloping platforms, and rimmed shelves, which are flat-topped platforms bordered by rims/barriers made of reef corals. Shallow seas developed patch reefs, in which the areas attacked by oceanic swell may be called “forereef” (Fig.  7.24). The energy distribution produces characteristic carbonate lithofacies and allows distinctive assemblages of foraminifera to flourish: • The reef itself, being shallowest, produced characteristic bioherms and calcirudites of corals and algae, cemented by sparitic matrix as the lime muds were flushed away by wave action. • In the forereef shelves, proximal biostromes of corals and algae were cemented by sparite and micrite carbonates. Distally from the reefs, coral debris diminished and micrite increased, cementing the calcarenitic biogenic debris. Below fair weather wave- base, micrite would accumulate, which contained scattered larger foraminifera swept in from the reef shelves. The proximal forereef shelf contained faunas dominated by Planorbulinella spp., cycloclypids, lepidocyclinids, operculinids and heterosteginids. • The shallow water carbonate facies are surrounded distally by deeper water sediment containing abundant planktonic foraminifera, which may constitute up to 35% vol- ume. The planktonic foraminifera, abundant in the micritic matrix, had flourished in the water column offshore and reflect relatively deeper water, and low energy. 588 Evolution and Geological Significance of Larger Benthic Foraminifera Lepidocyclina Planorbulinella Katacycloclypeus Amphistegina Marginopora Operculina Cycloclypeus Heterostegina Pseudotaberina Austrotrillina Figure 7.24. The facies range of the dominant Neogene foraminiferal taxa in a Tethyan carbonate shelf. Integrated reef/ramp model for Neogene carbonates. The ramp model is indicated by the blue dotted line. In the case of gently sloping ramp, the outer ramp lithofacies are made of mudstones and wackestones, while in the middle ramp mudstone with carbonate nodules would develop. 7.5 Palaeogeographic Distribution of the Neogene Larger Foraminifera 7.5.1 General characteristics of the distribution of Neogene larger foraminifera The end of the Oligocene witnessed very little extinction, worldwide just 13% of the larger benthic foraminifera disappeared at that time. As explained in Chapter  6, this stratigraphic boundary is probably mainly associated with plate tectonic events, and/ or gradual changes in climate (Berggren and Prothero, 1992). During the Early Miocene, modern patterns of atmospheric and ocean circulation began to be established. The proto- Mediterranean was still open, but was narrower than before (Fig. 7.25). The Early and Middle Miocene was a time of relatively warm average global temperatures (Fig. 7.26). However from the Middle Miocene, the cold phase became established as the isolation of Antarctica from Australia and South America became more pronounced, and meant the establishment of the circum- polar ocean circulation, which significantly reduced the mixing of warmer tropical water and cold polar waters, and further led to the buildup of the Antarctic ice cap. This is indicated by oxygen isotopic data obtained from deep- sea benthic foraminifera (Macleod, 2013, Fig. 7.27). Additionally, the Tibetan plat- form uplift (which began around the Eocene- Oligocene boundary, ~34 Ma) continued, and the Red Sea rifting accelerated (Aitchison et al., 2007; Thomas et al., 2006). However, the tectonic changes related to the widening of the Southern Ocean passages and the closing of the Tethys seaway, influenced circulation patterns that caused a flow The Cenozoic Larger Benthic Foraminifera: The Neogene 589 Figure 7.25. Palaeogeographic map showing the proto- Mediterranean still open but narrower than during the Early Miocene (by R. Blakey, http:// jan.ucc.nau.edu/~ rcb7/ paleogeographic.html). reversal through the Panama Seaway between the Oligocene and Miocene (Jakobsson, et al., 2007). The global thermohaline circulation in Oligocene and Miocene models are significantly different from the present day pattern. In particular, in the Oligocene the salinity contrast between the Atlantic and Pacific oceans was reduced because of water mass exchange through the low-l atitude connections between the two oceans (von der Heydt and Dijkstra, 2006). Although geochemical proxies suggest that the Drake Passage between South America and Antarctica was open earlier, in the Late Oligocene (32.8 Ma), to intermediate and deep water circulation (Lawver et al. 1992; Latimer and Filippelli, 2002), the continued opening of this gateway eventually led to the thermal isolation of Antarctica in the Early Miocene (Fig. 7.28) and the creation of the clockwise, strong Antarctic circumpolar current (Smith and Pickering, 2003). These tectonic changes during the Early Miocene influenced the biotic distributions within the oceans, in particular that of the shallow larger benthic foraminifera. These organisms, which contribute most to the Neogene biogenic carbonate sediments, have a distinct global pattern of provincialism. The four main bioprovinces known today are those of the Americas, South Africa, Western Tethys (or the Mediterranean), and the Indo- Pacific (e.g. Adams, 1967; Rosen, 1984; BouDagher- Fadel and Price, 2010a, 2013, 2014, 2017). The distribution of the larger benthic foraminifera throughout the Neogene 590 Evolution and Geological Significance of Larger Benthic Foraminifera 4 80 2 60 0 40 -2 20 -4 0 -6 10 15 20 25 30 35 -20 -8 S L B A C R -40 -10 Age (Ma) Figure  7.26. Variation in sea level and temperature during the Oligocene and Mid- Miocene based on Zachos et al. (2001), Miller et al. (2011) from BouDagher- Fadel and Price (2013). Figure 7.27. Cenozoic oxygen isotopic data in relation to global temperatures and their implications on physiographic events (from Macleod, 2013). is associated either with the Indo-P acific, or the Indo-P acific and Mediterranean (“Tethyan”), or the Mediterranean and West African, or the American provinces. There are also many cosmopolitan forms (Fig. 7.29). Although the extinctions (Fig 7.30) throughout the Neogene were not as pronounced as in other periods, each of the bioprovinces had its own patterns of evolutionary Sea Level change (m) Temp Anomaly (degree C)y The Cenozoic Larger Benthic Foraminifera: The Neogene 591 Figure 7.28. Palaeogeographic and tectonic reconstruction of the Early Miocene (by R. Blakey http://j an. ucc.nau.edu/ ~rcb7/ paleogeographic.html). lineages and extinctions, which at times were also punctuated by migration events from other provinces, examples of which are discussed in greater detail at the end of this section. In the Aquitanian, with the Tethyan seaway between the proto-M editerranean and the proto- Indian Ocean still open, 56% of the larger benthic foraminifera were cos- mopolitan, of which 41% of the new Miocene forms appeared globally (Fig.  7.31). Of the new forms in the Aquitanian, 35% were encrusting foraminifera, the acervu- linids, which in modern reefs are cosmopolitan, living at mid- depths in semi-c ryptic (gloomy), warm- water environments, encrusting corals (Dullo et al., 1990; Martindale 1992; Bosellini and Papazzoni, 2003). One of the main changes during this period was the appearance in Tethys of what were originally American lineages, such as those of the miogypsinids and lepidocyclinids (see Chart 7.1). Such lineages became very suc- cessful and evolved many times in the Indo-P acific (see the discussion at the end of Neogene Genera 80 70 60 50 40 30 20 Total Cosmopolitan 10 American Indo-Pacific Tethyan 0 Mediterranian Figure 7.29. The number of Neogene larger foraminifera genera plotted by province. Neogene Extinctions 20 18 16 14 12 10 8 6 Total 4 Cosmopolitan American 2 Indo-Pacific Tethyan 0 Mediterranian Figure  7.30 The number of extinctions of Neogene larger benthic foraminifera genera at the end of each stage. Aquitanian Aquitanian Burdigalian Burdigalian Langhian Langhian Serravallian Serravallian Tortonian Tortonian Messinian Messinian Pliocene Pliocene Pleistocene Pleistocene Holocene Holocene The Cenozoic Larger Benthic Foraminifera: The Neogene 593 Neogene Appearances 18 16 14 12 10 8 6 4 Total Cosmopolitan American 2 Indo-Pacific Tethyan 0 Mediterranian Figure 7.31. The appearances of new genera at the start of each Neogene stage boundary. this section) and played an important part in defining the biostratigraphy of the car- bonates of South East Asia, with Miogyspsina, Lepidocyclina, Eulepidina, Spiroclypeus appearing worldwide, along the tropical belt reaching as far south as New Zealand. Some forms, such as Cycloclypeus, originated in the Mediterranean, migrating raipdly towards South East Asia, but never appearing in America. Towards the end of the Aquitanian, global extinctions were low (Fig. 7.30), and only 8% of the Aquitanian genera became extinct, with the highest percentage in the Indo- Pacific (43%). In the Burdigalian much of the diversification of Miocene larger benthic foramin- ifera in Tethys is at the species level. Some new rotaliid genera made their first appear- ance in Tethys, with only a few occurring in the Indo-P acific province. The acervulinids and the planorbulinids were still common globally. The soritids became diverse, popu- lating warm seagrass environments in all provinces, while the nummulitids thrived in the reef and forereef together with the lepidocyclinids. However, towards the end of the Burdigalian, 29% of the larger foraminifera became globally extinct (Fig. 7.30). The already established cosmopolitan forms, however, were not significantly affected by this event (with only 5% becoming extinct). The larger benthic foraminifera that were most affected were the provincial genera, mainly those of the American province (with 50% of them becoming extinct) and the provincial Tethyan foraminifera (with 40% extinction). Two main groups of foraminifera that became extinct in the Americas, the lepidocyclinids and miogypsinids, were living on reefs in high energy environments. Aquitanian Burdigalian Langhian Serravallian Tortonian Messinian Pliocene Pleistocene Holocene 594 Evolution and Geological Significance of Larger Benthic Foraminifera On the other hand, the Soritoidea were not affected by this extinction, most probably because they mainly lived in sheltered habitats, and they had the opportunity to change their mode of life. They changed from living attached to phytal and non- phytal sub- strata to an epifaunal, free- living mode of life. They also changed their endosymbionts (see above), which enabled them to swap their attached habitat to become free living in slightly deeper oligotrophic environments when reefal conditions became difficult and lack of direct sunlight hampered their growth. The regional extinction of the larger benthic foraminifera in the Americas also coincided with the regional extinction of about half the Caribbean hermatypic corals, which died out during the latest Oligocene through Middle Miocene, about 24-1 6 Ma (Edinger and Risk, 1994). As was the case for the lepidocyclinids and miogypsinids, the majority of the corals that died out earlier in the Caribbean, were found in the Late Miocene of the Indo-P acific. However, unlike the corals, which are still extant in the Indo- Pacific, the lepidocyclinids and miogypsinids died out globally at the end of the Miocene. These regional extinction, confined to the Caribbean and western Atlantic, might be related to a contemporaneous eruption of the Columbia River basalts (Fig 7.32), which was at its most intense between 17 and 15 Ma (Courtillot and Renne, 2003). Figure  7.32. The Neogene world, showing the location of the Columbia River Flood Basalts and the Eltanin crater. The Cenozoic Larger Benthic Foraminifera: The Neogene 595 The CO2 emissions from this flood basalts could have triggered a cooling or acidifica- tion that may have affected the faunas on both patch reefs and shelf edge reefs. Except for a small number of deep water agglutinated foraminifera appearing in the Caribbean, very few new forms appeared globally in the Langhian (Fig.  7.31). Provincialism was rare on a genus level, with 58% of the larger benthic foraminifera being cosmopolitan and 27% belonging to Tethys. However, there was diversification on a species level in Tethys, and shallow water carbonates were teaming with foramin- ifera. The fusiform milioline Flosculinella bontangensis graded into Alveolinella forms at about the same time as the disappearance of Austrotrillina, occurring alongside Alveolinella praequoyi, within the Serravallian in the Indo- Pacific province. Also in the Serravallian, Cycloclypeus, which seemed to disappear in the earliest Miocene from the Mediterranean, seems to be common in the Indo-P acific and by the Serravallian it had developed distinct inflations along the test, as in Katacycloclypeus. In the Serravallian, plate collisions sealed off the eastern proto-M editerranean and the connections between the proto-M editerranean and the proto- Indian Ocean closed (see Fig 7.33) (see Henderson et al., 2008, 2010; Najman et al., 2008, 2010a, b). Evaporitic deposits formed in the Red Sea, the Mesopotamian basin and along the northern margin of Africa in the Sirt basin (Gvirtzman and Buchbinder, 1978; Rögl and Steininger, 1983; Jolivet et al., 2006). Many larger foraminifera became extinct in Tethys, where 60% of the global extinctions happened. Most noticeably the immigrants from the Americas, the lepidocyclinids and miogypsinids, after thriving successfully in the shallow high energy warm waters, were completely wiped out by the end of the Serravallian. Another casu- alty of the closure is Cycloclypeus, which seems to have disappeared from the proto- Mediterranean after the closure of the sea link, but continued to thrive up to the present day in the deep photic zone of the Indo- Pacific and Australia. In the Middle to Late Miocene the partial emergence of the Isthmus of Panama (~8.3– 7.9 Ma) disrupted the Atlantic to Pacific flow, closing the intermediate depth water connection (Roth et al., 2000). During the pre-c losure interval (7.6–4 .2 Ma), the Caribbean was under mesotrophic conditions, with little ventilation of bottom waters and low current velocities (Jain and Collins, 2007). This event was accompanied by the development of more provincialism, producing Caribbean endemists. Deep water textulariides, with inner complex structure, became common; they included 70% of the agglutinated larger foraminifera with inner structures found in the Caribbean. These foraminifera had structures, such as alveoles, and used to occupy shallow tropical waters of Tethys in the Early Jurassic to Cenomanian, with a brief resurgence of one of these groups (Coskinolina, Lituonella and Dictyoconus) in shallow tropical environments of the Early and Middle Eocene. In the Early Miocene, and in response to environmental stress in the Caribbean, they moved their habitats to deeper environments where they occurred in abyssal muds, poor in oxygen (Seiglie et al., 1986). In the Middle Miocene, they were mainly found in Central America (75%) with only a few occurring in West Africa and Western Mediterranean (12%), and only one genus occurring in both provinces (Liebusella). In the Tortonian, more provincialism occurred in the Americas, as new soritids were limited to the Caribbean. The Soritoidea evolved into new forms and exhibited the largest sizes in their history. These new forms are morphologically related to the Early Miocene Tethyan forms, as during or prior to the Middle Miocene, and Hottinger 596 Evolution and Geological Significance of Larger Benthic Foraminifera Figure  7.33. A map showing a permanent land barrier separating the proto- Mediterranean and proto- Indian Ocean regions developed in Middle Miocene times (by R. Blakey, http:// jan.ucc.nau.edu/ ~rcb7/ paleo- geographic.html). (2001) has suggested that the Tethyan soritids might have migrated to the Caribbean to replace the Eocene endemists However, given that historical migrations of LBF have always been from the Americas to Tethys (see below), this hypothesis may need further investigation. On the other hand, those in the Indo- Pacific were completely dominated by two soritids, Amphisorus, which appeared in the Late Oligocene and the extant new- comer, Marginopora, in the Middle Miocene. By the Late Messinian (5.65 Ma) most of the larger benthic foraminifera in the Mediterranean had become extinct (Fig.  7.30), as the Mediterranean experienced a salinity crisis and became partially desiccated. It has been suggested that this was caused by a tectonic event, with no specific climatic change being responsible for the evaporation (Jolivet et  al., 2006). In the Late Miocene period, the Gibraltar Straits The Cenozoic Larger Benthic Foraminifera: The Neogene 597 closed, and the Mediterranean evaporated forming a deep dry basin with a bottom at some places 2 to 3 miles (3.2 to 4.9 km) below the global sea- level (Clauzon et al., 1996). The end of this crisis corresponds to the breaching of the Gibraltar Straits that resulted in an extensive and rapid flooding of the Mediterranean basin (Loget et al., 2003). During this period, the Mediterranean larger benthic foraminifera were rare and localised and by the end of the Messinian 75% of them had become extinct. The foraminifera endemic to the Indo- Pacific were not very affected, however, as only 25% of them became extinct at this time (Fig. 7.30). The pre- closure interval of the Central American seaway (7.6–4 .2 Ma) saw enhanced seasonal input of phytodetritus with even more reduced ventilation, and enhanced dis- solution between 6.8 and 4.8 Ma (Jain and Collins, 2007). This lead to mesotrophic conditions and poor ventilation, abundance of the cool water and a continued drop in sea level that had begun in the Middle Miocene (Roth et al., 2000). These events might have caused the extinction of 75% of the Soritoidea, which were up to then endemic in the region. According to Hottinger (2001), these forms were more closely related to those extant in the Indo- Pacific province than those of the present day foraminifera in the Caribbean. In this case, they might not have become extinct but most probably migrated to the Indo- Pacific province to become the ancestral forms of modern Indo- Pacific soritids. At the onset of the Pliocene, new larger foraminifera were conspicuously absent from all bioprovinces, except in the Indo- Pacific, where a new lineage, restricted to the western and central tropical Pacific, emerged. These were the calcarinids, with a prefer- ence for shallow water, high- energy environments. They were particularly abundant on coral reef flats and widespread throughout the Great Barrier Reef Province of north- eastern Australia (Lobegeier, 2002). Their global distribution, according to Lobegeier (2002), seems to be limited by dispersal by oceanic currents rather than by habitat restriction. They are prevented from reaching Hawaii, and other islands lying to the east of 170o, by the westward flowing North and South equatorial currents and from the Indo- Malay region to the west or to the south by the Equatorial Countercurrent in the Indian Ocean. In the Caribbean, the total closure of the Central American seaway at 4.2 Ma caused the formation of the Isthmus of Panama, and allowed direct land- to-l and connection between North and South America. This closure blocked the Atlantic- Pacific water interchange, changing profoundly the ocean circulation, isolating the Arctic Ocean and initiating the northern polar ice (Jain and Collins, 2007). The closure of the seaway in the Caribbean produced a cascade of environmental con- sequences, including the reorganization of circulation in the Gulf of Mexico and much of the Caribbean, leading to reduced upwelling and palaeo- productivity (Allmon, 2001). In response, to the changes in ocean circulation and the decrease in the plankton in the Caribbean Sea, fast-g rowing oysters became extinct (Kirby and Jackson, 2004). Larger benthic foraminifera living in shallow- marine environments also reflected these changes by reorganizing from a suspension- feeder- dominated community to a more carbonate-r ich, phototrophic- based community. The increas- ing divergence of milioline and textulariine groups between the Atlantic and Pacific indicates differing habitats between shallow sea grass communities and deep water benthics. 598 Evolution and Geological Significance of Larger Benthic Foraminifera Extinctions at the end of the Pliocene were largely restricted to the Indo- Pacific. Some of these extinctions could have been triggered by the late Pliocene (2.15Myr) Eltanin asteroid impact in the Southern Ocean (see Fig 7.32). This is the only known deep ocean basin asteroid impact, and as such is suggested to have had major regional consequences (Gersonde et al., 1997). However, the late Neogene was also a time of exceptionally strong global cooling and oceanographic change (Zachos et al., 2001). Global cooling after ~3 Ma had for the first time a direct effect on trop- ical sea surface temperatures, resulting in high-a mplitude fluctuations in global ice volume and sea levels (Fedorov et al., 2006; Johnson et al., 2007). Other than the extinctions that may have also been triggered by climate change in the Indo- Pacific, the only other effect that this cooling seemed to have had on the tropical larger ben- thic foraminifera was a suppression of speciation and diversification (Fig.  7.31). Long ranging species continued to thrive and adapted themselves to cooler waters, and the only bioprovince which saw a small percentage of new species (Fig. 7.31) was that of the Indo- Pacific, where empty niches from the earlier extinctions became filled with new larger foraminifera. Indeed, 40% of the endemics in the Indo-P acific were new Pleistocene forms. It has been suggested that at the end of the Pleistocene (at 12.9 ka), an extra- terrestrial impact event over northern North America, caused abrupt environ- mental changes that destabilized the Laurentide Ice Sheet, triggered cooling (the Younger Dryas), and caused large- scale mammal extinctions and the end of the human Clovis Culture (Firestone et al., 2007). However, at this time the climate was already changing rapidly, with warming and changing patterns of rainfall, which in themselves could have triggered the terrestrial extinctions. However, whatever the cause of these changes they did not seem to affect the ecology of the marine tropical belt, as extinctions of larger benthic foraminifera were rare to non- existent at this time. In the Holocene, new forms of larger benthic foraminifera appeared through the tropical realm, with 70% of the foraminifera in the Indo-P acific realm being new, compared with 30% in the American bioprovince. Only 6% of the cosmopolitan fora- minifera are new appearances, with half of them being deeper water foraminifera (see Charts 7.2 and 7.3). Present day larger foraminifera play analogous roles in the eco- system, where the tropical belt is divided into two parts, different assemblages colo- nise different environments. In particular, the Indo- Pacific larger soritids, (Amphisorus and Marginopora) are substituted in the Caribbean by Archaias and Cyclorbiculina as porcelaneous discoidal epiphytes on tropical seagrasses (Langer and Hottinger, 2000). Modern tropical western Pacific sandy shoals and beaches can be nearly pure con- centrations of Calcarina, Baculogypsina or Amphistegina tests. Calcarinids can also range into the central Pacific and westward through the Indian ocean (Todd 1960; Hallock 1984; Röttger and Krüger, 1990). Analogous sediments in the Caribbean are characterized by robust peneroplids and soritids, and by thick- shelled miliolines (Brasier, 1975). Modern Indo-P acific faunas from deeper foreslopes are characterized by Heterostegina, Cycloclypeus, Operculina, and the flatter species of Amphistegina; Heterocyclina replaces Cycloclypeus and Calcarina in these deeper environments of the Red Sea- East African faunal province (Reiss and Hottinger, 1984, Renema, 2007). The Caribbean lacks deep- euphotic assemblage forms. Shallower foreslopes are dominated The Cenozoic Larger Benthic Foraminifera: The Neogene 599 by Amphistegina lessonii in the Indo-P acific and by A.  gibbosa in the Caribbean. Alveolinids occur on the shelf areas of both regions. Alveolinella in the Indo- Pacific and Borelis in both regions dominate quieter environments. In the Mediterranean area, Planorbulinella has been reported from the Oligocene to the end of the Miocene (Messinian), when the lineage ended (Freudenthal, 1969; Drooger, 1993). It has also been reported from the Miocene of Trinidad (Cushman and Jarvis, 1930). Planorbulinella is widespread in the Indo-W est Pacifc Ocean and at least two species occur from as far afield as the Red Sea and Hawaii (Renema, 2005). 7.5.2 Provincialism and migration of some Neogene larger benthic foraminifera It appears that the migration of a number of larger benthic foraminifera out of the Caribbean and into the Tethyan realm, and then onward to the Indo- Pacific and South Africa is an important, repeated characteristic of the Cenozoic. The migrations appear to correlate with global eustatic sea level falls (see BouDagher- Fadel and Price, 2010a, 2013). From Middle Eocene to Early Oligocene, American larger foraminifera were mainly endemics. By the earliest Oligocene, they had crossed the ocean perhaps being carried by currents on debris, or as planktonic gametes or zygotes, across the then narrow Atlantic from America to West Africa. Migration depends on mode of life and only those forms which could settle on floatable algae or terrestrial or volcanic debris, or that had long-l ived planktonic gametes or zygotes, could migrate. Migration stopped when sea level rose in the Rupelian. Repeated migration and geographical isolation drove the development of provincialism which became pronounced and char- acteristic of the Neogene assemblages. However, this provincialism can be highly com- plex, and indeed dynamic. This Chapter ends with a discussion of specific and recently studied examples of the complexity of bio- provincialism shown by the evolutionary trends and regional occurrence of the families Lepidocyclinidae (BouDagher- Fadel and Price, 2010a) and Miogypsinidae (BouDagher- Fadel and Price, 2013), which began in the Paleogene but culminated in the Neogene. The migration of lepidocyclinids and miogypsinids happened via three stages (Fig. 7.34). The larger foraminifera were carried first by currents on debris across the Atlantic from North America to North Africa and the Western Mediterranean in the Early Oligocene. Migration then must have occurred within the Mediterranean and along the Arabian coast by dispersal of foraminifera by algal rafting, planktonic gametes or zygotes, and long- shore drift, together with occasional migration from the Western Tethys on the Canary Current southwards towards tropical Africa, followed by the development of local stocks. They then spread rapidly in the Early Miocene, carried by ocean current from Arabia to the Indo- Pacific, or from Western Africa to South Africa. The migration of the lepidocyclinids out of the Caribbean and into Western Tethyan is an important event in the development of Indo- Pacific faunas, and appears to have occurred immediately after (or during) the Early to Middle Oligocene eusta- tic sea level falls (see Fig.  7.26). The oldest Lepidocyclinidae evolved in Central America (see Chapter 6) and are differentiated into two main lineages, which can be 600 Evolution and Geological Significance of Larger Benthic Foraminifera Figure  7.34. The migration of miogypsinids during Early Oligocene, shown by black arrows, from the Americas (1), to the Western Tethys (Mediterranean) (2), and on to the Indo-P acific (3), or to South Africa (4) (see BouDagher-F adel and Price, 2013). called Eulepidina and Lepidocyclina sensu lato, which includes L. (Lepidocyclina) and L. (Nephrolepidina), but from there they rapidly spread to Europe in the Oligocene. The established view is that the arrival of Lepidocyclina in Asia was an early to mid- Oligocene event, with some overlap of these forms with Nummulites (to give the Td faunas). During the middle Eocene to early Oligocene a series of sea level falls (Katz et  al., 2008)  enabled the migration of some species, such as Helicolepidina spiralis, Lepidocyclina (L.) pustulosa and L. (L.) rdouvillei, between the American province and the West African province (Brun and others, 1982; Ly and Anglada, 1991; Mello e Sousa and others, 2003; BouDagher-F adel, 2008). However, after the last major regres- sion in the early Oligocene, at 33.5 Ma, the rising sea level isolated the West African shelf from the American province, preventing further exchange between the Americans and Western African margins. Finally, they diffused in the late Rupelian, carried by ocean current from Arabia to the Indo- Pacific, where the equatorial conditions and abundant ecological niches favored rapid diversification and the development of a rich and complex range of species. Early Rupelian Tethyan lepidocyclinids evolved into a Lepidocyclina (N.) praemarginata and L. (Lepidocyclina) spp. assemblage (see The Cenozoic Larger Benthic Foraminifera: The Neogene 601 Fig. 7.35). These species, which imitated in their evolution the American Oligocene assemblages of L. (L.) yurnagunensis and L. (N.) braziliana (Plate 7.18, fig. 10), were first described by De Muller (1975) from the late Rupelian to early Chattian of Corfu. Since then, L. (N.) praemarginata was reported from the upper Rupelian beds in north- ern Spain and Chattian of Cyprus. While in the American province, the early Miocene assemblages were dominated by L. (L.) canellei. The American lepidocyclinids died out completely at the end of the Middle Miocene. In Tethys, the lepidocyclinids migrated eastward from the proto-M editerranean in the early Rupelian and reached the Indo-P acific towards the end of the Rupelian. L. (Lepidocyclina) spp. became extinct at the end of the Oligocene in the Mediterranean province. On the other hand, L. (Nephrolepidina) continued to thrive along the Mediterranean shelf, evolving different lineages into the Early Miocene (see BouDagher- Fadel and Price, 2010a). The earliest L. (Nephrolepidina) species in Tethys occurred in the late Chattian, and they were almost certainly the direct descendants of L. (Lepidocyclina) and L. (Nephrolepidina) of the Eocene and Oligocene of West Africa and America. During the early Miocene, the Tethyan seaway between the proto- Mediterranean and the proto-I ndian Ocean became narrower (BouDagher- Fadel, 2008), which restricted further migration between the two provinces. However, in the Indo-P acific the descendants of the Mediterranean lepidocyclinids continued to thrive successfully in the shallow high energy warm waters, evolving many independent par- allel lineages. The spatial and temporal separation of these lepidocyclinids from their ancestral Mediterranean forms precludes them from being the same species, but rather they are examples of parallel evolution, which occurred as the two bioprovinces devel- oped separately. The development of the miogypsinids is analogous to that of the lepidocycli- nids, and exploited the same migratory pathway. The miogypsinids were initially provincial, being found only in the Americans and survived there into the Early Miocene. BouDagher- Fadel and Price (2010b) have demonstrated that they origi- nated in the Americas from Neorotalia in the Early Oligocene (Rupelian and P18) (see Chapter 6). During the Early Oligocene a series of sea- level regressions (Katz et  al., 2008; Miller et  al., 2011)  reduced the effective width of the early Atlantic Ocean sufficiently to facilitate transoceanic migration of Neorotalia from the American province to the North African coast and into the Mediterranean. During this time, Mediterranean shallow water niches were still occupied by the Paleogene Nummulites. However, towards the end of the Early Oligocene (around 31– 29 Ma), the environmental stresses, perhaps associated with cooling and the large flood basalt event in Ethiopia and Yemen (see Chapter 6) contributed to the disappear- ance of the last Mediterranean Nummulites which provided an opportunity for new phylogenetic lineages of miogypsinids to fill the warm reefs of the Mediterranean (see BouDagher- Fadel and Price, 2013). The oldest representatives of the genus Miogypsina, appeared in the Oligocene in central America, but just before the Oligocene- Miocene boundary in Europe, where its evolutionary appearance occurs at the level of Miogysinoides formosensis in the latest Oligocene, but in the Indo-P acific province, the oldest Miogypsina gunteri is found with Miogypsinoides bantamensis at the start of the Miocene. M. gunteri shows a decrease in the size of the proloculus, which is one of several relapses (Drooger, newgenrtpdf 602 Evolution and Geological Significance of Larger Benthic Foraminifera Figure 7.35. Evolution of Lepidocyclina sensu lato in space and time in the Cenozoic. American province Tethys (Venezuella = V, Brazil = B) Lepidocyclina spp. L.(Lepidocyclina) L. (Nephrolepidina) B L. (Nephrolepidina) B V Lepidocyclina spp. L. (Lepidocyclina) O l i g o c e n e E a r l y M i o c e n e M i d d l e - L a t e E o c e n e M i d d l e M i o c e n e The Cenozoic Larger Benthic Foraminifera: The Neogene 603 1993) in the general principle of nepionic acceleration. In all three bioprovinces, there are later developments within the group of Miogypsinidae with lateral chamberlets (Miogypsina s.l.), towards forms in which the embryon tends to shift towards the cen- tre of the test. These forms are distinctly different in each of the bioprovinces:  the subgenera Helicosteginoides and Miogypsinita in Central America, Miolepidocyclina in Central America and the Western Tethys, and Lepidosemicyclina in the Indo-P acific. These forms appear at different times, but are each sufficiently different, and have their origins so close to contemporaneous Miogypsina, that migration does not have to be invoked As the morphologies of American and Mediterranean miogypsinids are seen to be crucially different, it follows that their evolutionary development was independent but closely parallel, as after the last major regression in the Early Chattian, the rising sea level and the continuing oceanic rifting effectively isolated the Mediterranean–W est African shelf from the American province (around 28 Ma), ending any flow of miogy- psinids from America to the Mediterranean. During the Late Oligocene and Early Miocene, successive forms of miogypsinid continued their migration eastward through the open seaway from the Mediterranean into the Indo-P acific, where they typically arrived a million years or so after their first Mediterranean appearance (Fig. 7.34). Once in the tropical setting of the Indo- Pacific, with its diverse paleogeography, the migrants gave rise to a richer diversity of local species than seen in the Mediterranean. However, in the late Burdigalian (around 17 Ma) the eastward migration between the Mediterranean and the Indo- Pacific was interrupted as tectonic processes narrowed and closed the seaway between the Mediterranean and Indian ocean (Von Rogl 1998; see Fig. 7.28). Prior to this first closure of the Eastern Tethys seaway in the Burdigalian (17 Ma), the miogyp- sinids had thrived in the warm climates of the Mediterranean, reaching their peak diversity, with a maximum number of species, in the Early Burdigalian, however, with the first closure of the Tethys seaway in the late Burdigalian, the miogypsinids became isolated, with limited ecological diversity, and only very few species survived the Burdigalian– Langhian boundary (15.9 Ma) (BouDagher- Fadel and Price, 2013). The first closure of the Tethyan sea was short-l ived, however, as a major global trans- gression in the Early Langhian (15.5– 16.5 Ma) flooded the Mediterranean from the Indo- Pacific. This transgression led to the extinction of the remaining Mediterranean miogypsinids within the Early Langhian, as they were replaced by algal– coral patch reefs, tropical mollusc fauna and flat forms of larger benthic foraminifera, such as Amphistegina (BouDagher- Fadel and Clark, 2006; BouDagher- Fadel, 2008). This trans- gression coincided with a Middle Miocene (14– 16.5 Ma) global warming (Fig. 7.26), which by contrast appeared to stimulate the development of further diversity of the miogypsinids in the Indo- Pacific province. In the late Langhian–e arly Serravallian (13– 14 Ma), a time of several regressions (Fig. 7.26), the short- lived marine reconnection between the Mediterranean and Indian ocean again closed (Von Rogl, 1998). This final closure coincided with the onset of global cooling (Fig. 7.26), and by the time the East Antarctic Ice Sheet was established (12 Ma), the miogypsinids had become extinct from the Indo- Pacific province (late Serravallian). This final miogypsinid extinction was also globally accompanied by the extinction of 60% of all other larger benthic foraminifera forms (BouDagher- Fadel, 2008). 604 Evolution and Geological Significance of Larger Benthic Foraminifera In contrast to the apparent isolation of Mediterranean forms from the Americas from the Early Chattian onwards, it seems that trans- Atlantic migration to South Africa remained possible up to the late Chattian (see BouDagher- Fadel and Price, 2013), thus allowing the direct migration of the early American miogypsinids (not just the Neorotalia seen in the Mediterranean) to South Africa. Subsequently, the South African miogypsi- nids did become isolated (perhaps again owing to sea- level rises or changes in oceanic cur- rents), and evolved lineages that were independent from their American ancestors. In the Burdigalian, South African species (Plate 7.11, figs 6-8 ) were still morphologically close to their American ancestors, but never evolved into the advanced forms of the megalo- spheric generations seen in the American, Mediterranean and Indo-P acific Miogypsina. This comparatively slow rate of evolution might reflect the lack of environmental diver- sity and the more temperate conditions in which the South African miogypsinids found themselves. Eventually, however, increased biological stress between the Burdigalian and Langhian (Figs 7.26 and 7.27), perhaps linked to rapid sea- level changes and global tectonic events, such as the eruption of the Columbia River flood basalt (15–1 7 Ma), caused the extinction of the South African miogypsinids and they were replaced by other forereef larger benthic foraminifera, such as Operculina and Heterostegina. On the other hand, in the American province, the extinction of the miogypsinids, coeval with a global warming event (Fig. 7.26), the so- called mid- Miocene Climatic Maximum (Fig. 7.27), resulted in their replacement by small nummulitic forms, such as Operculinoides species (see BouDagher- Fadel and Price, 2010b; BouDagher- Fadel et al., 2010). It should be noted however, that the claim of provincialism for some genera should be treated with caution, as limited observations may be due to poor preservation. In some cases, genera which were thought to be endemics to a typical province have later been found in another province. Thus, Discospirina, a milioline genus recorded as a fossil only from the Middle Miocene to Pleistocene of the Mediterranean, is still found living in the same area and into the North Atlantic Ocean (Adams, 1959, 1967, 1973; Adams et al., 1983). This genus has some biogeographical importance as it was consid- ered to have been the only genus of larger foraminifera that seemed “to have originated outside the Indo- Pacific region since Early Miocene times” (Adams, 1973, p.  465). Adams et al. (1983) used the absence of records of Discospirina from the Indian Ocean and Indo- Pacific region to support their contention that a permanent land barrier has separated the Mediterranean and Indian Ocean regions since Middle Miocene times (Fig. 7.33). However, Chaproniere (1991) collected similar forms from the Coral Sea, separated by a considerable distance from those recorded from the Mediterranean and Atlantic Ocean. He suggested, that their presence in the Coral Sea indicates that they must have escaped from the North Atlantic or the Mediterranean, either before it was closed to the Indian Ocean in the Serravallian, or after closure, possibly during a per- iod of high sea level. On the other hand, as Discospirina appears to have been able to live in deeper waters and therefore cooler temperatures, it may have been able to migrate from the southern Atlantic Ocean around the southern tip of Africa. Another alternative explanation is that this genus arose independently in the Coral Sea area. However, according to Chaproniere (1991) this appears unlikely due to the close sim- ilarity between the populations. The previous lack of records of this genus outside of the North Atlantic and Mediterranean regions may well be the result of the extreme delicacy of the very thin discoidal test, which would certainly be readily fragmented during post- mortem movement on the sea floor. The Cenozoic Larger Benthic Foraminifera: The Neogene 605 7.6 Epilogue In modern hydrocarbon exploration, it is important to understand the environment of deposition of any carbonate sequence so as to predict the presence, quality and thick- ness of reservoir, seal and source rocks. As we have seen throughout this volume, larger benthic foraminiferal associations, along with light-s ensitive calcareous algae, are use- ful as depth and age indicators. Their diversity can be used as an indicator of the locus of facies evolution (see Fig. 7.36), palaeogeography and palaeotemperature. The study of the systematics of larger foraminifera is an essential tool for biostratig- raphy. As seen in the previous chapters, in larger foraminifera growth history is recorded Teth. Am. Backreef Reef Forereef Figure 7.36. Characteristic foraminiferal assemblages in different parts of the reef ecology through time (Teth. = Tethys, Am. = Americas) for the provincialism characteristic of the Paleogene. Neogene Paleogene Mesoczi Palaeozcoi 606 Evolution and Geological Significance of Larger Benthic Foraminifera and preserved as a part of the test, and their evolution is recorded and defines the geo- logical column. Therefore, well-d efined lineages can be identified and traced through- out their history. They seem to follow Cope’s rule of a trend towards complexity (see Chapter 1). Primitive forms persist and survive harsh conditions and major extinction events, but also give rise gradually though time and space to more complex specialist forms that flourish in stable environments. Cope’s rule has been linked by many authors to being a reflection of “K-s election” and hypermorphosis (Gould, 1977). Larger fora- minifera thrive in many environments, and their delayed reproductive strategy leads to a long life span, reaching a hundred years in some large Nummulites. This combi- nation of a long life span, and abundance in the geological record make them a most valuable tool for biostratigraphy in shallow water environments. On the other hand, by having test features that continually change with time, larger foraminifera are useful in the study of the genetic and morphological basis of evolution. One important future development would be to combine studies of living forms with the valuable and rich fossil record in order to understand the relationship between genetic characteristics and morphology. Work on the origin of the foraminifera and on phylogenetic analysis of verified fora- miniferal DNA sequences are already underway (Wray et al., 1995; Holzmann et al., 2003). Genetic analyses are essential to understand phylogenetic relationships among larger foraminiferal genera and species. They will help to establish whether the char- acteristics developed by different forms are analogous (i.e. have the same appearance, but a different origin or function) or homologous (i.e. having the same origin). The ten- dency, noted in earlier chapters, towards evolutionary convergence is something which requires deeper analysis. Direct genetic comparison between the specimens from the European- Middle Eastern and the Far Eastern areas will also be essential to confirm the correlation recently made between the standard Planktonic Foraminiferal biozones and the geostratigraphic stages of Western Europe, that have enabled refinement of the “Letter Stages” of the Far East and recalibration of them with the European Stages (BouDagher- Fadel, 2013). This is helping to contribute to the understanding of tec- tonics and the timing of terrane/ microplate suturing as reflected by the distribution of present day faunas. For example, the final suturing of Arabia and Eurasia is dated as Late Oligocene and shows up well in the similarity of Miocene land faunas across the suture. The Tethys was therefore finally divided into the Mediterranean and the Indian Ocean at this time. Further investigation of larger foraminifera will show if the foraminifera of the Levant demonstrate increasing divergence from those of the east- ern part of Arabia and the Far East, and are shedding light on the Himalayan orog- eny (e.g. Najman et al., 2010a, b; BouDagher- Fadel et al., 2017; Najman et al., 2016; Hu et al., 2015). Culture experiments under controlled environmental conditions such as tempera- ture, salinity, dissolved oxygen content, nutrient content, trace element concentration, and isotopic enrichment (building upon the pioneering work of Röttger) should be encouraged in order to better interpret natural ecological behaviour and environmen- tal tolerances. The driving mechanisms behind larger foraminiferal evolution have been controlled by changes in palaeoenvironments, on a whole range of time scales from those driven by tectonics, through those associated with climate change and finally those affected by catastrophic volcanic or impact processes. Micropalaeontological The Cenozoic Larger Benthic Foraminifera: The Neogene 607 correlation will clarify the role played by biogeographical restrictions. Controlled experiments on, for example, the tolerance of different forms to water depth could enable estimates of such physical variables as subsidence rates to be made. The com- bination of depth and time calibration would help resolve some of the great sequence stratigraphy debates as to whether particular transgressions are precisely isochronous worldwide or are due to regional eustatic sea- level rises. Finally, it is widely held these days that human activity is destroying coral reefs worldwide. As highlighted by Hallock (2005) for example, more research on living fora- minifera will give us a greater understanding of contemporary reef ecologies and will help us to understand how reef- dwelling fauna respond to changes to environmen- tal parameters caused, for example, by climate change and/ or other human generated processes. As we have seen in this chapter, there is still much to discover about both the mode of life, evolutionary trends and the migratory habits of Neogene and existent forms. If this is the case for contemporary and geologically recent foraminifera, then it must be the case also for Early Cenozoic, Mesozoic and Palaeozoic forms. Therefore, although an attempt has been made in the preceding chapters to provide an overview of what is known about the fossil larger benthic foraminifera, it must be concluded that the study of these geologically important creatures, although mature, is far from complete, and should be pursued by future generations with vigour, as they are certainly highly versatile, biologically adaptive and flexible, and so still make a challenging subject for research. 3 1 2 7 6 4 5 B 9 A 8 10 11 12 14 13 15 16 17 18 19 20 Plate 7.1 Scale bars: Figs 1-5 , 7-1 0, 12-2 0 = 0.5mm; Fig. 6, 11 = 0.25mm. Figs 1-3 . Austrotrillina howchini (Sclumberger), closely spaced bifurcating alveoli, 1) Early Miocene (lower Tf), Pata Limestone, Australia, B- Form, NHM P47608; 2) Early Miocene, Lower Chake Beds, Pemba, Tanzania, NHM P22848; 3) Lower Miocene, Bairnsdale, Chowilla Dam Site, Pata Limestone, Australia, NHM P47587. Figs 4-7 . Austrotrillina striata Todd and Post, simple coarse alveoli, 4- 5, 7) Late Oligocene, Borneo, UCL coll; 6) topotype, Early Miocene, Bikini Island, NHM P47596. Figs 8- 9. Austrotrillina asmariensis Adams, numerous alveoli, Early Miocene (Aquitanian), Asmari Limestone, Luristan, Iran, NHM P47585. Fig. 10. A) Austrotrillina howchini (Sclumberger), B) Miogypsinoides sp., Early Miocene (lower Tf1), Loc. 238, Indonesia, UCL coll. Fig. 11. Austrotrillina brunni Marie, Early Miocene (Aquitanian), Nias, Indonesia, UCL coll. Fig. 12. Austrotrillina howchini (Sclumberger), Early Miocene (Tf1), Hadu Village, Coast Province, Kenya, NHM P43944. Figs 13- 14. Borelis pygmaeus Hanzawa, Early Miocene (Aquitanian), 13) equatorial section; 14) Loc. 205, Borneo, UCL coll. Fig. 15. Borelis melo (Fichtel and Moll), Middle Miocene, Turkey, NHM P49087. Fig. 16. Borelis haueri (d’Orbigny), type, Middle Miocene, Baden- Brickyard, Baden, Australia, UCL coll. Fig. 17. Borelis pulchra (d’Orbigny), Holocene, Mauritius, UCL coll. Fig.  18. Borelis melo (Fitchel and Moll), Middle Miocene (Serravallian), Barden, Australia, NHM coll. Fig. 19. Borelis curdica (Reichel), Miocene, Turkey, NHM coll. Fig. 20. Alveolinella praequoyi Wonders and Adams, holotype, early Middle Miocene (upper Tf1- Tf2), Darai Limestone, Papua New Guinea, NHM P52658. Plate 7.2 Scale bars:  Figs 1-6 , 8- 12, 14- 16  =  1mm; Fig.  7, 13, 18  =  0.5mm; Fig.  17  =  100μm. Figs 1, 5. Alveolinella sp., Miocene, Port Moresby Bay, Papua New Guinea, UCL coll., 1) solid specimen; 5) equa- torial section. Figs 2-4 . Alveolinella quoyi (d’Orbigny), Holocene, 2-3 ) Pacific; 4) Port Moresby Bay, Papua New Guinea, solid specimen, UCL coll. Figs 6- 7. Alveolinella praequoyi Wonders and Adams, holotype, early Middle Miocene (upper Tf1- Tf2), Darai Limestone, Papua New Guinea, NHM P52659- 60. Fig. 8. Alveolinella fennemai Checchia-R ispoli, late Early Miocene (Burdigalian), Borneo, UCL coll. Figs 9- 11. Flosculinella bontangensis (Rutten), 9) late Early Miocene (Burdigalian, lower Tf1), Borneo; 10-1 1) Middle Miocene (middle and upper Tf1), East Java, UCL coll. Fig.  12. Flosculinella globulosa (Rutten), Early Miocene (lowermost Burdigalian), Indonesia, UCL coll. Fig. 13. Dendritina rangi d’Orbigny, Early Miocene (Aquitanian), Kirkuk Well no.22, Iraq, NHM P39432. Fig.  14. Archaias angulatus (Fichtel and Moll), Holocene, offshore Belize, UCL coll. Fig.  15. Cyclorbiculina sp., Holocene, off Blackadare Bay, British Honduras, UCL coll. Figs 16- 17. Marginopora vertebralis Blainville, SEM pictures; 17) enlargement of an equatorial chamber of the fig. 16, Papua New Guinea, UCL coll. Fig. 18. Orbiculina sp., Early Miocene (Burdigalian), Upper Chake Beds (Datum Stratum), Wesha Road, Pemba, Tanzania, NHM P22853. 1 2 3 4 5 6 8 7 9 10 14 12 13 11 16 17 15 18 1 2 3 4 5 6 10 7 8 9 11 12 13 14 15 16 17 18 Plate 7.3 Scale bars:  Figs 1- 6, 9- 12  =  1mm; Figs 7- 8, 13, 15- 18 =0.5mm; Fig.  14  =  0.25mm. Figs 1- 5. Marginopora sp., Holocene, Papua New Guinea, on seagrass, 1- 2) SEM photograph; 2)  enlargement of chambers showing rectangular chamberlets, UCL coll. Figs 6-8 . Sorites sp., Holocene, Byblos, Lebanon, UCL coll. Figs 9 Archaias sp. Holocene, Funafuti sandy beach, Tuvalu, South Pacific, UCL coll. Fig. 10. Discospirina italica (Costa), Late Miocene, Pakhna Formation, Anaphotia, Cyprus, UCL coll. Fig.  11. Cyclorbiculina compressa d’Orbigny, Holocene, off Belize, UCL coll. Fig. 12. Cyclorbiculina sp. Funafuti Sandy Beach, Tuvalu, South Pacific, UCL coll. Fig. 13. Archaias kirkukensis Henson, Kirkuk Well no. 57 Iraq, NHM P39633. Fig.  14. Peneroplis sp., Holocene, offshore Byblos, Lebanon, UCL coll. Figs 15-1 6. Peneroplis thomasi Henson, paratypes, Early Miocene (Aquitanian), Kirkuk Well no.57, Iraq, NHM P39634. Fig. 17. Planorbulinella larvata (Parker and Jones), Early Miocene, Java, UCL coll. Fig. 18. Praerhapydionina delicata Henson, common in both Aquitanian and Burdigalian assemblages of the Bahamas (Fischer coll., University of Geneva). Plate 7.4 Scale bars: Figs 1- 5; 7- 12 = 1mm; Figs 6, 13- 15 = 0.5mm. Figs 1- 6, 12. Pseudotaberina malabarica (Carter), 1- 3) Early Miocene, Malabar coast, India’s southwestern coast, NHM P29875, 1) axial section, 2) equatorial section, 3) enlargement of fig. 2; 4-6 ) Early Miocene (Burdigalian), Darai Limestone, Papua New Guinea, UCL coll; 12) Early Miocene (Burdigalian), Java, UCL coll. Figs 7- 8. Alanlordia banyakensis Banner and Samuel, paratypes, Middle Miocene (middle Serravallian), Banyak, Indonesia, NHM P52877- 79. Figs 9- 11. Alanlordia niasensis Banner and Samuel, topotypes, Middle Miocene (middle Serravallian), Nias, Indonesia, UCL coll., 9) with Planostegina sp. Figs 13-1 4. Sphaerogypsina globulus (Reuss), Middle Miocene (Serravallian), Cina Village, Indonesia, UCL coll. Fig.  15. Sphaerogypsina sp., Early Miocene (Burdigalian), Darai Limestone, Australia, this species differs from S. globulus by having thicker walls, more distinctly perforate chambers, not arranged in radial rows but alternating in annuli, and from Discogypsina vesicularis (Parker and Jones) but sub- spherical growth pattern. 1 3 4 2 5 7 8 6 9 10 11 12 13 14 15 2 3 1 6 7 4 5 10 9 11 8 15 13 14 12 16 17 A B 18 19 20 Plate 7.5 Scale bars:  Figs 1- 14  =  0.5mm; Figs 15- 16, 18, 20  =  1mm; Frig. 17, 19  =  0.3mm Fig.  1. Planorbulinella kinabatangenensis Renema, holotype, Lower Miocene (Te1-4 ) Batu Temenggong Besar, Lower Kinabatangan River, North Borneo, NHM NB9023a. Figs 2-5 . Planorbulinella solida Belford, Early Miocene (Tf), Nias, Indonesia, UCL coll. Fig.  6. Planorbulinella batangenensis adamsi Renema, Early Miocene, from loose block about 17 feet above Lower Kinabatangah River, Batu Temenggong Besar, North Borneo, NB9030a, Adams coll., NHM P67225. Figs 7- 9. Planorbulinella larvata (Parker and Jones), Early Miocene, Java, UCL coll. Fig.  10. Elphidium/C ellanthus craticulatus, Jabal Terbol, Lebanon, UCL coll. Figs 11- 14. Calcarina sp., 11) Miocene, Borneo; 12-1 4) Pliocene, Philippines, showing canaliferous spines, UCL coll. Figs 15- 19. Cycloclypeus eidae Tan Sin Hok, Lower Miocene, Kinabatangan River, Sabah, North Borneo, 15)  axial section, NHM coll., NB9066; 16- 17) NHM P49518, 17)  enlargement of early part of test of fig.  16; 18- 19) NHM NB9067,19) enlargement of early part of test of fig.  18. Fig.  20. Thin sec- tion photomicrograph of A) L. (Nephrolepidina) ferreroi Provale, B) Cycloclypeus carpenteri Brady, Middle Miocene(Langhian), Nias, Sumatra, UCL coll. Plate 7.6 Scale bars: Figs 1-4  = 1mm; Fig. 5- 7, 9- 13 = 0.5mm; Fig. 8 = 0.3mm; Fig. 14 = 0.2mm. Figs 1-4 . Cycloclypeus carpenteri Brady, 1-3 ) offshore Borneo, UCL coll; 4) Miocene, Funafuti, Tuvalu, UCL coll. Fig 5. Cycloclypeus sp., Dutch New Guinea, NHM P22790. Figs 6, 8, 11. Cycloclypeus indopacificus Tan Sin Hok, Early Miocene (Burdigalian, lower Tf1), East Borneo, UCL coll. Fig. 9. Lepidocyclina murrayana, Cycloclypeus cf. orbitoitades Schmutz, Miocene, Dutch New Guinea, NHM P22791. Figs 7, 10. Cycloclypeus pillaria BouDagher-F adel, 7) Middle Miocene (Serravallian, Tf2), Bulu Mampote, Indonesia, NHM 67125; 8)  Early Miocene (Burdigalian, upper Te), Loc. 620/ 14, Borneo, UCL coll. Figs 12- 13. Katacycloclypeus annulatus (Martin), 12) Miocene (Serravallian, Tf2), Futuna Limestone, Vanua Mbalavu Island, Fiji, NHM P50298; 13) Middle Miocene, Java, NHM P36377. Fig. 14. Amphistegina sp., Holocene, offshore Lebanon, SEM photograph, enlargement of the apertural face showing pustules surrounding the aperture, UCL coll. 1 2 3 4 5 6 7 8 11 9 10 12 13 14 1 2 3 4 5 6 8 7 9 Plate 7.7 Scale bars:  Figs 1- 9  =  1mm. Figs 1- 5. Katacycloclypeus annulatus (Martin), 1- 3) Miocene (Serravallian, Tf2), Darai Limestone, Papua New Guinea; 4- 5) late Middle Miocene Fiji, NHM P50462- 3. Figs 6- 7. Katacycloclypeus martini (van der Vlerk), 6)  Miocene (Serravallian, Tf2), Sumatra; 7)  Miocene Serravallian (Tf2), East Java, UCL coll. Fig.  8. Katacycloclypeus annulatus (Martin), Miogypsina sp., Amphistegina sp., Victoriella sp., Nias, Sumatra. Fig. 9. Cycloclypeus indopacificus Tan Sin Hok, Dutch New Guinea, NHM P22785. 1 2 3 4 11 5 6 7 8 9 Plate 7.8 Scale bars: Fig. 1- 4, 7-9  = 1mm; Figs 5- 6 = 0.5mm. Fig. 1. L. (Nephrolepidina) ferreroi Provale, Miogypsina tani Drooger, Miocene (early Langhian, middle Tf1), Nias, Sumatra, UCL coll. Fig.  2. Miogypsina sp., Miocene (early Langhian, middle Tf1), Nias, Sumatra, UCL coll. Fig. 3. Austrotrillina how- chini (Sclumberger), Borelis sp., Sorites sp., Miocene (Tf1), Coast Province, Kenya, NHM P43946. Fig. 4. Spiroclypeus sp., Oligocene, East Java, UCL coll. Fig. 5. A packstone of Miogypsinella ubaghsi (Tan Sin Hok), Miocene (early Aquitanian, early Te5), Darai Limestone, Papua New Guinea, UCL coll. Fig.  6. Lepidocyclina gigantea (Martin), Miocene (Serravallian, Tf3), Darai Limestone, Papua New Guinea, UCL coll. Fig.  7. Katacycloclypeus annulatus (Martin), Lepidocyclina sp., Miocene (Serravallian, Tf2), Borneo, UCL coll. Figs 8- 9. Eulepidina ephippioides (Jones and Chapman), and Heterostgina (Vlerkina) borneensis van der Vlerk, Chattian, Mosque Quarry, Borneo, UCL coll. Plate 7.9 Scale bars: Figs 1- 4, 6, 11-1 5, 17 = 1mm; Figs 5, 7-1 0, 16 = 0.5mm.. Figs 1-2 . Borodinia sp., Early Miocene (Aquitanian), Java, UCL coll. Fig.  3. Gypsina sp., Carpenteria sp., algae, Early Miocene (Burdigalian), Loc. 251, Borneo, UCL coll. Figs 4-5 . Homotrema rubrum (Lamarck), Holocene, Indo-P acific, exact locality unknown, 4) vertical section from a specimen attached to Acropora millepora specimens NHM coll; 5) John Murray exp.1956, NHM coll. Figs 6- 7. Carpenteria proteiformis Goës, Holocene, off Barbados, 100 fathoms, NHM coll., 6) vertical section; 7) horizontal section of tip of right hand chamber. Figs 8- 9. Rupertina stabilis Wallich, Holocene, cold area Faroe Channel, 8) vertical, 9) basal horizontal section. Figs 10- 13. Miniacina miniacea (Pallas), Holocene, 10) Tizard Bank, China Sea, Heron- Allen- Earland coll.; 11) John Murray expedition, Station 178, 91m, NHM coll; 12- 13) Early Miocene, Java, UCL coll. Fig. 14. Sporadotrema cylindricum (Carter), Maldives Islands, Murray expedition, Station 149, NHM coll. Fig. 15. Thin section photomicrograph of A) Lepidosemicyclina sp., B) Carpenteria sp., Early Miocene (Burdigalian), Borneo, UCL coll. Fig. 16. Thin section photomicrograph of A) L. (Nephrolepidina) sumatrensis Brady, B) Victoriella sp., Early Miocene (Burdigalian), East Java, UCL coll. Fig.17. Sporadotrema mesentericum (Carter), Holocene, Maldives Islands, John Murray expedition, Station112, 113m, NHM coll. 3 1 2 4 4 5 6 8 9 7 11 10 14 13 12 B A B A 16 15 17 1 2 3 5 4 6 A B 8 9 7 10 13 11 12 14 Plate 7.10 Scale bars: Figs 1, 5- 8, 11, 13- 14 = 1mm; Figs 2- 4, 9- 10 12 = 0.5mm. Figs 1- 4. Carpenteria sp., 1) Early Miocene, East Java; 2). Early Miocene (Burdigalian), Borneo; 3) Early Miocene (Aquitanian), Borneo; 4) Middle Miocene, UCL coll. Fig.5. Baculogypsina sp., Holocene, forereef, Coral Sea, Papua New Guinea, UCL coll. Fig. 6. Baculogypsina baculatus (Montfort), Holocene, New Zealand, NHM coll. P83. Fig. 7. Baculogypsina sphaerulata Parker and Jones, paralectotype, Holocene, Ex. ZF 3598, Rewa Reef, Fiji, Parker coll. Fig. 8. Calcarina hispida Brady var. pulchra Chapman, Holocene, Funafuti Atoll, Ellis Islands, UCL coll. Fig. 9. Planorbulinella larvata (Parker and Jones), Holocene, John Murray expedition, Station103, 101m, NHM coll. Figs 10- 12. Victoriella sp., 10) Childs Bank Burdigalian, Ka1, South Africa, McMillan coll. 11) Early Miocene (Burdigalian), Borneo, UCL coll.; 12) Java, UCL coll. Fig. 13. Biarritzina sp., Early Miocene, Brazil, UCL coll. Fig. 14. Miniacina miniacea (Pallas), Holocene, John Murray expedition, Station 178, 91m, NHM coll. Plate 7.11 (13 Scale bars: Figs 1-1 5 = 0.5mm; Fig. 16 = 0.25mm. Fig. 1. Miogypsinodella corsicana Ferrandini et al., paratype, figured by Ferrandini et al. (2010b), Miocene (Burdigalian), Bonifacio, Corsica, France. Figs 2- 4. Miogypsinodella primitiva BouDagher-F adel and Lord, Miocene (Burdigalian), Nias, Sumatra, UCL coll. Fig. 5. Miogypsina antillea (Cushman), Early Miocene, Brazil, UCL coll. Fig. 6. Miogypsina southernia BouDagher-F adel and Price, paratype figured by BouDagher-F adel and Price (2013), Burdigalian, Childs Bank, Ka1, 360m, South Africa, McMillan UCL coll. MF275. Fig. 7. Miogypsina mcmillania BouDagher- Fadel and Price, paratype figured by BouDagher- Fadel and Price (2013), Burdigalian, Childs Bank, Ka1, 400m, South Africa, McMillan coll. UCL MF243. Fig.  8. Miogypsina africana BouDagher-F adel and Price, paratype figured by BouDagher- Fadel and Price (2013), Burdigalian, Childs Bank, Ka1, 36m, South Africa, McMillan UCL coll. MF250. Fig. 9. Miogypsina samuelia BouDagher-F adel and Price, holotype figured by BouDagher- Fadel and Price (2013), Nias, Indonesia. Fig. 10. Miogypsina niasiensis BouDagher- Fadel and Price, holotype figured by BouDagher-F adel and Price (2013), Miocene (Early Langhian, Tf1), Nias, Indonesia, McMillan UCL coll. MF205. Fig.  11. Miogypsina cushmani Vaughan, Miocene (latest Burdigalian), Castelsardo section (northern Sardinia), equivalent of Cala di labra Formation, Ferrandini UCL coll. MF259–6 0. Fig.  12. Lepidosemicyclina thecideaeformis Rutten, Miocene (Burdigalian), Upper Chake Beds, Pemba, Tanzania, NHM P22849. Fig. 13. Miogypsina sabahensis BouDagher- Fadel, Lord and Banner, holotype, Miocne (early Burdigalian), North East Borneo, NHM P66907. Fig. 14. Miolepidocyclina mexicana (Nuttall), Miocene (Aquitanian), Brazil, UCL coll. Fig. 15. Miolepidocyclina excentrica Tan Sin Hok, Cala di Labra Formation, Cala di Ciappili section (Bonifacio area), Ferrandini coll. Fig. 16. Miogypsina indonesiensis Tan Sin Hok, Middle Miocene (Tf3), Yule Island Region, SW Papua, UCL coll. 2 1 3 5 7 4 6 9 8 10 12 11 13 14 15 16 2 1 3 4 8 7 5 6 a 9 b10 11 Plate 7.12 Scale bars:  Figs 1- 11  =  1mm. Fig.  1. Lepidocyclina (Nephrolepidina) morgani Lemoine and Douvillé, Miocene (Aquitanian), Nerthe, near Marseille Petit Nid, France, Ferrandini UCL coll. Figs 2- 3. Lepidocyclina delicata Scheffen, Miocene (Serravallian, Tf2), Darai Limestone, Papua New Guinea, UCL coll. Figs 4-5 . Lepidocyclina angulosa quadrata Provale, Miocene (Serravallian), Darai Limestone, Papua New Guinea, UCL coll. Fig. 6. Lepidocyclina (Nephrolepidina) angulosa Provale, NCS-3 , Nam Con Son basin, Vietnam, Vova’s coll. Fig. 7. Lepidocyclina banneri BouDagher-F adel, Noad and Lord, holotype, Miocene (Aquitanian), Borneo, NHM P66905. Fig.  8. Nephrolepdina radiata (Martin), Miocene (Serravallian), Darai Limestone, Papua New Guinea, UCL coll. Fig. 9. Lepidocyclina (Nephrolepidina) suwanneensis Cole, Miocene (Aquitanian), Retrench member, Cipero Formation, Trinidad, UCL coll. Fig. 10. Lepidocyclina (Nephrolepidina) sondaica Yabe and Hanzawa, Miocene (Aquitanian), Loc. 205, Borneo, UCL coll. Fig. 11. a) Lepidocyclina (Nephrolepidina) brouweri Rutten, b) Lepidosemicyclina sp., Miocene (Burdigalian, early Tf1), Borneo, UCL coll. Plate 7.13 Scale bars: Figs 1- 21 = 0.5mm. Figs 1- 3. Miogypsinoides dehaarti (van der Vlerk), 1- 2) Miocene (Burdigalian), Waterfall Section, Christmas Island, Indian Ocean, NHM 6764 1295; 3)  late Miocene (Aquitanian), Sulawesi, UCL MF167. Figs 4- 6. Miogypsina gunteri Cole, 4)  Miocene (Aquitanian), top of Superga Mountain, Aman, about 2km S.E.  of Superga, NHM coll.; 5)  Miocene (Aquitanian), Superga Mountain, Aman, about 2 km SE of Superga, UCL coll. MF257; 6)  Miogypsina globulina (Michellotti), Miocene (early Langhian), Cala di Labra Formation, Cala di Ciappili section (Bonifacio area), Corsica, France, Ferrandini UCL coll. Fig. 7. Miogypsina orientalis BouDagher-F adel et al., Miocene (late Burdigalian), Kalimantan, East Borneo, UCL coll. Fig.  8. Miogypsina kotoi Hanzawa, Miocene (Burdigalian), Darai Limestone, Papua New Guinea, UCL coll. Figs 9- 13. Miogypsina borneensis Tan Sin Hok, 9- 10) Miocene (Aquitanian, upper Te), Darai Limestone, Papua New Guinea; SEM photographs of a microspheric specimen half dissected showing embryonic apparatus and equatorial chambers; 10) enlarge- ment of an equatorial chamber of (9) showing modifications (called egg-h olders by Hottinger, 2006) to house symbionts; 11- 13) Miocene (early Burdigalian, lower Tf1), UCL coll. Fig.  14. Miogypsina bifida Rutten, Miocene, Sungai, Boengaloen, East Borneo, NHM. Van Vessem coll. BB 469 – 1913. Fig. 15. Miogypsina sp., Miocene (Burdigalian), Morocco, UCL coll. Senn27778a. Fig. 16. Miogypsina cushmani Vaughan, Miocene (latest Burdigalian) (note in Chart 7.1, this form appears later in time in the Far East), Castlesardo sec- tion, equivalent of Cala di Labra formation, Northern Sardinia, Ferrandini UCL coll. Fig. 18. Miogypsina indonesiensis Tan Sin Hok, Miocene (Serravallian), Nias, UCL MF204. Fig.  19. Miogypsina mediterra- nea Brönnimann, Miocene (late Burdigalian), Cala di Ciappili section, Bonifacio area, southern Corsica, France, Ferrandini UCL coll. MF243. Fig. 19. Miogypsina intermedia Drooger, Miocene (late Burdigalian), Capo Testa, Ferrandini UCL coll. Fig. 20. Miogypsina subiensis BouDagher- Fadel and Price, paratype fig- ured by BouDagher-F adel and Price (2013), Miocene (Burdigalian), Subis Formation, Borneo. Fig.  21. Miolepidocyclina excentrica Tan Sin Hok, Cala di Labra Formation, Cala di Ciappili section, Bonifacio area, southern Corsica, France, Ferrandini UCL coll. 1 2 3 4 5 7 6 8 9 10 11 12 13 14 15 16 17 18 19 20 21 4 1 2 3 5 6 7 8 9 11 10 12 13 14 15 16 19 17 18 20 Plate 7.14 Scale bars:  Figs 1- 20  =  0.5mm. Figs 1- 2. Neorotalia tethyana BouDagher- Fadel and Price, Miocene (Aquitanian), Loc. 621/ 3, Borneo. Fig. 3. Miogypsinella borodinensis Hanzawa, Late Oligocene, Loc. 204, Borneo, UCL coll. Figs 4- 5. Miogypsinella ubaghsi (Tan Sin Hok), Miocene (early Aquitanian, Te5), Darai Limestone, Papua New Guinea, UCL coll. Fig. 6. Boninella sp., Operculina sp., Lepidocyclina (L.) isolepidinoides van der Vlerk, Late Oligocene, Loc. 239, Borneo, UCL coll. Fig. 7. Paleomiogypsina sp., Late Oligocene (Chattian), Loc. 621/3 , Borneo, UCL coll. Figs 8-1 0. Miogypsinella bornea BouDagher- Fadel and Price, 8)  Early Miocene (Aquitanian), Borneo, UCL coll.; 9- 10) figured by BouDagher- Fadel and Price (2013) from Miocene (early Chattian– early Aquitanian), Java, 9) holotype; 10) paratype. Fig. 11. Miogypsinoides formosensis Yabe and Hanzawa, Miocene (early Aquitanian), Nerthe area, near Marseille, Petit Nid section, Formation pararecifale du Cap de Nautes, France, Ferrandini UCL coll. MF252 Figs 12-1 4. Miogypsinoides dehaarti (van der Vlerk), Miocene (Burdigalian), 12, 14) Cyprus; 13), Nias, Sumatra, 201 AS, UCL coll. Figs 15-1 7. Miogypsinoides bantamensis Tan Sin Hok, Miocene (earliest Burdigalian), 15,17) figured by Ferrandini et  al. (2010) and BouDagher-F adel and Price (2013) from Castlesardo sec- tion, equivalent of Cala di Labra formation, Northern Sardinia, Ferrandini UCL/F errandini coll. MF253; 16) figured by Ferrandini et al (2010), from the Nerthe area, near Marseille, Petit Nid section, Formation pararécifale du Cap de Nautes, France, UCL/F errandini coll. Fig. 18. Miogypsina mcmillania BouDagher- Fadel and Price, paratype, figured by BouDagher-F adel and Price (2013), Miocene (Burdigalian), Childs Bank, Ka1, 400m, South Africa, UCL MKF240. Fig. 19. Miogypsinoides mauretanica Bronnimann, figured by Ferrandini et al. (2010) from, Castlesardo section, Northern Sardinia. Fig. 20. Miogypsina triangulata BouDagher- Fadel and Price, paratype, figured by BouDagher- Fadel and Price (2010b), Early Miocene (N4), offshore Brazil, UCL coll. Plate 7.15 Scale bars: Figs 1, 4-6 , 8- 14 = 1mm; Figs 2- 3, 7, 15 = 0.5mm. Fig. 1. Operculinoides panamen- sis (Cushman), Miocene (Aquitanian), Brazil, UCL coll. Fig. 2. Operculina sp., Early Miocene, Sentosa, UCL coll. Fig. 3. Operculinoides antiguensis Vaughan and Cole, Early Miocene, Juana Diaz Formation, La Rambla, Puerto Rico, NHM 47351-2 . Figs 4- 5. Heterostegina sp., Miocene (Te 1- 4), Batu Temongong, Besar, Lower Kinabatangan River, North Borneo, 4)  NHM NB9023; 5)  NHM NB9020. Fig.  6. Heterostegina (Vlerkina) borneensis van der Vlerk, Heterostegina (Vlerkina) sp., Early Miocene, Loc. 205, Borneo, UCL coll. Fig. 7. Tansinhokella yabei (van der Vlerk), Miocene, S. Soembal, E. Borneo, NHM P45042. Fig. 8. L. (Nephrolepidina) nephrolepidinoides BouDagher- Fadel and Lord, Heterostegina (Vlerkina) borneensis van der Vlerk, Early Miocene, Loc. 239, Borneo, UCL coll. Figs 9- 12. Thin section photomicrograph of Spiroclypeus sp., L. (Nephrolepidina) sp., L. (Nephrolepidina) brouweri Rutten, Miocene (early Burdigalian, Te5), Indonesia, UCL coll. Fig. 13. Spiroclypeus tidoenganensis van der Vlerk, Miocene, S. Patoeng (Antjam) E.  Borneo, NHM P45039. Figs 14-1 5. Discogypsina discus (Goës), Miocene, exposure 4, Kinabatangan River, North Borneo, NHM N.B. 9050. 1 2 3 7 6 4 5 10 8 9 11 14 13 12 15 1 2 3 5 6 4 8 7 9 10 9 14 11 12 15 13 17 18 16 Plate 7.16 Scale bars: Figs 1-5 , 15, 17 = 0.5mm; Figs 6-1 4, 16, 18 = 1mm. Fig. 1. Amphistegina lessonii d’Orbigny, Holocene, Mauritius, NHM coll. Figs 2-5 . L. (Nephrolepidina) sp. of an amphilepidine type, Miocene (late Aquitanian or Burdigalian), Lower Chake Beds, Pemba, Tanzania 2) NHM P22846; 3) NHM P22844; 4) NHM P22483; 5) NHM 22845. Figs 6- 7. Lepidocyclina (Nephrolepidina) batesfordensis Crespin, topotypes of tribliolepidine embryo, Miocene, Lepidocyclina Limestone, Upper Quarry, Batesford, Victoria, Australia, NHM P36060. Figs 8-9 . Lepidocyclina verbeeki (Newton and Holland), Early Miocene, Nias, Sumatra, NHM P45073. Figs 10-1 1. Lepidocyclina (Nephrolepidina) martini (Schlumberger), Miocene, Yule Island, Papua New Guinea, UCL coll. Figs 12- 13. Lepidocyclina (Nephrolepidina) rutteni quadrata van der Vlerk, showing a quadrate proloculus, Miocene (Serravallian, Tf2), Darai Limestone, Papua New Guinea, UCL coll. Fig.  14. Lepidocyclina murrayana Jones and Chapman, lectotype, Miocene, Christmas Island, NHM P22441. Fig. 15. L. (Nephrolepidina) inflata Provale, Miocene (Burdigalian, Tf1), Papua New Guinea, UCL coll. Fig. 16. Lepidocyclina acuta (Rutten), Miocene (Late Burdigalian), Papua New Guinea, UCL coll. Fig. 17. L. (Nephrolepidina) aff. epigona Schuber, Miocene (Tf3), Darai Limestone, Papua New Guinea, UCL coll. Fig. 18. L. (Nephrolepidina) soebandii van der Vlerk, Miocene (Burdigalian, upper Te), Loc. 205, Borneo, UCL coll. Plate 7.17 Scale bars: Figs 1- 2, 8, 10, 14, 16 = 0.5mm; Fig. 3 = 0.02mm; Fig. 4- 7, 9, 11- 13, 15 = 1mm. Fig. 1. Lepidocyclina sp., Miocene (Serravallian, Tf2), Darai Limestone, Papua New Guinea, UCL coll. Figs 2- 3. L. (Nephrolepidina) sp., 3) enlargement of fig. 2, Early Miocene, Java, UCL coll. Fig. 4. Eulepidina undosa laramblaensis Eames et al., holotype, Early Miocene, Juana Diaz formation, La Rambla, Puerto Rico, NHM P47333. Fig. 5. Lepidocyclina (Nephrolepidina) aquitaniae Silvestri, Miocene (Messinian), Spain, UCL coll. Fig. 6. Lepidocyclina canellei Lemoine and Douville, Oligocene, base of Antigua Limestone, Antiga, NHM P28629. Fig. 7. Lepidocyclina isolepidinoides van der Vlerk, Late Oligocene (lower Te), Borneo, UCL coll. Figs 8- 9. L. (Nephrolepidina) bikiniensis pumilipapilla Cole, Late Oligocene- Early Miocene, Juana Diaz Formation, La Rambla Puerto Rico, NHM P47346,9. Fig. 10. L. (Nephrolepidina) rutteni van der Vlerk, Miocene (Serravallian), Borneo, UCL coll. Figs 11- 12. Lepidocyclina mantelli (Morton), Early Oligocene, Marianna Limestone, Little Stave Creek, Alabama, 11)  axial section, NHM P47331; 12)  equatorial sec- tion, NHM P47326. Fig.  13. Lepidocyclina omphalus Tan Sin Hok, Miocene (Serravallian, Tf2), Papua New Guinea, UCL coll. Fig.  14. Thin section photomicrograph of Rodophyte sp., L. (Nephrolepidina) brouweri, Planorbulinella sp., Miocene (Tf1), Sumatra, UCL coll. Fig.  15. L. (Nephrolepidina) ferreroi Provale, Miocene (Burdigalian), sample 3813 from mollusc shell Limestone at Loa Duri (W. Outcrop) 100m from River East Kalimantan, Borneo, UCL coll. Fig.  16. L. (Nephrolepidina) ferreroi Provale, Miocene (Langhian), Sumatra, UCL coll. 1 2 3 4 5 6 7 9 11 8 10 12 14 13 14 15 16 1 2 3 5 6 4 8 9 14 7 10 11 12 Plate 7.18 Scale bars: Figs 1- 5, 9 = 1mm; Figs 6, 7- 8, 10- 12 = 0.5mm. Fig. 1. Lepidocyclina ngampelensis Gerth, Miocene (Serravallian, Tf2), Darai Limestone, Papua New Guinea, UCL coll. Fig. 2. Lepidocyclina murrayana Jones and Chapman, Early Miocene, Dutch New Guinea, NHM P22786. Fig. 3 Lepidocyclina (Nephrolepidina) angulosa Provale, Nam Con Son basin, Vietnam, Vova’s coll. NCS-3 . Fig. 4. Lepidocyclina aff. volucris Scheffen, Middle Miocene (Tf2), Darai Limestone, Papua New Guinea, UCL coll. Figs 5-7 . Lepidocyclina (Nephrolepidina) marginata (Michelotti) = Lepidocyclina (Nephrolepidina) tournoueri Lemoine and Douvillé, 5)  Oligocene, Cyprus, UCL coll.; 6)  Zakas section, Greece, NHM P51963; 7)  Oligocene, Oligocene (Rupelian, P20), Iran, Rahaghi NHM coll. Figs 8-9 . Lepidocyclina (Nephrolepidina) bikiniensis Cole, Juana Diaz formation, La Rambla, Puerto Rico, UCL coll. Fig. 10. Lepidocyclina (Nephrolepidina) bra- ziliana BouDagher-F adel and Price, holotype figured by BouDagher-F adel and Price, (2010a) from the early Oligocene (Rupelian, P18– P19), offshore Brazil, UCL 1MF. Fig. 11. Lepidocyclina (Lepidocyclina) boetonen- sis van der Vlerk, showing an isolepidine embryo, Middle Oligocene, Papua New Guinea, UCL coll. Fig. 12. Lepidocyclina (Nephrolepidina) sumatrensis Brady, Nam Con Son basin, Vietnam, Vova’s coll. NCS- 3. Plate 7.19 Scale bars:  Figs 1-1 0 =1mm. Figs 1- 3. Eulepidina badjirraensis Crespin, Early Miocene (Burdigalian, Te5), Mosque Quarry, Loc. 205, Borneo, UCL coll. Fig. 4. Eulepidina inaequalis Jones and Chapman, Indian Ocean, NHM P22539. Fig. 5. Eulepidina sp., Oligocene, Falling Waters Sink, Florida, NHM coll. Fig.  6. Eulepidina formosides (Douville), Oligocene, Spain, NHM P38679. Fig.  7. Eulepidina andrewsiana (Nuttall), Christmas Island, Indian Ocean, NHM P22533. Fig. 8. Eulepidina dilatata Michelotti, Oligocene (late Rupelian), Mesohellenic Basin, Greece, SFN coll. Fig. 9. Lepidocyclina (Nephrolepidina) brouweri Rutten, Miocene (early Burdigalian, Te5), Borneo, Loc. 205, UCL coll.. Fig.  10. Lepidocyclina (Nephrolepidina) sp. aff. gibbosa, Early Miocene (early Burdigalian, Te5), Borneo, Loc. 205, UCL coll. 1 4 5 2 6 7 3 9 8 10 Plate 7.20 Scale bars: Figs 1-7  = 1mm; Fig. 8 = 0.5mm. Figs 1-2 . Lepidocyclina (Lepidocyclina) canellei Lemoine and Douville´, offshore Brazil, UCL 10– 12MF. Figs 3- 5. L. (Nephrolepidina) sumatrensis Brady, 3- 4) Early Miocene (Te5), Malinau River, N. of Mount Mulu, N.W. Sarawak Borneo, NHM P46529; 5) Darai Limestone, Papua New Guinea, UCL coll. Fig.  6. Lepidocyclina (Lepidocyclina) isolepidinoides van der Vlerk, Oligocene, Dubai, NHM P51965. Fig. 7. L. (Nephrolepidina) pilifera Scheffen, Miocene (Burdigalian), Darai Limestones, Papua New Guinea, UCL coll. Fig. 8. L. (Nephrolepidina) sp., solid specimen, Miocene (Burdigalian), Borneo, UCL coll. Plate 7.21 Scale bars: Figs 1-8  = 1mm. Fig. 1. Eulepidina favosa Vaughan, Early Miocene, Juana Diaz formation, La Rambla, Puerto Rico, NHM P47332. Fig. 2. Eulepidina badjirraensis Crespin, Early Miocene (Burdigalian, Te5), Mosque Quarry, Loc. 205, Borneo, UCL coll. Fig. 3. Eulepidina parkinsonia BouDagher- Fadel and Price, holotype figured by BouDagher- Fadel and Price (2010a), Miocene (Aquitanian), Hodges Bay, Antigua, West Indies, NHM P22753. Fig. 4. Eulepidina undosa laramblaensis Eames et al., holotype, early Miocene, Juana Diaz Formation, La Rambla, Puerto Rico. NHM P47353 Fig. 5-7 . Eulepidina formosa (Schlumberger), Miocene (Aquitanian), Java, UCL coll. Fig. 8. Eulepidina eodilatata Douvillé, Oligocene (Rupelian), lower part of the Mesolouri section, marly facies, Greece, SFN coll. 2001z0153/ 0006. References Adams, C.G., 1959. Geological distribution of Discospirina (foraminifera) and occurrence of D. italica in the Miocene of Cyprus. Palaeontology 1, 364– 368. Adams, C.G., 1965. The foraminifera and stratigraphy of the Melinau Limestone, Sarawak, and its importance in Tertiary correlation. Quart. J. Geol. Soc. Lond. 121, 283– 338. Adams, C.G., 1967. Tertiary Foraminifera in the Tethyan, American, and Indo-P acific Provinces. In: Adams, C.G., Ager, D. (Eds), Aspects of Tethyan Biogeography, Systematics Association Publication, special publication. London, Vol. 7, pp. 195–2 17. Adams, C.G., 1968. A revision of the foraminiferal genus Austrotrillina Parr. British Museum (Natural History), Bulletin (Geology) 16, 73– 97. Adams, C.G., 1970. A reconsideration of the East Indian letter classification of the Tertiary. Bull. Brit. Mus. (Nat. Hist.) Geol. 19, 87– 137. Adams, C.G., 1973. Some tertiary foraminifera. In: Hallam, A. (Ed.), Atlas of Paleobiogeography. Elsevier, pp. 453– 468. Adams, C.G., 1984. Neogene larger foraminifera, evolutionary and geological events in the context of datum planes. In: Ikebe, I., Tsuchi, R. (Eds), Pacific Neogene Datum Planes. University of Tokyo Press, Tokyo, pp. 47–6 8. Adams, C.G., 1987. On the classification of the Lepidocyclinidae (foraminifera) with redescriptions of the unrelated Paleocene genera Actinosiphon and Orbitosiphon. Micropaleontology 33, 289– 317. Adams, C.G., Gentry, A.W., Whybrow, P.J., 1983. Dating the terminal Tethyan event. In:  Meulenkamp, J.E. (Ed.), Reconstruction of Marine Paleoenvironments. Utrecht Micropaleontological Bulletins,, 30, 273– 298. Adams, C.G., Racey, A., 1992. The occurrence and palaeobiogeographical significance of the foraminiferid Yaberinella from the Eocene of Oman. Paleontology 35, 237–2 45. Adams, R.M., Rosenzweig, C., Peart, R.M., Ritchie, J.T., McCarl, B.A., Glyer, J.D., Curry, R.B., Jones, J.W., Boote, K.J., Allen, L.H., Jr., 1990. Global climate change and U.S. agriculture. Nature 345, 219– 224. Advocaat, E. L., Hall, R., White, L. T., Watkinson, I. M., Rudyawan, A., Boudagher- Fadel, M. K., 2017. Miocene to recent extension in NW Sulawesi, Indonesia. Journal of Asian Earth Sciences. doi:10.1016/j .jseaes.2017.07.023 Aguirre, J., Riding, R., 2005. Dasycladalean algal biodiversity compared with global variations in temperature and sea level over the past 350 Myr. Palaios 20, 581– 588. Aitchison, J.C., Ali, J.R., Davis, A.M., 2007. When and where did India and Asia collide? J.  Geophys. Res.-S ol. Ea. 112(Article number), p. B05423. Albrich, S., Frijia, G., Parente, M., Caus, M., 2014. The evolution of the earliest representatives of the genus Orbitoides: Implications for Upper Cretaceous biostratigraphy. Cretaceous Research 51, 22–3 4. Alegret, L., Arenillas, I., Arz, J.A., Diaz, C., Grajales-N ishimura, J.M., Melendez, A., Molina, E., Rojas, R., Soria, A.R., 2005. Cretaceous–P aleogene boundary deposits at Loma Capiro, Central Cuba, evidence for the Chicxulub impact. Geology 33, 721– 724. Alexander, S.P., 1985. The cytology of certain benthonic foraminifera in relation to test structure and function. PhD thesis, University of Wales. Al-H abeeb, K.H., 1977. Foraminifera in the carboniferous limestones in South Wales. PhD thesis, University College of Swansea, University of Wales. Allemann, F., Schroeder, R., 1972. Spiroconulus perconigi n. sp. a new Middle Jurassic foraminifer of Oman and Spain. Revista Espan˜ ola de Micropaleontologia 30, 199– 209. Allmon, W.D., 2001. Nutrients, temperature, disturbance, and evolution: A model for the late Cenozoic marine record of the western Atlantic. Paleogeography, Paleoclimatology, Paleoecology, 166, 9– 26. Almazán- Vázquez, E., Buitrón- Sánchez, B.  E., Vachard, D., Mendoza-M adera, C., Gómezespinosa, C., 2007. The late Atokan (Moscovian, Pennsylvanian) chaetetid accumulations of Sierra Agua Verde, Sonora (NW Mexico), compo- sition, facies and palaeoenvironmental signals. In: Álvaro, J. J., Aretz, M., BoulvaIn: F., Munnecke, A., Vachard, D., and VennIn: E. (eds) Palaeozoic Reefs and Bioaccumulations, Climatic and Evolutionary Controls. Geological Society, London, Special Publications, 275, 189–2 00. Altiner, D., Groves, J. R., O¨zkan- Altiner, S., 2005. Calcareous foraminiferal recovery from the end- Permian mass extinc- tion, southern Turkey. Abstracts, Albertiana, International Symposium on Triassic Chronostratigraphy and Biotic Recovery, Chaohu, China, 23–2 5, 14–1 7. 640 References Altiner D., and Savini, R., 1997. New species of Syzrania from the Amazonas and Solimões basins (north Brazil), remarks on the generic and suprageneric position of syzraniid foraminifers. Revue de Paléobiologie, 16, 7– 20. Altiner, D., Septfontaine, M., 1979. Micropaléontologie, stratigraphie et environnement de deposition d’une série jurassique à facies de plate- forme de la région de Punarbaşi (Taurus oriental, Turquie). Revue de Micropaléontologie, 22, 3– 18. Altiner, R.D., Rettori, R. 2003. Origin and early evolutionary radiation of the Order Lagenida (Foraminifera). Journal of Paleontology, 77, 831– 843. Amodiao, S., 2006. Foraminifera diversity changes and paleoenvironmental analysis, the Lower Cretaceous shallow- water carbonates of San Lorenzello, Campanian Apennines, southern Italy. Facies, 52, 53– 67. An W., Hu, X., Garzanti, E., BouDagher- Fadel, M., Wang, J., Sun, G., 2014. Xigaze forearc basin revisited (South Tibet):  Provenance changes and origin of the Xigaze Ophiolite. Geological Society of American Bulletin, 126. 1595– 1613 An, W., Hu, X., Garzanti, E., BouDagher- Fadel, M. K., Wang, J., Sun, G., 2015. Erratum to Xigaze forearc basin revis- ited (South Tibet): Provenance changes and origin of the Xigaze Ophiolite, [Geological Society of America, 126, 11/ 12, 1595–1 613, 1984] 10.1130/B 31020.1. Bulletin of the Geological Society of America, 127, 409. doi:10.1130/ B25514.1. Anderson, O.R., Spindler, M., Bé, A.W.H., Hemleben, Ch., 1979. Trophic activity of planktonic foraminifera. J.  Mar. Biolog. Assoc. UK 59, 791– 799. Angiolini, L., Gaetani, M., Muttoni, G., Stephenson, M.H., Zanchi, A., 2006. Tethyan oceanic currents and climate gradi- ents 300 Ma ago. Geology 35, 1071– 1074. Archibald, J.M., Longet, D., Pawlowski, J., Keeling, P., 2003. A novel polyubiquitin structure in cercozoa and foraminifera, evidence for a new eukaryotic supergroup. Molecular Biology and Evolution, 20, 62– 66. Arnaud-V anneau, A., 1975. Reflexion sur le mode de vie de certains Orbitolinides (Foraminifères) barremo-a ptiens de l’Urgonien du Vercors. Compte Rendu des Seances de la Société de Physique et d’Histoire Naturelle de Genève, 10, 126–1 30. Arnaud- Vanneau, A., 1980. Micropaléontologie, paléoécologie et sédimentologie d’une plate- forme carbonatée de la marge passive de la Téthys. Géologie alpine, 11, 1– 874. Arnaud-V anneau, A., and Peybernés, B., 1978. Les représentants Ėocrétacés du genre Nautiloculina Mohler, 1938 (Foraminifera, Fam. Lituolidae?) dans les chaines Subalpines septentrionales (Vercors) et les Pyrénées Franco- Espagnoles. Revision de Nautiloculina cretacea Peybernés, 1976 et déscription de Nautiloculina bronnimanni n.  sp.. Geobios, 11, 67–8 1. Arnold, Z.M., 1978. An allogromiid ancestor of the miliolidean foraminifera. Journal of Foraminiferal Research, 8, 83– 96. Azema, J., Chabrier, G., Fourcade, E., and Jaffrezo, M., 1977. Nouvelles données micropaléontologiques, stratigraphiques et paléogéographiqies sur le Portlandien et le Néocomien de Sardaigne. Revue de Micropaléontologie, 20, 125–1 39. Baldauf, S.L., 2003. The deep roots of Eukaryotes. Science 300, 1703– 1706. Bambach, R. K., 2006. Phaenerozoic Biodiversity Mass Extinctions. The Annual Review of Earth and Planetary Science, 34, 117–1 55. Banerjee, A., Boyajian, E., 1997. Selectivity of foraminiferal extinction in the Late Eocene. Paleobiology 23, 347–3 57. Banerjee, A., Yemane, K., Johnson, A., 2000. Foraminiferal Biostratigraphy of Late Oligocene- Miocene Reefal Carbonates in Southwestern Puerto Rico. Micropaleontology, 46 327– 342. Banner, F. T., Lord, A. R., & BouDagher-F adel, M. K., 1999. The Terra Limestones Member (Miocene) of Western Cyprus. Greifswalder Geowissenschaftliche Beiträge, 6, 503– 515. Banner, F. T., Whittaker, J. E., Boudagher-F adel, M. K., & Samuel, A. (1997). Socotraina, a new hauraniid genus from the upper Lias of the Middle East (Foraminifera, Textulariina). Revue de Micropaleontologie, 40, 115–1 23. Banner, F.T., 1970. A synopsis of the Spirocyclinidae. Revista Espanola de Micropaleontologia 2, 243–2 90. Banner, F.T., Finch, E.M., Simmons, M.D., 1990. On Lithocodium Elliott (calcareous algae); its paleobiological and strati- graphical significance. Journal of Micropaleontology 9, 21– 36. Banner, F.T., Highton, J., 1990. On Everticyclammina Redmond (foraminifera), especially E. kelleri (Henson). Journal of Micropaleontology. 9, 1– 14. Banner, F.T., Hodgkinson, R.L., 1991. A revision of the foraminiferal subfamily Heterostegininae. Revista Espanola de Micropaleontologia 23, 101– 140. Banner, F.T., Knight-J ones, E.W., and Wright, J.M., 1994. Protozoa, the marine fauna of the British Isles and North- West Europe. In: Hayward, P.J., and Ryland, J.S. (eds.), Introduction to Arthropods and Protozoans. Oxford Science Publications, 36– 50. Banner, F.T., Pereira, P.G., 1981. Some biserial and triserial agglutinated smaller foraminifera, their wall structure and its significance. Journal of Foraminiferal Research, 11, 85– 117. Banner, F.T., Samuel, A., 1995. Alanlordia, a new genus of acervuline foraminifera from the Neogene of Indonesia. J. Micropalaeontol. 14, 107– 117. Banner, F.T., Simmons, M.D., 1994. Calcareous algae and foraminifera as water depth indicators, an example from the early cretaceous carbonate of northeast Arabia. In:  Simmons, M.D. (Ed.), Micropalaeontology and Hydrocarbon Exploration in the Middle East. Chapman and Hall, London, pp. 243– 252. References 641 Banner, F.T., Simmons, M.D., Whittaker, J.E., 1991. The Mesozoic chrysalidinidae (foraminifera, textulariacea) of the Middle East, the Redmond (Aramco) taxa and their relatives. Bull. Brit. Mus. Nat. Hist. (Geol.) 47, 101– 152. Banner, F.T., Whittaker, J.E., 1991. Redmond’s ‘‘new lituolid foraminifera’’ from the Mesozoic of Saudi Arabia. Micropaleontology 37, 41– 59. Baranova, D.V., Kabanov, P.B., 2003. Facies distribution of fusulinoid genera in the Myachkovian (Upper Carboniferous, upper Moscovian) of southern Moscow region. Rivista Italiana di Paleontologia e Stratigrafia 109, 225–2 39. Barattolo, F., Bigozzi, A., 1996. Dasycladaceans and depositional environments of the Upper Triassic–L iassic carbonate platform of the Gran Sasso. Facies 35, 163– 208. Barker, R.W., Grimsdale, T.F., 1936. A contribution to the phylogeny of the orbitoidal foraminifera, with descriptions of new forms from the Eocene of Mexico. Journal of Paleontology, 10, 231– 247. Bartholdy, J., 2002. The Architecture of Nummulites (Foraminifera) reexamined, PhD thesis, Freie Universita¨ t Berlin, 64, 1–2 2. Basov, V.A., Kuznetsova, K.I., 2000. Dynamics of diversity and evolutionary trends of Jurassic foraminifers. Stratigr. Geol. Correl. 8, 593–6 07. Bassoullet, J.P., Boutakiout, M., 1996. Haurania (Platyhaurania) subcompressa nov. subgen., nov. sp., new imperforate larger foraminifera (Hauraniidae) from lower Toarcian of septentrional Morocco (South Rif Ridges). Rev. Micropaleontol. 39, 27–4 0. Bassoullet, J.P., Boutakiout, M. and Echarfaoui, H., 1999. Deux nouveaux genres, Palaeocyclammina et Ijdranella, fora- minifères (Textulariina) d’un niveau Liasique à Orbitopsella praecursor (Gümbel) du Moyen Atlas (Maroc). Revue de Micropaléontologie, 42, 213– 230. Beavington-P enney, S.J., Racey, A., 2004. Ecology of extant nummulitids and other large benthic foraminifera, applications in palaeoenvironmental analysis. Earth Sci. Rev. 67, 219– 265. Beerling, D.J., Berner, R.A., 2002. Biogeochemical constraints on the Triassic–J urassic boundary carbon cycle event. Global Biogeochem. Cycles 16, 1– 13. Bender, H., Hemleben, C., 1988. Calcitic elements secreted by agglutinated foraminifers grown in laboratory culture. J. Foramin. Res. 18, 42– 45. Benedetti A., Pignatti J., 2013. Conflicting evolutionary and biostratigraphical trends in Nephrolepidina praemarginata (Douvillé, 1908) (Foraminiferida). Historical Biology, 25, 363– 383, ISSN: 0891- 2963, doi: 10.1080/ 08912963.2012.713949 Benton, M.J., 2002. Cope’s rule. In:  Pagel, M.  (Ed.), Encyclopedia of Evolution. Oxford University Press, New  York, pp. 209– 210. Benton, M.J., Twitchett, R.J., 2003. How to kill (almost) all life, the end-P ermian extinction event. Trends Ecol. Evol. 18, 358–3 65. Berggren, W.A., 1972. Cenozoic biostratigraphy and paleobiogeography of the North Atlantic. In:  Laughton, A.S., McKenzie, D.P., Sclater, J.G. (Eds), Initial Reports of the Deep Sea Drilling Project, Volume X/ I, Washington, DC: U.S. Govt. Printing Office, pp. 965–1 000. Berggren, W.A., Pearson, P. N., 2005, A revised tropical to subtropical Paleogene planktonic foraminiferal zonation: Journal of Eoraminiferal Research, 35, 279– 298. Berggren, W.A., Prothero, D.R., 1992. Eocene–O ligocene climatic and biotic evolution: An overview. In: Prothero, D.R., Berggren, W.A. (Eds), Eocene– Oligocene Climatic and Biotic Evolution. Princeton University Press, Princeton, pp. 1– 28. Bergquist, H.R., 1971. Biogeographical review of Cretaceous foraminifera of the Western Hemisphere. North American Paleontological Convention, Chicago, 1969, Proceedings, pp. 1565– 1609. Berner, R.A., 2002. Examination of hypotheses for the Permo–T riassic boundary extinction by carbon cycle modelling. Proc. Natl. Acad. Sci. USA 99, 4172–4 177. Berner, R.A., Beerling, D.J., Dudley, R., Robinson, J.M., Wildman, R.A., Jr., 2003. Phanerozoic atmospheric oxygen. Earth Planet. Sci. Lett. 31, 105– 134. Betzler, C., 1997. Ecological controls on geometries of carbonate platforms: Miocene/ Pliocene shallow-w ater microfaunas and carbonate biofacies from the Queensland Plateau (NE Australia). Facies 37, 147– 166. Blackburn, T.J., Olsen, P.E., Bowring, S.A., McLean, N.M., Kent, D.V., Puffer, J., McHone, G., Rasbury, E.T., and Et-Touhami, M., 2013, Zircon U-Pb geochronology links the end-Triassic extinction with the Central Atlantic Magmatic Province: Science, 340, 941–945, doi:10.1126/science.1234204. Bonis, N.R., Kürschner, W.M., Krystyn, L., 2009. A detailed palynological study of the Triassic–Jurassic transition in key sec- tions of the Eiberg basin (Northern Calcareous Alps, Austria). Review of Palaeobotany and Palynology 156, 376–400. Bonnefous, J., Bismuth, H., 1982. Les facies carbonates de plate-f orme de l’Éocene moyen et supérieur dans l’offshore tunisien nord- oriental et en mer pélagienne, implications paléogéographiques et analyse micropaléontologique. Bulletin des Centres de Recherches Exploration- Production Elf-A quitaine 6, 337– 403. Boomer, I., Lord, A., Crasquin, S., 2008. The extinction of the Metacopina (Osatracoda). Senckenbergiana lethaea 88, in press. Bornemann, A., Norris, R., Friedrich, O., Beckmann, B., Schouten, S., Sinninghe Damste´, J., Vogel, J., Hofmann, P., Wagner, T., 2008. Isotopic evidence for glaciation during the cretaceous supergreen- house. Science 319, 189–1 92. 642 References Bosellini, A., Broglio Loriga, C., 1971. I ‘‘Calcari Grigi’’ di Rotzo (Giurassico inferiore, Altopiano di Asiago) e loro inquadra- mento nella paleogeografia e nella evoluzione tettonico-s edimentaria delle Prealpi venete. Annali del Universitario Ferrara 5, 1–6 1. Bosellini, F.R., Papazzoni, C.A., 2003. Palaeoecological significance of coral- encrusting foraminiferan associations: A case- study from the Upper Eocene of Northern Italy. Acta Palaeontol. Pol. 48, 279– 292. Bosellini, F.R., Russo, A., Vescogni, A., 2002. The Messinian Reef Complex of the Salento Peninsula (Southern Italy): Stratigraphy, facies and paleoenvironmental interpretation. Facies 47, 91– 112. Bosence, D., Procter, E., Aurell, M., Kahla, A. B., BouDagher-F adel, M. K., & et al. (2009). A tectonic signal in high- frequency, peritidal carbonate cycles? a regional analysis of liassic platforms from western Tethys. Journal of Sedimentary Research, 79 (6), 389– 415. Boslough, M.B., Chael, E.P., Trucano, T.G., Crawford, D.A., Campbell, D.L., 1996. Axial focusing of impact energy in the Earth’s interior, a possible link to flood basalts and hotspots. In: Ryder, G., Fastovsky, D., Gartner, S. (Eds), The Cretaceous– Tertiary Event and Other Catastrophes in Earth History. Geological Society of America. special paper, 307, 541– 550. BouDagher- Fadel, M.K., 2000. Benthic foraminifera of the Jurassic- Early Cretaceous of Tethys. International Workshop on North African Micropaleontology for Petroleum Exploration 1, 27– 28. BouDagher- Fadel, M.K., 2001. The tax- onomy and evolution of the foraminiferal genus Buccicrenata Loeblich and Tappan. Micropaleontology 47, 168– 172. BouDagher- Fadel, M. K., 2001. The taxonomy and evolution of the foraminiferal genus Buccicrenata Loeblich and Tappan. Micropalaeontology, 47 (2), 168- 172. BouDagher- Fadel, M.K., 2002. The stratigraphical relationship between planktonic and larger benthic foraminifera in Middle Miocene to Lower Pliocene carbonate facies of Sulawesi, Indonesia. Micropaleontology 48, 153–1 76. BouDagher-F adel, M. K., 2008. Evolution and Geological Significance of Larger Benthic Foraminifera. Developments in Palaeontology and Stratigraphy, 21. ELSEVIER, 1– 544. BouDagher- Fadel, M.K., 2013. Diagnostic First and Last Occurrences of Mesozoic and Cenozoic Planktonic Foraminifera ((Professional Papers Series, pp. 1– 4).). London, UK: UCL Office of the Vice-P rovost Research. Boudagher- Fadel, M.K., 2015. Biostratigraphic and Geological Significance of Planktonic Foraminifera (Updated 2nd Edition). London: UCL Press. doi:10.14324/ 111.9781910634257 BouDagher-F adel, M. K., 2016. Genotypes of Septatrocholina and Alzonorbitopsella, two new Jurassic foraminifera: subse- quent designations. Micropaleontology, 62, 87– 89. BouDagher-F adel, M.  K., Banner, F.  T., 1997. The revision of some genus- group names in Tethyan Lepidocyclininae. Paleopelagos, 7, 3– 16. BouDagher-F adel, M.K., Banner, F.T., 1999. Revision of the stratigraphic significance of the Oligocene– Miocene ‘‘Letter- Stages’’. Rev. Micropaleontol. 42, 93– 97. BouDagher- Fadel, M.K., Banner, F.T., Whittaker, J.E. 1997. Early Evolutionary History of Planktonic Foraminifera, British Micropalaeontological Society Publication Series, Chapman and Hall, 269p. BouDagher-F adel, M. K., Bosence, D., 2007. Early Jurassic benthic foraminiferal diversification and biozones in shallow- marine carbonates of western Tethys. Senckenbergiana lethaea, 87, 1– 39. Boudagher- Fadel, M. K., Hu, X., Price, G. D., Sun, G., Wang, J. - . G., BouDagher-F adel, M.  K., Lokier, S., 2005. Significant Miocene larger foraminifera from south central Java. Revue de Micropaléontologie, 24 (1), 291– 309. BouDagher- Fadel, M.K., Lord, A. R., 2000. The evolution of Lepidocyclina (L.) isolepidinoides, L. (Nephrolepidina) nephro- lepidinoides, L. (N.) brouweri in the Late Oligocene- Miocene of the Far East. Journal of Foraminiferal Research, 30, 71– 76. BouDagher- Fadel, M.K., Lord, A.R., 2002. Larger foraminifera of the Jurassic Western Neotethys Ocean. Archaeology and History in Lebanon, 15, 87– 94. BouDagher-F adel, M.K., Lord, A.R., Banner, F.T., 2000. Some Miogypsinidae (foraminiferida) in the Miocene of Borneo and nearby countries. Revue de Paléobiology 19, 137– 156. BouDagher- Fadel, M.  K., Lord, A.  R., 2006. Illusory stratigraphy decoded by Oligocene- Miocene autochthonous and allochthonous foraminifera in the Terra Member, Pakhna Formation (Cyprus). Stratigraphy, 3, 217– 226. BouDagher- Fadel, M.K., Noujaim Clark, G., 2006. Stratigraphy, paleoenvironment and paleogeography of Maritime Lebanon: a key to Eastern Mediterranean Cenozoic history. Stratigraphy, 3, 81– 118. BouDagher- Fadel, M.K., Price, G. D., 2009. Loftusia persica: an Eocene Lazarus occurrence?. Micropaleontology, 55, 75–8 5. BouDagher- Fadel, M.K., Price, G. D., 2010a. Evolution and paleogeographic distribution of the lepidocyclinids. Journal of Foraminiferal Research, 40, 79– 108. BouDagher- Fadel, M.K., Price, G. D., 2010b. American miogypsinidae: An analysis of their phylogeny and biostratigraphy. Micropaleontology, 56, 567– 586. BouDagher- Fadel, M.K., Price, G. D., Koutsoukos, E. A. M., 2010c. Foraminiferal biostratigraphy and paleoenvironments of the Oligocene- Miocene carbonate succession in Campos Basin, southeastern Brazil. Stratigraphy, 7, 283- 299. BouDagher-F adel, M.K., Price, G. D., 2013. The phylogenetic and palaeogeographic evolution of the miogypsinid larger benthic foraminifera. Journal of the Geological Society, 170, 185– 208. doi:10.1144/ jgs2011- 149 BouDagher- Fadel, M.K., Price, G. D., 2014. The phylogenetic and palaeogeographic evolution of the nummulitoid larger benthic foraminifera. Micropaleontology, 60, 483– 508. References 643 Boudagher-Fadel, M. K., Hu, X., Price, G. D., Sun, G., Wang, J. -. G., An, W., 2017. Foraminiferal biostratigraphy and palaeoenvironmental analysis of the mid-Cretaceous limestones in the southern Tibetan plateau. Journal of Foraminiferal Research, 47, 188-207. doi:10.2113/gsjfr.47.2.188 BouDagher- Fadel, M.K, Price, G.  D., 2017. The paleogeographic evolution of the orthophragminids of the Paleogene. Journal of Foraminiferal Research, 47, 337–357. BouDagher-F adel, M. K., Price, G. D., Hu, X., Li, J., 2015. Late Cretaceous to early Paleogene foraminiferal biozones in the Tibetan Himalayas, and a pan- Tethyan foraminiferal correlation scheme. Stratigraphy, 12, 67– 91. Boudagher- Fadel, M., Rose, E. P. F., Bosence, D. W. J., Lord, A. R., 2001. Lower Jurassic foraminifera and calcified micro- flora from Gibraltar, western Mediterranean. Palaeontology, 44 (4), 601– 621. BouDagher- Fadel, M.K., Wilson, M., 2000. A  revision of some larger Foraminifera of the Miocene of South- East Kalimantan. Micropaleontology, 42 (2), 153– 166. Boussac., J., 1906. Developement et Morphologie de quelques Foraminifères de Priabona. Bulletin Société Géologique de France 6, 88–9 7. Bowser, S. S. and Travis, J. L., 2002. Reticulopodia: structural and behavioral basis for the suprageneric placement of granu- loreticulosan protists. J. Foraminiferal Res., 32, 440–447. Braga, J. C., García-G omez, R., Jiménez, A. P., and Rivas, P., 1981. Correlaciones en el Lías de las Cordilleras Béticas. In: Almela Samper, A., and Meléndez Meléndez, B., (eds), Curso de conferencias sobre el Programa Internacional de Correlación Geológica, Part 2. Real Academia de Ciencias Exactas, Físicas y Naturales, Madrid, 161– 181. Brasier, M.D., 1975. An outline history of seagrass communities. Palaeontology 18, 681– 702. Breitfeld, H.  T., Hall, R., Galin, T., Forster, M.  A., & BouDagher-F adel, M.  K., 2016. A  Triassic to Cretaceous Sundaland–P acific subduction margin in West Sarawak, Borneo. Tectonophysics, 694, 35– 56. doi:10.1016/ j.tecto.2016.11.034. Brönnimann, P., 1951. Remarks on the embryonic chambers of Upper Eocene Asterocyclinidae of Trinidad, B.W.I. Eclogae geologicae Helvetiae 44, 474– 486. Brönnimann, P., 1954a. Upper cretaceous orbitoidal foraminifera from Cuba. I.  Sulcorbitoides n.  gen. Contribution Cushman Foundation for Foraminiferal Research 5, 55– 61. Brönnimann, P., 1954b. Upper cretaceous orbitoidal foraminifera from Cuba II. Vaughanina Palmer. Contrib. Cushman Foundation for Foraminiferal Research 5, 91– 105. Brönnimann, P., 1958. New pseudorbitoidadae from the Upper Cretaceous of Cuba, with remarks on encrusting foramin- ifera. Micropaleontology 4, 165– 186. Brönnimann, P., 1966. Pseudotextulariella courtionensis n. sp., from the Valanginian of well Courtion 1, Courtion, Canton of Fribourg, Switzerland. Archives Sciences, Geneve 19, 265– 278. Brun, L., Butterlin, J., and Montell, L., 1982. Decourverte de Lépidocyclines (Foraminières) d’âge Eocène dans le Golfe de Guinée. Implications Paléobiogéographiques. Cahiers de Micropaléontologie, 2, 1– 109. Burki F, Kaplan M, Tikhonenkov D V, Zlatogursky V, Minh BQ, Radaykina L V, Smirnov A, Mylnikov P, Keeling PJ, Keeling PJ, 2016. Untangling the early diversification of eukaryotes: a phylogenomic study of the evolutionary origins of Centrohelida, Haptophyta and Cryptista, 1– 10. Bursch, J.G., 1947. Mikropalaontologische Untersuchungen des Tertiars von Gross Kei (Molukken). Schweizerische Palaeontologische Abhandlungen 65, 1– 56. Butterlin, J., 1987. Origine et e´ volution des Lepidocyclines de la re´ gion des Caraibes. Comparisons et relations avec les Lepidocyclines des autres re´ gions du monde. Rev. Micropaleontol. 29, 203– 219. Buxton, M.W.N., Pedley, H.M., 1989. A standardized model for Tethyan Tertiary cabonate ramps. J. Geol. Soc. Lond. 246, 746–7 48. Cahuzac, B., 1984. Les faunes de Miogypsinidae d’Aquitaine meridionale (France), International Symposium Benthic Foraminifera (Pau, April 1983), 1984, pp. 117– 129. Cahuzac, B., Poignant, A., 1991. Morphologie des éspèces de Pararotalia et de Miogypsinoides (Foraminiferida) dans l’Oligocène d’Aquitaine me´ ridionale. Geobios, Mem. Sp. 13, 69– 78. Caplan, M.L., Bustin, R.M., 1999. Devonian– Carboniferous Hangenberg mass extinction event, widespread organic- rich mudrock and anoxia: causes and consequences. Paleogeography, Paleoclimatology. Paleoecology. 148, 187–2 07. Carpenter, W.B., 1850. On the microscopic structure of Nummulina, Orbitolites and Orbitoides. Quart. J. Geol. Soc. Lond. 6, 181–2 36. Carpenter, W.B., Parker, W.K., Jones, T.R., 1862. Introduction to the study of the foraminifera. Ray Soc. Lond. 319. Castellarin, A., 1972. Evoluzione tettonica sinsedimentaria del limite tra ‘‘Piattaforma veneta’’ e ‘‘Bacino lombardo’’ a nord di Riva del Garda. Giornale di Geologia 38, 11– 212. Caudri, C.M.B., 1972. Systematics of the American Discocyclinas. Eclogae geologicae Helvetiae 65, 211– 219. Caudri, C.M.B., 1974. The larger foraminifera of Punta Mosquito, Margarita Island, Venezuela. Verhandlungen der Naturforschenden Gesellschaft in Basel 84, 293– 318. Caudri, C.M.B., 1996. The larger foraminifera of Trinidad (West Indies). Eclogae geologicae Helvetiae 89, 1137–1 309. Caus, E., 1988. Upper Cretaceous larger foraminifera, paleoecological distribution. Revue de Paleobiologie, Switzerland, volume special, 2, 417–4 19. Cavalier-S mith, T., 1998. A revised six-k ingdom system of life. Biol. Rev. Camb. Philos. Soc. 73, 203– 266. 644 References Cavalier- Smith, T., 2002. The phagotrophic origin of eukaryotes and phylogenetic classification of protozoa. Int. J. Syst. Evol. Microbiol. 52, 297– 354. Cavalier- Smith, T., Chao, E.E., 1997. Sarcomonad ribosomal RNA sequences, rhizopod phylogeny, and the origin of eug- lyphid amoebae. Archiv. Fuer. Protistenkunde. 147, 227– 236. Chapelle G., Peck, L.S., 1999. Polar gigantism dictated by oxygen availability. Nature, 399, 114–115. Chaproniere, G.C.H., 1975. Palaeoecology of Oligo– Miocene larger foraminiferida, Australia. Alcheringa 1, 37–5 8. Chaproniere, G.C.H., 1980. Biometrical studies of early Neogene larger foraminiferida from Australia and New Zealand. Alcheringa 4, 153–1 81. Chaproniere, G.  C. H., 1983. Tertiary larger foraminiferids from the northwestern margin of the Queensland Plateau, Australia; in Palaeontological papers 1983. Bureau of Mineral Resources, Australia, Bulletin 217, 31– 57. Chaproniere, G.C.H., 1984. The Neogene larger foraminiferal sequence in the Australian and New Zealand regions, and its relevance to the East Indies letter stage classification. Palaeogeog. Palaeoclimatol. Palaeoecol. 46, 25– 35. Chaproniere, G.C.H., 1991. Discospirina (miliolina, foraminiferida) from the Coral Sea off Northeastern Australia: A new record. Journal of Paleontology, 65, 332– 334. Chen, Z.Q., Benton, M.J. 2012. The timing and pattern of biotic recovery following the end- Permian mass extinction. Nature Geoscience 5: 375–3 83. Chen, D., Tucker, M.E., 2003. The Frasnian– Famennian mass extinction, insights from high- resolution sequence stratigra- phy and cyclostratigraphy in South China. Palaeogeog. Palaeoclimatol. Palaeoecol. 193, 87– 111. Cherchi, A., Schroeder, R., 1982. Dictyoconus algerianus n.  sp., grand Foraminifère de l’Aptien supeérieur du plaque Africaine (marge septentrionale). Comptes Rendus de l’Académie des Sciences, Paris, Ser. II 295, 77– 82. Cherchi, A., Schroeder, R., Zhang, B.-G ., 1984. Cyclorbitopsella tibetica n.gen., n.sp., a lituolacean foraminifer from the Lias of southern Tibet. In: Oertli, H.J. (Ed.), Benthos ‘83, Second International Symposium on Benthic Foraminifera. Pau, Elf Aquitaine, Esso Rep and Total CFP, Pau and Bordeaux, pp. 159– 165. Chernysheva, N.E., 1948. Archaedisus i blizkikh k nemu formakh iz nizhnego Karbona SSSR. Trudy Instituta Geologischeskikh Nauk, Akademiya Nauk SSSR 62, 150– 158. Chiocchini, M., Farinacci, A., Mancinelli, A., Molinari, V., Potetti, M., 1994. Biostratigrafia a foraminiferi, dasicladali e calpionelle delle successioni carbonatiche mesozoiche dell’Appennino centrale (Italia). In:  Mancinelli, A., (Ed.), Biostratigrafia dell’Italia centrale. Studi Geologici Camerti, Volume Speciale 1994 (A), 9– 129. Chiocchini, M., Mancinelli, A., Romano, A., 1984. Stratigraphic distribution of benthic foraminifera in the Aptian, Albian, and Cenomanian carbonate sequences of the Aurunci and Ausoni Mountains (Southern Lazio, Italy). Benthos 83, 167–1 81. Ciry, R., 1964. À propos de Meandropsina larrazeti Mun.- Chalm., generotype d’un genre nouveau Larrazetia Ciry. Rev. Micropaleontology, 6, 185– 195. Clauzon, G., Suc, J.P., Gautier, F., Berger, A., Loutre, M.F., 1996. Alternate interpretation of the Messinian salinity crisis, controversy resoved? Geology 24, 363– 366. Cloud, P., 1976. Beginnings of biospheric evolution and their biogeochemical consequences. Paleobiology 2, 351–3 87. Cohen, K.M., Finney, S.C., Gibbard, P.L., Fan, J.-X ., 2013, The ICS International Chronostratigraphic Chart, Episodes, 36, 199– 204. Cole, W.S., 1941. Stratigraphic and paleontologic studies of wells in Florida – no 1. Fla Geol. Surv. Bull. 19, p. 91. Cole, W.S., 1957. Variation in American Oligocene species of epidocyclina. Bull. Am. Paleonotol. 38, 31–5 1. Cole, W.S., 1963. Tertiary larger foraminifera from Guam. US Geol. Surv. Prof. Pap. 403E, p. 28. Cole, W.S., 1964. Orbitoididae; Discocyclinidae; Lepidocyclinidae. In: Loeblich, A.R., Jr., Tappan, H. (Eds), Sarcodina, Chiefly ‘‘Thecamoebians’’ and Foraminiferida, Treatise on Invertebrate Paleontology (R.C. Moore, Ed.), Geological Society of America and University of Kansas Press, Part: C. Protista 2, pp. C710– C724. Cole, W.S., Bermudez, P.J., 1944. New foraminiferal genera from the Cuban Middle Eocene. Bull. Amer. Paleont. 28, (113), 333– 334. Conway Morris, S., 2003. Life’s Solution:  Inevitable Humans in a Lonely Universe. Cambridge University Press, Cambridge, 464p. Cope, E.D., 1896. The Primary Factors of Organic Evolution. Open Court Publishing Company, Chicago, 547p. Copper, P., 1989. Enigmas in phanerozoic reef development. Memoire association Australas. Palaeontology 8, 371–3 85. Copper, P., 2002. Reef development at the Frasnian/F amennian mass extinction boundary. Paleogeography, Paleoclimatology, Paleoecology, 28, 1– 39. Cosijn, A.J., 1938. Statistical Studies on the Phylogeny of Some foraminifera: Cycloclypeus and Lepidocyclina from Spain, Globorotalia from the East- Indies. Doctoral thesis, University of Leiden, pp. 1– 70. Ćosović V., Drobne, K., Moro, A., 2004. Paleoenvironmental model for Eocene foraminiferal limestones of the Adriatic carbonate platform (Istrian Peninsula). Facies 50, 61– 75. Cossey, P.J, Mundy, D.J.C., 1990. Tetrataxis, a loosely attached limpet- like foraminifer from the Upper Paleozoic, Lethaia, 23, 311– 322. Courgeon, S., Jorry, S.  J., Camoin, G.  F., BouDagher- Fadel, M.  K., Jouet, G., Révillon, S., Bachèlery P, Pelleter E, Borgomano J, Poli E, Droxler, A.  W., 2016. Growth and demise of Cenozoic isolated carbonate platforms:  new insights from the Mozambique Channel seamounts (SW Indian Ocean). Marine Geology, 380, 90– 105. doi:10.1016/ j.margeo.2016.07.006. References 645 Courgeon S, Jorry SJ, Jouet G, Camoin G, BouDagher-F adel MK, Bachelery P, Caline B, Boichard R, Revillon S, Thomas Y, Thereau E, Guerin C., 2017. Impact of tectonic and volcanism on the Neogene evolution of isolated carbonate plat- forms (SW Indian Ocean). Sedimentary Geology, 355, 114– 131. doi:10.1016/ j.sedgeo.2017.04.008. Courtillot, V.E., Renne, P.R., 2003. On the ages of flood basalt events. C. R. Geosci. 335, 113– 140. Cowen, R., 1983. Algal symbiosis and its recognition in the fossil record. In: Teveszm, J.S., McCallp, L. (Eds), Biotic Interactions in Recent and Fossil Benthic Communities. Plenum Press, New York, pp. 431– 478. Cózar, P., Rodríguez, 2003. The palaeoecological distribution of the endothyroids (foraminifera) in the Guadiato area (SW Spain, Mississippian). Palaeogeography, Palaeoclimatology, Palaeoecology, 201, 1–1 9. Cózar, P., Vachard D., 2001. Dainellinae, Subfam.nov. (early Carboniferous Foraminiferida), review and new taxa. Geobios, 34, 5: 505– 526. Cramer, B.S., Kent, D.V., 2005. Bolide summer: The Paleocene/ Eocene thermal maximum as a response to an extraterrestrial trigger. Paleogeography, Paleoclimatology, Paleoecology, 224, 144–1 66. Cummings, R.H., 1956. Revision of upper Paleozoic textulariid foraminifera. Micropaleontology 2, 201– 242. Cushman, J.A., Jarvis, P.W., 1931. Some new Eocene Foraminifera from Jamaica. Contrib. Cushman Found. Foramin. Res. 7, 75– 78. Cuvillier, J., Foury, G., Pignatti Morano, A.G., 1968. Foraminifères nouveaux du Jurassique supérieur du Val Cellina (Frioul Occidental, Italie). Geologica Romana 7, 141– 156. D’Archiac, A., 1850. Histoire des progrés de la Géologie de 1834 à 1849. Formation nummulitique de l’Espagne 3, p. 304. Dain, I. G. 1958. New genera and species of foraminifera. Trudy Vsesoyuznogo Nauchno-I ssled Geologorazved Institution, Gostoptekhizdat, 9, 4– 81. Davies, A. M., 1935. Tertiary faunas: A textbook for oilfield paleontologists and students of geology. The composition of Tertiary faunas. Allen and Unwin, London, Vol. 1, 1406p. Davydov, V.I., 1984. On the problem of origin of Schwagerins. Paleontol. J. 3– 16 (In Russian). Davydov, V.I., 2011. Taxonomy, nomenclature and evolution of the early schubertellids (Fusulinida, Foraminifera). Acta Palaeontol. Pol. 56, 181–1 94. Davydov, V.I., 2013. Climate fluctuations within the western Pangean tropical shelves - the Pennsylvanian/ Permian record from benthic foraminifera. In: Lucas, S.G., DiMichele, W.A., Barrick, J.E., Schneider, J.W., Spielmann, J.A. (Eds.), The Carboniferous– Permian Transition. Bulletin New Mexico Museum of Natural History and Science, 73–7 8. Davydov, V., 2014. Warm Water Benthic Foraminifera Document the Pennsylvanian-p ermian Warming and Cooling Events – The Record from the Western Pangea Tropical Shelves”. Palaeogeography, Palaeoclimatology, Palaeoecology, 414, 284– 295. Davydov, V.I., Arefifard, S., 2007. Permian fusulinid fauna of Gondwanan affinity from Kalmard Region, East- Central Iran and its significance for the tectonics and paleogeography. Paleontol. Electron, 40 (http://p alaeoelectronica.org/2 007_ 2/0 0124/ index.html). Davydov, V.I., Wardlaw, B.R., Gradstein, F.M., 2004. The carboniferous period. In: Gradstein, F.M., Ogg, J.G., Smith, A.G. (Eds), A Geologic Time Scale. Cambridge University Press, Cambridge, pp. 222–2 48. De Bock, J.F., 1976. Studies on some Miogypsinoides – Miogypsina s.s. associations with special reference to morphological features. Scripta geologica 36, 1– 137. De Castro, P., 1971. Osservazioni su Raadshoovenia van den Bold, e i suoi col nuovo genere Scandonea (Foraminiferida, Miliolacea). Bollettino della Societa dei Naturalisti in Napoli 80, 161– 235. De Gracianski, P.-C ., Hardenbol, J., Jacquin, Th., Vail, P.R., 1998. Mesozoic–C enozoic Sequence Stratigraphy of European Basins, Society of Economic Paleontologists and Mineralogists Special Publication, SEPM, Tulsa, OK, Vol. 60, 786pp. De Mulder, E.F.J., 1975. Microfauna and sedimentary-t ectonic history of the Oligo–M iocene of the Ionian Islands and western Epirus (Greece). Utrecht Micropaleontological Bulletins, 13, p. 147. Deckart, K., Féraud, G., Bertrand, H., 1997. Age of Jurassic continental tholeiites of French Guyana, Surinam and Guinea: Implications for the initial opening of the central Atlantic Ocean. Earth Planet. Sci. Lett. 150, 205–2 20. Decrouez, D., Moullade, M., 1974. Orbitolinides nouveaux de l’Albo- Cenomanien de Grece. Arch. Sci. Geneve 27, 75– 92. Deloffre, R., Hamaoui, M., 1969. Biostratigraphie des ‘‘Breches de Soumoulou’’ et description de Pseudobroeckinella soumoulouensis n.  gen., n.  sp., Foraminifère du Cretacé supérieur d’Aquitaine. Bulletin Centre Recherche Pau- SNPA 3, 5– 31. Dettmering, C., Röttger, R., Hohenegger, J., Schmaljohann, R., 1998. The trimorphic life cycle in foraminifera: Observations from cultures allow new evaluation. European Journal of Protistology, 34, 363– 368. Dickens, G. R., 2001. Carbon addition and removal during the Late Palaeocene Thermal Maximum: Basic theory with a preliminary treatment of the isotope record at ODP Site 1051, Blake Nose. In: Kroon, D., Norris, R.D., Klaus, A.  (Eds), Western North Atlantic Palaeogene and Cretaceous Paleoceanography, Geological Society of London (Special Publication), London, Vol. 183, 293– 305. Dilley, F.C., 1973. Cretaceous larger foraminifera. In: Hallam, A. (Ed.), Atlas of Palaeobiogeography. Elsevier, Amsterdam, pp. 403–4 19. Dolenec, T., Pavsic, J., Lojen, S., 2000. Ir anomalies and other elemental markers near the Palaeocene– Eocene boundary in a flysch sequence from the Western Tethys (Slovenia). Terra Nova 12, 199– 204. 646 References Douglass, R.C., 1977. The development of fusulinid biostratigraphy. In: Kauffman, E. G., Hazel, J. E., (Eds), Concepts and Methods of Biostratigraphy. Dowden, Hutchinson, and Ross, Inc., Stroudsbard, Pennsylvania, pp. 463–4 81. Drobne, K., 1974. Les grandes Miliolides des couches paleocenes de la Yougoslavie du Nord- Ouest (Idalina, Fabularia, Lacazina, Periloculina). Razprave IV, razr, SAZU 17, 129–1 84. Drobne, K., 1975: Hottingerina lukasi n. gen., n. sp. (Foraminiferida) du Paleocene moyen provennant du Nord- Ouest de la Yougoslavie. Razprave, 4.razr. SAZU 18, 242– 253. Drobne, K., 1988. Elements structuraux et repartition stratigraphique des grands miliolides de la famille des Fabulariidae. Revue de Palé obiologie, Vol. Special, Benthos ‘86, Vol. 2, pp. 643– 661. Drooger, C.W., 1952. Study of American Miogypsinidae. Doctoral thesis, University of Utrecht, p. 80. Drooger, C.W., 1955. Remarks on cycloclypeus. Proc. Koninklijke Nederlandse Akadamie Wetenschappen, Amsterdam, Ser. B 58, 415–4 33. Drooger, C.W., 1963. Evolutionary trends in Miogypsinidae. In: Evolutionary Trends in Foraminifera. Drooger, C.W., 1983. Environmental gradients and evolutionary events in some larger Foraminifera. Utrecht Micropaleontological Bulletins, 30, 255– 271. Drooger, C.W., 1984. Evolutionary partterns in lineages of orbitoidal foraminifera. Palaeontol., Proc. B 87, 103–1 30. Drooger, C.W., 1993. Radial foraminifera, morphometrics and evolution. Verhandelingen der Koninklijke Nederlandse Akademie van Wetenschappen, Afd. Natuurkunde, Erste Reeks, Amsterdam, deel 41, 241p. Drooger, C.W., Raju, D.S.N., 1973. Protoconch diameter in the Miogypsinidae. Proc. Koninklijke Nederlandse Akadamie Wetenschappen, Amsterdam, Ser. B 76, 206– 216. Drooger, C.W., Roelofsen, J.W., 1982. Cycloclypeus from Ghar Hassan, Malta. Proceedings of Koninklijke Nederlandse Akadamie Wetenschappen, Amsterdam, Ser. B- 85, pp. 203– 218. Drooger, C.W., Rohling, E.J., 1988. Lepidocyclina migration across the Atlantic. Proceedings Koninklijke Nederlandse Akadamie Wetenschappen, Amsterdam, Ser.B, 91, pp. 39– 52. Dullo, W., Moussavian, E., Brachert, T.C., 1990. The foralgal crust facies of the deeper fore reefs in the Red Sea, a deep diving survey by submersible. Geobios 23, 261– 281. Eames, F.E., Banner, F.T., Blow, W.H., Clarke, W.J., 1962. Fundamentals of Mid-T ertiary Stratigraphical Correlation. Cambridge University Press, Cambridge, 163p. Edinger, E.N., Risk, M.J., 1994. Oligocene–M iocene extinction and geographic restriction of Caribbean corals, roles of tur- bidity, temperature, and nutrients. Palaios 9, 576– 598. Eldholm, O., Coffin, M.F., 2000. Large igneous provinces and plate tectonics. In: Richards, M., Gordon, R., Van der Hilst, R. (Eds), The History and Dynamics of Global Plate Motions, American Geophysical Union Geophysical Monograph, Washington, DC, Vol. 121, pp. 309– 326. Eldholm, O., Thomas, E., 1993. Environmental impact of volcanic margin formation. Earth Planet. Sci. Lett. 117, 319–3 29. Elliott, J.M., Logan, A., and Thomas, L.H. 1996. Morphotypes of the foraminifera Homotrema rubrum (Lamarck), distri- bution and relative abundance on reefs in Bermuda. Bulletin of Marine Science, 58, 261–2 76. Erwin, D.H., 1993. The Great Paleozoic Crisis: Life and Death in the Permian. Columbia University Press, 327p. Elsevier, Amsterdam, pp. 315– 349. Erwin, D.H., 1996. Understanding biotic recoveries, extinction, survival, and preservation during the end- Permian mass extinction. In: Jablonski, D., Erwin, D.H., Lipps, J.H. (Eds), Evolutionary Paleobiology. University of Chicago Press, Chicago, pp. 398–4 18. Erwin, D.H., Bowring, S.A., Jin, Y.G., 2002. End-P ermian mass extinctions, a review. In: Koeberl, C., MacLeod, K.G. (Eds), Catastrophic Events and Mass Extinctions, Impacts and Beyond, Geological Society of America, Special Paper, Vol. 356, pp. 363– 383. Erwin, D.H., Droser, M.E., 1993. Elvis taxa. Palaios 8, 623– 624. Fagerstrom, J.A., 1987. The Evolution of Reef Communities. Wiley, New York, 600p. Farinacci, A.  1996. A  new nezzazatid foraminifera in theSalento Upper Campanian limestones (southern Italy). Palaeopelagos, 5, 129–1 37. Farinacci, A, & Yeniay, G. 1994. Tekkeina anatoliensis n.gen. n.sp., a new foraminifer from Susuz Dag, Western Taurus, Turkey. Palaeopelagos, 4, 47– 59. Fedorov, A.V., Dekens, P.S., McCarthy, M., Ravelo, A.C., de Menocal, P.B., Barreiro, M., Pacanowski, R.C., Philander, S.G., 2006. The Pliocene paradox (mechanisms for a permanent El Nin˜ o). Science 312, 1485– 1489. Feldmann, M., McKenzie, J.A., 1998. Stromatolite– thrombolite associations in a modern environment, Lee Stocking Island, Bahamas. Palaios 13, 201– 212. Ferrandez-C anadell, C., 2002. Multicellular-l ike compartmentalization of cytoplast in fossil larger foraminifera. Lethaia 35, 121– 129. Ferrandez-C anadell, C., Serra Kiel, J., 1992. Morphostructure and paleobiology of Discocyclina Guembel, 1870. J. Foramin. Res. 22, 147– 165. Ferrandini, M., BouDagher-F adel, M.  K., Ferrandini, J., Oudet, J., Andre, J.  -. P., 2010. New observations about the Miogypsinidae of the Early and Middle Miocene of Provence and Corsica (France) and northern Sardinia (Italy). Annales de Paleontologie, 96, 67– 94. doi:10.1016/ j.annpal.2011.04.002. Ferrero, E., 1987. Miogypsinidi della serie oligo- miocenica della Collina di Torino. Bollettino della Societa Paleontologica Italiana 26, 119– 150. References 647 Ferrero, E., Maia, F., Tonon, M., 1994. Biostratigrafia dei livelli bioclastici del Monferrato. Atti Ticinensi di Scienre della Terra, Serie Speciale 1, pp. 227– 230. Firestone, R. B., West, A., Kennett, J. P., Becker, L., Bunch, T. E., Revay, Z. S., Schultz, P. H., Belgya, T., Kennett, D. J., Erlandson, J. M., Dickenson, O. J., Goodyear, A. C., Harris, R. S., Howard, G. A., Kloosterman, J. B., Lechler, P., Mayewski, P.  A., Montgomery, J., Poreda, R., Darrah, T., Hee, S.  S. Que, Smitha, A.  R., Stich, A., Topping, W., Wittke, J. H., and Wolbach, W. S., 2007. Evidence for an extraterrestrial impact event 12,900 years ago that contributed to megafaunal extinctions and the Younger Dryas cooling. Proceedings of the National Academy of Sciences 104, 16016–1 6021. Fischer, G., Arthur, M.A., 1977. Secular variations in the pelagic realm. Soc. Econ. Paleontolog. Mineralog. 25, 19–5 0. Flakowski, J., Bolivar, I., Fahrni, J., Pawlowski, J., 2005. Actin phylogeny of Foraminifera. Journal of Foraminiferal Research, 35, 93– 102. Flügel, E., 1994. Pangean shelf carbonates, controls and paleoclimate significance of of Permian and Triassic reefs. In: Klein, G. de V., (Ed.), Pangea, Paleoclimate, Tectonics and Sedimentation During Accretion, Zenith, and Breakup of a Supercontinent, GSA special paper, Boulder, CO, Vol. 288, pp. 247– 266. Fourcade, E., 1966. Murciella cuvillieri n. gen., n. sp., nouveau Foraminifere du Sénonien supérieur du SE de L’Éspagne. Revue de Micropaléontolgie 9, 147– 155. Fourcade, E., Mouty, M., Teherani, K.K., 1997. Levantinella nov.gen. et re´ vision du genre Mangashtia Henson, grands foraminife` res du Jurassique et du Creétacé du Moyen- Orient. Geobios 30, 179–1 92. Fourcade, E., Neumann, M., 1966. A  propos des genres Labyrinthina Weynschenk 1951 et Lituosepta Cati 1959. Rev. Micropaleontol. 8, 233–2 39. Fourcade, E., Raoult, J.-F ., Vila, J.-M ., 1972. Debarina hahounerensis n. gen., n. sp., nouveau Lituolide (Foraminife` re) du Cre´ tace´ infe´ rieur constantinois (Algerie). C. R. Acad. Sci. Paris, 274, 191– 193. Fowell, S.J., Olsen, P.E., 1993. Time- calibration of Triassic/ Jurassic microfloral turnover, eastern North America. Tectonophysics 222, 361– 369. Fowell, S.J., Traverse, A., 1995. Palynology and age of the upper Blomidon formation, Fundy Basin, Nova Scotia. Rev. Palaeobot. Palynol. 86, 211– 233. Freudenthal, Th., 1969. Stratigraphy of neogene deposits in the Khania Province, Crete, with special reference to foramin- ifera of the family Planorbulinidae and the genus Heterostegina. Utrecht Micropaleontological Bulletins, 1, 1– 208. Freudenthal, Th., 1972. On some larger orbitoidal foraminifera in the Tertiary of Senegal and Portuguese Guinea, Proceedings of the 4th African Micropal. Coll., (Abidjan, 1970), pp. 144– 162. Frost, S.H., Langenheim, R.L., 1974. Cenozoic Reef Biofacies. Tertiary Larger Foraminifera and Scleractinian Corals from Mexico. Northern Illinois University Press, De Kalb, pp. 1– 388. Fugagnoli, A., 1999. Cymbriaella, a new foraminiferal genus (Textulariina) from the Early Jurassic of the Venetian Prealps (Northeastern Italy). Revue de Micropaleontologie 42, 99– 110. Fugagnoli, A., 2000. First record of Everticyclammina Redmond 1964 (E. praevirguliana n. sp.; Foraminifera) from the Early Jurassic of the Venetian Prealps (Calcari Grigi, Trento Platform, Northern Italy). Journal of Foraminiferal Research, 30, 126– 134. Fugagnoli, A., 2004. Trophic regimes of benthic foraminiferal assemblages in Lower Jurassic shallow water carbonates from Northeastern Italy (Calcari Grigi, Trento Platform, Venetian Prealps). Paleogeography, Paleoclimatology, Paleoecology, 205, 111– 130. Furrer, V., and Septfontaine, M., 1977. Nouvelles données biostratigraphiques (à l’aide des foraminiferès) dans le Dogger. Faciés briançonnais des Préalpes médianes romandes (Suisse). Eclogae geol. Helv., 70, 717– 737. Gallagher S.J., 1998. ControÌs on the distribution of calcareous Foraminifera in the Lower Carboniferous of Ireland. Marine Micropaleontology, 34, 187– 211. Gallagher, S.J., Somerville, I.D., 2003. Lower Carboniferous (late Viséan) platform development and cyclicity in south- ern Ireland: foraminiferal biofacies and lithofacies evidence. Rivista Italiana di Paleontologia e Stratigrafia, 109, 159– 171 Galloway, J.J., 1933. A  Manual of the Foraminifera. James Furman Kemp Memorial ser. publ. 1, Principia. Press, Bloomington, Indiana, 183p. García- Hernández, M., González- Donoso, J.M., Linares, A., Rivas, P., Vera, J. A., 1978. Características ambientales del Lías inferior y medio en la Zona Subbética y su significado en la interpretación general de la Cordillera. In: Universidad De Granada, Reunión sobre la Geodinámica de la Cordillera Bética y Mar de Alborán, Granada, Spain, May 12–1 4, 1976. University of Granada, Granada, 125– 157. Gargouri, S., Vachard, D., 1988. On Hemgordiopsis and other porcellaneous foraminifera from Jebel Tebaga (Upper Permian), Tunisia. Revue de Paléobiologie, Special vol. 2, Benthos ‘86, 56– 688. Gast, R.J., Caron, D.A., 1996. Molecular phylogeny of symbiotic dinoflagellates from Foraminifera and Radiolaria. Molecular Biology and Evolution, 13, 1192– 1197. Gattuso, J.- P., Buddemeier, R.W., 2000. Calcifcation and CO2. Nature 407, 311– 312. Gaździcki, A., Trammer, J., and Zawidzka, K., 1975. Foraminifers from the Muschelkalk of Southern Poland, Acta Geologica Polonica, 25, 285– 298. 648 References Gersonde, R., Kyte, F.T., Bleil, U., Diekmann, B., Flores, J.A., Gohlk, K., Grahl, G., Hagen, R., Kuhn, G., Sierro, F.J., Völker, D., Abelmann, A., and Bostwick, J.A., 1997. Geological record and reconstruction of the late Pliocene impact of the Eltanin asteroid in the Southern Ocean. Nature, 390, 357– 363. Ghose, B.K., 1977. Paleoecology of the Cenozoic reefal foraminifers and algae - a brief review. Palaeogeography, Palaeoclimatology, Palaeoecology, 22, 231– 256. Glaessner, M.F., 1945. Principles of Micropaleontology. Melbourne University Press, 296p. Glikson, A., 2005. Asteroid/ comet impact clusters, flood basalts and mass extinctions: Significance of isotopic age overlaps. Earth Planet. Sci. Lett. 236, 933– 937. Gohrbandt, K.H.A., 1966. Some Cenomanian foraminifera from northwest Libya. Micropaleontology 12, 65–7 0. Gold, D. P., Burgess, P. M., BouDagher- Fadel, M. K., 2017a. Carbonate drowning successions of the Bird’s Head, Indonesia. Facies. doi:10.1007/s 10347-0 17- 0506-z . Gold, D.  P., White, L.  T., Gunawan, I., BouDagher- Fadel, M.  K., 2017b. Relative sea- level change in western New Guinea recorded by regional biostratigraphic data. Marine and Petroleum Geology, 86, 1133–1 158. doi:10.1016/ j.marpetgeo.2017.07.016. González-D onoso, J.  M., Linares, A., and Rivas, P., 1974. El Lías inferior y medio de Poloria (serie del Zegri, Zona Subbética, norte de Granada). Estudios Geológicos, 30, 639– 654. Görög, A., Arnaud Vanneau, A., 1996, Lower Cretaceous Orbitolinas from Venezuela: Micropaleontology, v. 42, p. 65–7 8. Gould, S.J., 1977. Ontogeny and Phylogeny. Belknap Press of Harvard University Press, Cambridge, 357p. Gould, S.J., 2002. The Structure of Evolutionary Theory. Belknap Press of Harvard University Press, Cambridge, 1433p. Grimsdale, T.F., 1959. Evolution in the American Lepidocyclinidae (Cainozoic Foraminifera), an interim view. Proc. Koninklijke Nederlandse Akadamie Wetenschappen, Amsterdam, Ser. B 62, 8– 33. Grønlund, H., Hansen, H.J., 1976. Scanning electron microscopy of some recent and fossil nodosariid foraminifera. Bulletin of the Geological Society of Denmark, 25, 121– 134. Groussin, M., Pawlowski, J., Yang, Z., 2011. Bayesian relaxed molecular clock analysis of divergence times: implications for the speciation timing in Foraminifera. Molecular Phylogeny and Evolution 61, 157– 166. Groves, J.R., 2000. Suborder Lagenina and other smaller foraminifers from Uppermost Pennsylvanian- Lower Permian rocks of Kansas and Oklahoma. Micropaleontology 46(4), 285– 326. Groves, J.R., 2005. Fusulinid wall structure in the Profusulinella–F usulinella evolutionary transition. Bulletins of American Paleontology, 369, 199– 218. Groves, J.R., Altiner, D., 2004. Survival and recovery of calcareous foraminifera pursuant to the end- Permian mass extinc- tion. C. R. Palevol. 4, 487– 500. Groves, R., Altiner, D., Rettori, R., 2003. Origin and early evolutionary radiation of the order Lagenida (foraminifera). Journal of Paleontology, 77, 831– 843. Groves, J.R., Altiner, D., Rettori, R., 2005. Decline and recovery of lagenide foraminifers in the Permian– Triassic bound- ary interval (Central Taurides, Turkey). Paleontological Society Memoir 62 [supplement to Journal of Paleontology, 79(4)], 38. Groves, J.R., Boardman, D.R., 1999. Calcareous smaller foraminifers from the Lower Permian Council Grove Group near Hooser, Kansas. Journal of Foraminiferal Research, 29, 243– 262. Groves, J.R., Kulagina, E., Villa, E., 2007. Diachronous appearances of the Pennsylvanian fusulinid/ Profusulinella in Eurasia and North America. Journal of Paleontology, 81, 227– 237. Groves, J.R. and Lee, A. 2008. Accelerated rates of foraminiferal origination and extinction during the Late Paleozoic ice age. Journal of Foraminiferal Research, 38, 74– 84. Groves, J.R., Nemyrovska, T.I., Alekseev, A.S., 1999. Correlation of the type Bashkirian Stage (Middle Carboniferous, South Urals) to the Morrowan and Atokan series of the midcontinental and Western United States. Journal of Paleontology, 73, 529– 539. Groves, J.R., Rettori, R., Altiner, D., 2004. Wall structures in selected Paleozoic lagenide foraminifera. Journal of Paleontology, 78, 245– 256. Groves, J.R., Wahlman, G.P., 1997. Biostratigraphy and evolution of Upper Carboniferous and Lower Permian smaller foraminifers from the Barents Sea (offshore Arctic Norway). Journal of Paleontology, 71, 758– 779. Gŭsić, I., Velić, I., 1978. Lituolipora polymorpha n. gen., n. sp. (Foraminiferida Lituolacea?) from the Middle Liassic of the Outer Dinarids in Croatia and the establishment of a new family, Lituoliporidae. Geol. Vjesn. Inst. Geol. Istraz., 30, 73–9 3. Gutnic, M., Moullade, M., 1967. Donnees nouvelles sur le Jurassique et le Crétacé inferieur du Barla Dag au sud de Senitkent (Taurus de Pisidie, Turquie). Ankara Maden Tetkik ve Arama Enstitusu. Bull., 69, 60– 78. Gvirtzman, G., Buchbinder, B., 1978. The late tertiary of the coastal plain of Israel and its bearing on the history of the southeastern Mediterranean coast. Initial reports DSDP 42(part 2), 1192– 1222. Hageman, S.A., Kaesler, R.L., 1998. Wall structure and growth of fusulinacean foraminifera. Journal of Paleontology, 72, 181–1 90. Hallam, A., 1961. Cyclothems, transgressions and faunal changes in the Lias of northwest Europe. Trans. Edinburgh Geol. Soc., 18, 132– 174. Hallam, A., 1978. Eustatic cycles in the Jurassic. Palaeogeography, Palaeoclimatology, Palaeoecology, 23, 1– 32. References 649 Hallam, A., 1986. The Pliensbachian and Tithonian extinction events. Nature 319, 765– 768. Hallam, A., 1990. The end- Triassic mass extinctionevent. In: Sharpton, V. L., and Ward, P. D., (eds), Global catastrophes in Earth history, Geological Society of America, Special Paper, 247, 577– 583. Hallam, A., 1995. Oxygen- restricted facies of the basal Jurassic of north west Europe. Hist. Biol. 10, 247–2 57. Hallam, A., Wignall, P.B., 1997. Mass Extinctions and Their Aftermath. Oxford University Press, Oxford, 307p. Hallam, A., Wignall, P.B., 2000. Facies change across the Triassic– Jurassic boundary in Nevada, USA. J. Geol. Soc. 156, 453–4 56. Hallet, D., 1966. A study of some microfossils from the Yoredale series of Yorkshire. PhD thesis, University College London. Hallock, P., 1979. Trends in test shape in large, symbiont-b earing foraminifera. Journal of Foraminiferal Research, 9, 61–6 9. Hallock, P., 1981. Production of carbonate sediments by selected large benthic foraminifera on two Pacific coral reefs. J. Sediment. Petrol. 51, 467– 474. Hallock, P., 1984. Distribution of selected species of living algal symbiont- bearing foraminifera on two Pacific coral reefs. Journal of Foraminiferal Research, 14, 250– 261. Hallock, P., 1985. Why are larger foraminifera large? Paleobiology 11, 195– 208. Hallock, P., 1999. Advantages of algal symbiosis. In: Sen Gupta, B.K. (Ed.), Modern Foraminifera. Kluwer Academic, New York, pp. 123– 139. Hallock, P., 2000. Larger foraminifera as indicators of coral- reef vitality. Environ. Micropaleontol. 15, 121–1 50. Hallock, P., 2001. Coral reefs, carbonate sediment, nutrients, and global change. In:  Stanley, G.D. (Ed.), Ancient Reef Ecosystems, their Evolution, Paleoecology and Importance in Earth History. Kluwer Academic/ Plenum Publishers, New York, pp. 388–4 27. Hallock, P., 2005. Global change and modern coral reefs: New opportunities to understand shallow-w ater carbonate depo- sitional processes. Sediment. Geol. 175, 19– 33. Hallock, P., Glenn, E.C., 1985. Paleoenvironmental analysis of foraminiferal assemblages, a tool for recognizing deposi- tional facies in lower Miocene reef complexes. Journal of Paleontology, 59, 1382– 1394. Hallock, P., Glenn, E.C., 1986. Larger foraminifera, a tool for paleoenvironmental analysis of Cenozoic carbonate facies. Palaios 1, 55–6 4. Hallock, P., Hansen, H.J., 1979. Depth adaptation in Amphistegina, change in lamellar thickness. Bulletin of the Geological Society of Denmark, 27, 99– 104. Hallock, P., Röttger, R., Wetmore, K., 1991. Hypotheses on form and function in foraminifera. In: Lee, J.J. and Anderson, O.R., (eds), Biology of Foraminifera. Academic Press, New York, 41– 72. Hallock, P., Talge, H.K., 1994. A predatory foraminifer, Floresina amphiphaga, n. sp., from the Florida Keys. Journal of Foraminiferal Research, 24, 210-213. Hamaoui, M., Fourcade, E., 1973. Revision de Rhapydionimae (alveolinidae, foraminiferes). Bull. Cent. Rech. Pau-S PNA 7, 361–4 35. Hansen, H.  J., Buchardt, B. 1977, Depth distribution of Amphistegina in the Gulf of Elat. Utrecht Micropal. Bull., 15, 205– 224. Hanzawa, S., 1957. Cenozoic foraminifera of Micronesia. Geol. Soc. Am. Mem. 66, p. 163. Hanzawa, S., 1964. The phylomorphogeneses of the Tertiary foraminiferal families Lepidocylinidae and Miogypsinidae. Tohoku Univ. Sci. Repts., 2nd. ser. (Geol.), 35, 295– 313. Hanzawa, S., 1965. The ontogeny and the evolution of larger Foraminifera. Tohoku Univ. Sci. Repts., ser. 2 (Geol.), 36, 239–2 56. Haq, B.U., Al-Q ahtani, A.M., 2005. Phanerozoic cycles of sea-l evel change on the Arabian platform. GeoArabia 10, 127–1 60. Harney J.N., Hallock P., Talge H.K.0, 1998. Observations on a trimorphic life in Amphistegina gibbosa populations from the Florida Keys. Journal of Foraminiferal Research, 28, 141– 147. Harries, P.J., Little, C.T.S., 1999. The early Toarcian (Early Jurassic) and the Cenomanian–T uronian (Late Cretaceous) mass extinctions, similarities and contrasts. Paleogeography, Paleoclimatology, Paleoecology, 154, 39– 66. Haynes, J.R., 1965. Symbiosis, wall structure and habitat in foraminifera. Contributions of the Cushman Foundation of Foraminiferal Research, 16, 40– 43. Haynes, J.R., 1981. Foraminifera. MacMillan, London, 433p. Hemleben, C.H., Be, A.W.H., Anderson, O.R., Tuntivate-C hoy, S., 1977. Test morphology, organic layers and chamber formation of the planktonic foraminifer Globorotalia menardii (D’Orbigny). Jour. Foram. Res., 7, 1– 25. Henderson, A. L., Najman, Y., Parrish, R. R., Carter, A., BouDagher-F adel, M. K., & et al., 2008. Constraints to the timing of India- Eurasia collision determined from the Indus Group: a reassessment. Himalayan Journal of Sciences, 5 (7), 64-? . doi:10.3126/ hjs.v5i7.1266. Henderson, A. L., Najman, Y., Parrish, R., BouDagher- Fadel, M. K., Barford, D., Garzanti, E., and Andò, S., 2010. Geology of the Cenozoic Indus Basin sedimentary rocks: Paleoenvironmental interpretation of sedimentation from the west- ern Himalaya during the early phases of India- Eurasia collision. Tectonics, 29, TC6015. doi:10.1029/ 2009TC002651. Henson, F.R.S., 1948. Larger imperforate Foraminifera of South- Western Asia. Brit. Mus. (Nat.Hist.), London, p. 127. Henson, F.R.S., 1950. Middle Eastern Tertiary Peneroplidae (Foraminifera), with Remarks on the Phylogeny and Taxonomy of the Family. Doctoral thesis, University of Leiden, p. 70. 650 References Hesselbo, S.P., Robinson, S.A., Surlyk, F., Piasecki, S., 2002. Terrestrial and marine extinction at the Triassic–J urassic boundary synchronized with major carbon- cycle perturbations, a link to initiation of massive volcanism? Geology 30, 251– 254. Higgins, J.A., Schrag, D.P., 2006. Beyond methane, towards a theory for Paleocene– Eocene thermal maximum. Earth Planet. Sci. Lett. 245, 523–5 37. Hofker, J. 1927. The foraminifera of the Siboga Expedition, Pt. 1, Tinoporidae, Rotaliidae, Nummulitidae, Amphisteginidae. E. J. Brill, Leiden, 74 p. Hofker, J., 1965. Methods for study and preparation of foraminifera. In:  Kummel, B., Raup, D.  (Eds), Handbook of Paleontological Techniques. Freeman W.H. and Co., San Francisco, pp. 251– 256. Hohenegger, J., 2000. Coenoclines of larger foraminifera. Micropaleontology 46 (Suppl., 127–1 51. Hohenegger, J., 2004. Depth coenoclines and environmental consideration of Western Pacific larger foraminifera. Journal of Foraminiferal Research, 34, 9– 33. Hohenegger, J., Piller, W., 1975, Wandstrukturen und Grossgliederung der foraminiferen, Sitzungsberichten Osterreichisch Akademie der Wissenscaften, Mathematisch- naturwissenschaftliche Klasse, Abteilung I, 184, 67–9 6. Hohenegger, J., Yordanova, E., 2001. Depth-t ransport functions and erosion– deposition diagrams as indicators of slope inclination and time-a veraged traction forces, application in tropical reef environments. Sedimentology 48, 1025– 1046. Hohenegger, J., Yordanova, E., Hatta, A., 2000. Remarks on western Pacific nummulitidae (foraminifer- ida). Journal of Foraminiferal Research, 30, 3– 28. Holzmann, M., Hohenegger, J., Pawlowski, J., 2003. Molecular data reveal parallel evolution in nummulitid foraminifera. Journal of Foraminiferal Research, 33, 277– 284. Hottinger, L., 1960. Uber paleocaene und eocaene Alveolinen. Eclogae geologicae Helvetiae 53, 265–2 83. Hottinger, L., 1967. Foraminiferes imperfores du Mesozoique marocain. Notes et Mémoires Service Géologique. 209, 1– 179. Hottinger, L., 1969. The foraminiferal genus Yaberinella Vaughan, 1928, with remarks on its species and its systematic pos- ition. Eclogae geologicae Helvetiae 62, 743– 749. Hottinger, L., 1973. Selected Paleogene larger foraminifera. In: Hallam, A. (Ed.), Atlas of Palaeobiogeo- graphy. Elsevier, Amsterdam, pp. 443–4 52. Hottinger, L., 1977. Foraminifères operculiniformes. Mem. Mus. Natl. Hist. Nat. (Paris), C 40, 1– 159. Hottinger, L., 1978. Comparative anatomy of elementary shell structures in selected larger Foraminifera. Hottinger, L., 1979. Comparative anatomy of elementary shell structures in selected larger foraminifera. In: Headley, R.H., Adams, C.G. (Eds), Foraminifera. Academic Press, London, Vol. 3, pp. 203– 266. Hottinger, L., 1982. Larger foraminifera, giant cells with a historical background. Naturwissenschaften 69, 361–3 71. Hottinger, L., 1996. Sels nutritifs et biosedimentation. Mem. Soc. Geol. France n.s. 169, 99– 107. Hottinger, L., 1997. Shallow benthic foraminifera1 assemblages as signals for depth of their deposition and their limitations. Bulletin de la Socie´ te´ Ge´ ologique de France 168, 491– 505. Hottinger, L., 2000. Functional morphology of benthic foraminiferal shells, envelopes of cells beyond measure. Micropaleontology 46(Suppl. 1), 57–8 6. Hottinger, L., 2001. Archaiasinids and related porcelaneous larger foraminifera from the Late Miocene of the Dominican Republic. Journal of Paleontology, 75, 475– 512. Hottinger, L., 2006. Illustrated glossary of terms used in foraminiferal research. Carnets de Géologie / Notebooks on Geology, Memoir 2006/ 02. Hottinger, L., 2007. Revision of the foraminiferal genus Globoreticulina Rahaghi, 1978, and of its associated fauna of larger foraminifera from the late Middle Eocene of Iran. Notebooks on Geology (Brest), CG2007-A 06, 1– 51. Hottinger, L., 2009. The Paleocene and Earliest Eocene foraminiferal family Miscellaneidae: neither nummulitids nor rotali- ids. Carnets de Géologie (Brest), article 2009/ 06, 1– 41. Hottinger, L., 2013. Micropaleontology in Basel (Switzerland) during the twentieth century, rise and fall of one of the smaller fields of the life sciences. In Bowden, A.J., Gregory, F.J., and Henderson, A.S. (eds.):  Landmarks in Foraminiferal Micropaleontology: History and Development. The Micropalaeontological Society Special Publications No. 6, Geological Society Publishing House, Bath, U.K., 317– 335. Hottinger, L., Caus, E., 1982. Marginoporiform structure in Ilerdorbis decussatus n. gen. n. sp., a Senonian, agglutinated, discoidal foraminifer. Eclogae Geologicae Helvetiae 75, 807– 819. Hottinger, L., Dreher, D., 1974. Differentiation of Protoplasm in Nummulitidae (Foraminifera) from Elat, Red Sea. Mar. Biol. 25, 41–6 1. Hottinger, L., 1978. Comparative Anatomy of Elementary Shell Structures in Selected Larger Foraminifera. In: Hedley, R.H., and Adams, C.G., (eds), Foraminifera 3. Acad.Press, London, 203–2 66. Hottinger, L., Drobne, K., 1980. Early Tertiary conical imperforate foraminifera. Rasprave IV. razr., SAZU 22, 188– 276. Hottinger, L., 1982. Larger Foraminifera, Giant Cells with a Historical Background. Naturwissenschaften, 69, 361–3 71. Hottinger, L., Drobne, K., 1989. Late Cretaceous, larger complex Miliolids (foraminifera) endemic in the pyrenean faunal province. Facies 21, 99– 134. Hottinger, L., Drobne, K., Caus, E., 1989. Late Cretaceous, Larger, Complex Miliolids (Foraminifera) Endemic in the Pyrenean Faunal Province. Facies (Erlangen), 21, 99– 134 References 651 Hottinger, L., Halicz, E., Reiss, Z., 1991. The foraminiferal genera Pararotalia, Neorotalia, and Calcarina, taxonomic revi- sion. Journal of Paleontology, 65, 18– 33. Hottinger, L., Halicz, E., Reiss, Z., 1993. Recent Foraminifera from the Gulf of Aqaba, Red Sea. Opera Slovenian Acad. Sci. Arts (SAZU), 33, 1–1 79. Hottinger, L., Leutenegger, S., 1980. The structure of calcarinid foraminifers. Schweiz. Palaont. Abh. Basel. 101, 115–1 54. Hottinger, L., Romero, J., Caus, E., 2001. Architecture and revision of the pellatispirines, planispiral canaliferous foramin- ifera from the Late Eocene Tethys. Micropaleontology 47(Suppl. 2), 35– 77. Hottingerina from the Paleocene of Turkey. Micropaleontology, 45, 113– 137. Hu, X., Wang, J., Boudagher- Fadel, M. K., Garzanti, E., Wei, A., 2013. Transition from forearc basin to syn- collisional basin in Southern Tibet (Paleocene Cuojiangding Group): Implication to timing of the India-A sia initial collision and of Yarlung Zangbo ophiolite emplacement. Acta Geologica Sinica (English Edition), 87 (Supp), 27–2 8. Hu, X., Wang, J., BouDagher-F adel, M., Garzanti, E., An, W., 2015. New insights into the timing of the India-A sia collision from the Paleogene Quxia and Jialazi formations of the Xigaze forearc basin, South Tibet. Gondwana Research, 32, 76– 92. doi:10.1016/j .gr.2015.02.007. Hughes, G.W.E., 2004. Palaeoenvironments of lower Aptian agglutinated foraminifera of Saudi Arabia. In: Bubik, M., Kaminski, M.  (Eds), Proceedings of the Sixth International Workshop on Agglutinated Foraminifera. Grzybowski Foundation, special publication, 8, 197– 207. Hughes, G.W., 2014. Micropalaeontology and palaeoenvironments of the Miocene Wadi Waqb carbonate of the northern Saudi Arabian Red Sea Gulf. GeoArabia, 19, 59–108. Isozaki, Y., 2009. Integrated “plume winter” scenario for the double- phased extinction during the Paleozoic-M esozoic tran- sition: G-L B and P-T B events from a Panthalassan perspective. Journal of Asian Earth Sciences 36, 459–4 80. Jain, S., Collins, L.S., 2007. Trends in Caribbean paleoproductivity related to the Neogene closure of the Central American Seaway. Mar. Micropaleontol. 63, 57– 74. Jakobsson, M., Backman, J., Rudels, B., Nycander, J., Frank, M., Mayer, L., Jokat, W., Sangiorgi, F., O’Regan, M., Brinkhuis, H., King, J., Moran, K., 2007. The early Miocene onset of a ventilated circulation regime in the Arctic Ocean. Nature 447, 986– 990. Jenkyns, H.C., 1988. The early Toarcian (Jurassic) anoxic event, stratigraphic, sedimentary, and geochemical evidence. Am. J. Sci. 288, 101– 151. Jenkyns, H.C., 1999. Mesozoic anoxic events and paleoclimate. Zentralblatt /u r Géologie und Palâontologie, 1997, 943– 949. Jenkyns, H.C., 2003. Evidence for rapid climate change in the Mesozoic– Palaeogene greenhouse world. Royal Society of London Philosophical Transactions Ser. A, 361, 1885– 1916. Jenkyns, H.C., Clayton, C.J., 1997. Lower Jurassic epicontinental carbonates and mudstones from England and Wales, chemostratigraphic signals and the Early Toarcian anoxic event. Sedimentology 44, 687– 706. Jenkyns, H.C., Wilson, P.A., 1999. Stratigraphy, paleoceanography and evolution of cretaceous Pacific guyots, relics from a greenhouse earth. Am. J. Sci. 299, 341– 392. Jin- zhang, S., 1990. Development of fusuline foraminifera in China. Stud. Benthic Foraminifera, Benthos 90, 11–2 2. Johnson, C.C., Barron, E., Kauffman, E., Arthur, M., Fawcett, P., Yasuda, M., 1996. Middle cretaceous reef collapse linked to ocean heat transport. Geology 24, 376– 380. Johnson, C.C., Kauffman, E.G., 1990. Originations, radiations and extinctions of cretaceous rudistid bivalve species in the Caribbean province. In: Kauffman, E.G., Walliser, O.H. (Eds), Extinction Events in Earth History. Springer- Verlag, New York, pp. 305– 324. Johnson, K.G., Todd, J.A., Jackson, J.B.C., 2007. Coral reef development drives molluscan diversity increase at local and regional scales in the late Neogene and Quaternary of the southwestern Caribbean. Paleobiology 33, 24–5 2. Jolivet, L., Augier, R., Robin, C., Suc, J.- P., Rouchy, J.M., 2006. Lithospheric- scale geodynamic context of the Messinian salinity crisis. Sediment. Geol. 188– 189, 9–3 3. Jones, A.P., Price, G.D., Price, N.J., De Carli, P.S., Clegg, R.A., 2002. Impact induced melting and the development of large igneous provinces. Earth Planet. Sci. Lett. 202, 551– 561. Jones, B., Pickering, K. T., BouDagher-F adel, M. K., Matthews, S., 2003. Micropalaeontological characterisation of subma- rine fan/ channel sub- environments, Ainsa system, south- central Pyrenees, Spain. The Micropalaeontological Society Special Publications, Recent Development 55– 68. Jones, E. J. W., BouDagher- Fadel, M. K., & Thirlwall, M.F., 2002. An investigation of seamount phosphorites in the Eastern Equatorial Atlantic. Marine Geology, 183 (1– 4), 143– 162. doi:10.1016/ S0025- 3227(01)00254-7 . Jorry, S., Hasler, C.-A ., Davaud, E., 2006. Hydrodynamic behaviour of Nummulites, implications for depositional models. Facies 52, 221–2 35. Kaiho, K., Kajiwara, Y., Nakano, T., Miura, Y., Kawahata, H., Tazaki, K., Ueshima, M., Chen, Z.Q., Shi, G.R., 2001. End- Permian catastrophe by a bolide impact, evidence of a gigantic release of sulphur from mantle. Geology 29, 815–8 18. Kaiser, S.I., Steuber, T., Becker, R.T., Joachimski, M.M., 2006. Geochemical evidence for major environmental change at the Devonian– Carboniferous boundary in the Carnic Alps and the Rhenish Massif. Palaeogeography, Palaeoclimatology, Palaeoecology, 240, 146– 160. Kalia, P., Banerjee, A., 1995. Ocurrence of the genus Eoconuloides Cole and Bermudez 1944 in Rajasthan, India and its paleobiogeographic significance. Micropaleontology 41, 69– 76. 652 References Kaminski, M., 1910. The year 2010 classification of the agglutinated foraminifera. Micropaleontology, 60, 89– 108. Kauffman, E.G., Erwin, D.H., 1995. Surviving mass extinctions. Geotimes 14, 14– 17. Kauffman, E.G., Johnson, C.C., 1988. The morphological and ecological evolution of Middle and Upper Cretaceous reef- building rudistids. Palaios 3, 194– 216. Kalvoda, J., 1989. Tournaisian events in Moravia and their significance. Courier Forschungsinstitut Senckenberg 117, 353–3 58. Kalvoda, J., 2002. Late Devonian–e arly Carboniferous foraminiferal fauna: zonations, evolutionary events, paleobiogeog- raphy and tectonic implications. Folia Facultatis Scientiarium Naturalium Universitatis Masarykianae Brunensis, Geologia 39, 1–2 13. Katz, M.E., Miller, K.G., Wright, J.D., Wade, B.S., Browning, J.V., Cramer, B.S., Rosenthal, Y., 2008, Stepwise transition from greenhouse to icehouse climates: Nature Geoscience, 1, 329– 334, doi: 10.1038/ ngeo179. Keller, G., Stinnesbeck, W., Adatte, T., Stueben, D., 2003. Multiple impacts across the Cretaceous–T ertiary boundary. Earth Sci. Rev. 62, 327–3 63. Kelley, S.P., Gurov, E., 2002. Boltysh, another end-C retaceous impact. Meteorit. Planet. Sci. 37, 1031–1 043. Kerr, R.A., 2006. Creatures great and small are stirring the ocean. Science 313, p. 1717. Kiessling, W., 2002. Secular variations in the Phanerozoic reef ecosystem. In: Kiessling, W., Flu¨ gel, E., Golonka, J. (Eds), Phanerozoic Reef Patterns. SEPM Special Publications, Tulsa, Vol. 72, pp. 625– 690. Kiessling, W., Flügel, E., and Golonka, J., 2003. Patterns of Phanerozoic carbonate platform sedimentation. Lethaia, 36, 195–2 26. Kirby, M., Jackson, J., 2004. Extinction of a fast-g rowing oyster and changing ocean circulation in Pliocene tropical America. Geology 32, 1025– 1028. Knoll, A.H., Bambach, R.K., Canfield, D.E., Grotzinger, J.P., 1996. Comparative earth history and Late Permian mass extinction science 273, 452–4 57. Knoll, A.H., Bambach, R.K., Payne, J.L., Pruss, S., Fischer, W.W., 2007. Paleophysiology and end-P ermian mass extinction. Earth Planet. Sci. Lett. 256, 295– 313. Kochansky-D evidé, V., 1970,: Die Kalkalgen des Karbons vom Velebit-G ebirge (Moskovien und Kassimovien). Palaeont. Jugosl., 10, 32 p. Krebs, CJ., 1972. Ecology: The experimental analysis of distribution and abundance. New York: Harper and Row. Krüger, R., 1994. Untersuchungen zum Entwicklungsgang nrezenter Nummulitiden, Heterostegina depressa, Nummulites venosus und Cycloclypeus carpenteri. Hochschulschriften, 70, 1– 98. Küpper, K., 1954. Notes on Cretaceous larger foraminifera. I.  Genus Orbitoides in America. Contributions from the Cushman Foundation for Foraminiferal Research, 5, 63– 67. Kurbatov, V.V., 1971. Foraminiferal basis for the Jurassic section of Kugitanga and contiguous regions. In: Paleontologicheskoe obosnovanie opornykh razrezov yurskoy sistemy Uzbekistana I sopredel’nukh rayonov. Sbornik, 10, 117–1 32. Kuroda, J., Ogawa, N., Tanimizu, M., Coffin, M., Tokuyama, H., Kitazato, H., Ohkouchi, N., 2007. Contemporaneous massive subaerial volcanism and late Cretaceous oceanic anoxic event 2. Earth Planet. Sci. Lett. 256, 211–2 23. Laagland, H., 1990. Cycloclypeus in the Mediterranean Oligocene. Utrecht Micropaleontological Bulletins, 39, p. 171. Langer, M., 1995. Oxygen and carbon isotopic composition of recent larger and smaller foraminifera from the Madagan Lagoon (Papua New Guinea). Mar. Micropaleontol. 26, 215– 221. Langer, M.R., Hottinger, L., 2000. Biogeography of selected ‘‘larger’’ foraminifera. Micropaleontology 46(Suppl.  1), 105– 127. Langer, M.R., Lipps, J.H., 2003. Foraminiferal distribution and diversity, Madang Reef and Lagoon, Papua New Guinea. Coral Reefs 22, 143– 154. Larsen, A. R., Drooger, C. W. 1977: Relative thickness of the test in the Amphistegina species of the Gulf of Elat. Utrecht Micropaleontologica Bulletins, 15, 225–240. Latimer, J.C., Filippelli, G.M., 2002. Eocene to Miocene terrigenous inputs and export production, geochemical evidence from ODP Leg 177, Site 1090. Paleogeography, Paleoclimatology, Paleoecology, 182, 151– 164. Lawver, L.A., Gahagan, L.M., Coffin, M.F., 1992. The development of paleoseaways around Antarctica. In: Kennett, J.P., Warnke, D.A. (Eds), The Antarctic Paleoenvironment, A Perspective on Global Change, American Geophysical Union Monograph, Vol. 56, pp. 7– 30. LeCalvez, J., 1950. Recherches sur les foraminiferes, II. Place de la meiose et sexualite´. Archis de Zoologie Experimentale et Generale 87, 211– 243. Lee, J.J. 1990. 29. Phylum Granuloreticulosa (Foraminifera). In L. Margulis et al. (eds.), Handbook of Protoctista: 524–5 48. Lee, J.J., Anderson, O.R., 1991. Symbiosis in Foraminifera. In: Lee, J.J., Anderson, O.R. (Eds), Biology of Foraminifera. Academic Press, New York, pp. 157– 220. Lee, J.J., Burnham, B., Cevasco, M.E., 2004. A new modern soritid foraminifer, Amphisorus saurensisn. sp., from the Lizard Island group (Great Barrier Reef, Australia). Micropaleontology 50, 357–368. Lee, J.J., Morales, J., Bacus, S., Diamont, A., Hallock, P., Pawlowski, J., Thorpe, J., 1997. Progress in characterizing the endosymbiotic dinoflagellates of soritid foraminifera and related studies on some stages in the life cycle of Marginopora vertebralis. Journal of Foraminiferal Research, 27, 254– 263. References 653 Lehmann, C., Osleger, D.A., Montan˜ ez, I.P., Sliter, W., Arnaud- Vanneau, A., Banner, J., 1999. Evolution of Cupido and Coahuila carbonate platforms, Early Cretaceous, Northeastern Mexico. GSA Bull. 111, 1010– 1029. Leppig, U., 1992. Functional anatomy of Fusulinids (foraminifera): Significance of the polar torsion illustrated in Triticites and Schwagerina (Schwagerinidae). Palaontologische Zeitschrift 66, 39– 50. Less, G., 1987. Paleontology and stratigraphy of the European orthophragminae. Geologica Hungarica, Ser. Paleontologica Fasc. 51, p. 373. Less, G., 1998. The zonation of the Mediterranean Upper Paleocene and Eocene by Orthopfragminae. In: Hottinger, L., Drobne, K. (Eds), Paleogene Shallow Benthos of the Tethys, Slovenian Academy of Sciences and Arts, Ljubljana, Vol. 2, pp. 21– 43. Less, G., Kovács, L.O., 1995. Age-e stimates by European paleogene orthophragminae using numerical evolutionary corre- lation. Geobios 29, 261– 285. Less, G., Özcan, E., Báldi-Beke, M., Kollányi, K., 2007. Thanetian and early Ypresian Orthophragmines (foraminifera, Discocyclinidae and Orbitoclypeidae) from the central western Tethys (Turkey, Italy and Bulgaria) and their revised taxonomy and biostratigraphy. Rivista Italiana di Paleontogia e Stratigrafia 113, 415–4 48. Leupold, W., van der Vlerk, I.M., 1931. The tertiary. Leidsche Geologische Mededelingen 5, 611–6 48. Leutenegger, S., 1977. Ultrastructure de Foraminifères pérforés et impérforés ainsi que de leurs symbiotes. Cah. Micropal. 3, 1– 52. Leutenegger, S., 1984. Symbiosis in benthic foraminifera: Specifity and host adaptations. Journal of Foraminiferal Research, 14, 16– 35. Leven, E., 1993. Early Permian fusulinids from the Central Pamir. Revista Italiana di Paleontologia e Stratigrafia, 104, 3–4 2. Leven, E., 1997. Permian stratigraphy and fusulinida of Afghanistan with their paleogeographic and paleotectonic implica- tions. In: Stevens, C.H., Baars, D.L. (Eds), special paper, Geological Society of America, Special Paper, Washington, DC, Vol. 316, pp. 1– 135. Leven, E., 2003. Diversity dynamics of fusulinid genera and main stages of their evolution. Stratigraphy and Geological Correlation, 11, 220– 230. Leven, E., 2009. Origin of Higher Fusulinids of the Order Neoschwagerinida Minato et Honjo, 1966. Stratigraphy and Geological Correlation, 18, 290– 297. Leven, E., Gorgij, M.N., 2006. Upper Carboniferous–P ermian stratigraphy and fusulinids from the Anarak region, central Iran. Russ. J. Earth Sci. 8, p. ES2002. Leven, E., Korchagin, O.A., 2001. Permian– Triassic biotic crisis and foraminifers. Stratigr. Geol. Correl. 9, 364–3 72. Levinton, J.S., 1970. The paleoecological significance of opportunistic species. Lethaia 3, 69– 78. Levy, A., 1977. Revision micropalontologique des Soritidae actuels Bahamiens. Un nouveau genre, Androsina. Bull. Centre Recherche Exploration- Production ELF- Aquitaine 1, 393– 449. Li, J., Hu, X., Garzanti, E., BouDagher- Fadel, M., 2016. Shallow- water carbonate responses to the Paleocene– Eocene thermal maximum in the Tethyan Himalaya (southern Tibet): Tectonic and climatic implications. Palaeogeography, Palaeoclimatology, Palaeoecology, 466, 153– 165. doi:10.1016/j .palaeo.2016.11.026. Lipina, O.A,. 1965. Sistematika Turneyellid. (Systematics of the Tournayellidae). Akad. Nauk SSSR, Geologicheskii Irzstitut, Moscou., Trudy, 130, 1– 115. Little, C., Benton, M., 1995. Early Jurassic mass extinction: A global long- term event. Geology 23, 495– 498. Lobegeier, M.K., 2002. Benthic foraminifera of the family Calcarinidae from Green Island Reef, Great Barrier Reef Province. Journal of Foraminiferal Research, 32, 201– 216. Loeblich, A.R., Jr., Tappan, H., 1949. Formainifera from the Walnut Formation (lower Cretaceous) of northern Texas and southern Oklahoma. Journal of Paleontology, 23, 245– 266. Loeblich, A.R., Jr., Tappan, H., 1961. Suprageneric classification of the Rhizopodea. Journal of Paleontology, 35, 245– 330. Loeblich, A.R., Jr., Tappan, H., 1964. Protista 2, Sarcodina, chiefly ‘‘Thecamoebians’’ and Foraminiferida. In: Moore, R.C. (Ed.), Treatise on Invertebrate Paleontology, University of Kansas Press, Kansas, Part C, Vols. 1 and 2. Loeblich, A.R., Jr., and Tappan, H., 1984. Suprageneric classification of the Foraminiferida (Protozoa). Micropaleontology, 30, 1–7 0. Loeblich, A.R., Jr., and Tappan, H., 1985. Some new and redefined genera and families of agglutinated foraminifera I. Journal of Foraminiferal Research, 15, 91– 104. Loeblich, A.R., Jr., Tappan, H., 1986. Some new and redefined genera and families of Textulariina, Fusulinina, Involutinina, and Miliolina (Foraminiferida). Journal of Foraminiferal Research, 16, 334– 346. Loeblich, A.R., Jr., Tappan, H., 1988. Foraminiferal Genera and their Classification. Van Nostrand Reinhold, New York, two volumes, 2047pp. Loget, N., van Den Driessche, J., Davy, P., 2003. How did the Messinian crisis end up? Terra Nova 17, 414– 419. Lord, A.  R., Harrison, R.  W., BouDagher-F adel, M., Stone, B.  D., Varol, O., 2009. Miocene mass-t ransport sedi- ments, Troodos Massif, Cyprus. Proceedings of the Geologists’ Association, 120 (2– 3), 133– 138. doi:10.1016/ j.pgeola.2009.08.001. Louis-S chmid, B., Rais, P., Bernasconi, S.M., 2007. Detailed record of the mid- Oxfordian (Late Jurassic) positive carbon- isotope excursion in two hemipelagic sections (France and Switzerland), A  plate tectonic trigger? Paleogeography, Paleoclimatology, Paleoecology, 248, 459– 472. 654 References Lucas, G., 1939. Dictyoconus cayeuxi n. sp. Foraminifère de grande taille de l’Aalénien de l’Oranie. Compte rendu sommaire des séances de la société géologiques de la France 353– 355. Lunt, P., Allan, T., 2004. Larger foraminifera in Indonesian biostratigraphy, calibrated to isotopic dating. GRDC Museum Workshop on Micropaleontology, Bandung, June, p. 109. MacArthur, R.H., 1955. Fluctuations of animal populations, and a measure of community stability. Ecology 35, 533–5 36. MacGillavry, H.J., 1962. Lineages in the genus Cycloclypeus carpenter (foraminifera) I and II. Proc. Koninklijke Nederlandse Akadamie Wetenschappen, Amsterdam 65 (Ser. B), 429– 458. Macgillavry, H.J., 1978. Foraminifera and parallel evolution- How or why? Geologieen Mijnbouw 57, 385– 394. Macleod, N., 2013. The great extinctions: what causes them and how they shape life. The Natural History Museum, London, 192p. Macleod, N., Rawson, P. F., Forey, P. L., Banner, F. T., BouDagher-F adel, M. K., Bown, P. R., . . . Young, J. R. (1997). The Cretaceous-T ertiary biotic transition. Journal of the Geological Society, 154, 265–2 92. Mahoney, J.J., Duncan, R.A., Tejada, M.L.G., Sager, W.W., Bralower, T.J., 2005. Jurassic– Cretaceous boundary age and mid-o cean- ridge-t ype mantle source for Shatsky Rise. Geology 33, 185– 188. Mamet, B., Zhu, Z., 2005. Carboniferous and Permian algal microflora, Tarim Basin (China). Geologica Belgica 8, 3–1 3. Márquez, L., 2005. Foraminiferal fauna recovered after the Late Permian extinctions in Iberia and the westernmost Tethys area. Palaeogeography, Palaeoclimatology, Palaeoecology, 229, 137–  157. Márquez, L., Trifonova, E., 2000. Tasas evolutivas de algunos subórdenes de foraminíferos triásicos del área occidental del Thetys. Revista Española de Micropaleontología, 32, 1–1 9. Marriner, N., Morhange, C., Boudagher-F adel, M., Bourcier, M., & Carbonel, P., 2005. Geoarchaeology of Tyre’s ancient northern harbour, Phoenicia. Journal of Archaeological Science, 32, 1302– 1327. Marriner, N., Morhange, C., Rycx, Y., BouDagher-F adel, M. K., Bourcier, M., Carbonel, P., Goiran J-P , Noujaim-C lark, G., 2005. Holocene coastal dynamics along the Tyrian peninsula, palaeo-g eography of the northern harbour. Bulletin d’Archéologie et d’Architecture Libanaises, BAAL, Hors- Série 2, 61–8 9. Martin, K., 1880. Tertiarschichten auf Java. Leiden, E.J. Brill, Lief, 3, Paleont. Theil (1879– 1880), Netherland, 150– 164. Martindale, W., 1992. Calcified epibionts as palaeoecological tools, examples from the recent and Pleistocene reefs of Barbados. Coral Reefs 11, 167– 177. Martini, R., Rettori, R., Urosevic, D., Zaninetti, L., 1995. Le genre Piallina Rettori and Zaninetti (Foraminifère) dans des calcaires à Turriglomines du Trias (Carnien) de Serbie orientale (Domaine Carpatho-B alkanique). Revue de Pale´ obiologie 14, 411– 415. Marzoli, A., Bertrand, H., Knight, K., Cirilli, S., Buratti, N., Vérati, C., Nomade, S., Renne, P.R., Youbi, M., Martini, R., Allenbach, K., Rapaille, C., Zaninetti, L., Bellieni, G., 2004. Synchrony of the central Atlantic magmatic province and the Triassic– Jurassic boundary climatic and biotic crisis. Geology 32, 973–9 76. Marzoli, A., Renne, P.R., Piccirillo, E.M., Ernesto, M., Bellieni, G., De Mind, A., 1999. Extensive 200-m illion-y ear-o ld con- tinental flood basalts of the central Atlantic Magmatic Province. Science 284, 616– 618. Masse, J.-P ., 1976. Les calcaires urgoniens de Provence (Valanginian-A ptien). Stratigraphie, paléontologie, les paléoenvi- ronnements et leur évolution. Thése Doct. D’Etat, Univ. Aix- Marseille, France, Vol. I, II, p. 445. Mateu- Vicens, G, Hallock, P, Brandano, M. 2009. Test shape variability of Amphistegina d’Orbigny 1826 as a paleobathy- metric proxy: application to two Miocene examples. In: Demchuk T, Gary A, editors. Geologic problems solving with microfossils, 67– 82. (SEPM Special Publication 93). Mateu-Vicens, G., Pomar, L., Ferrandez-Ca~nadell, C., 2012. Nummulitic banks in the upper Lutetian “Buil level”, Ainsa basin, South Central pyrenean zone: the impact of internal waves. Sedimentology 59, 527–552. Matsumaru, K., 1991. On the evolutionary classification of the Family Lepidocyclinidae (Foraminiferida). Trans. Proc. Paleontol. Soc. Japan, new ser. 164, 883– 909. Matsumaru, K., 1996. Tertiary larger foraminifera (foraminiferida) from the Ogasawara Islands, Japan. Matteucci, R., Schiavinotto, F., 1985. Deux nouvelles espèces de la lignée Cycloclypeus sans ornement de la région Méditerranéenne Proceedings of the Koninklke Nederlandse Akademie van Wetenschappen. Series B. Palaeontology, geology, physics and chemistry, Netherland, 88, 123– 130. Maxwell, W.G.H., 1968. Atlas of the Great Barrier Reef. Elsevier, Amsterdam, 258p. Maxwell, W.G.H., Day, R.W., Fleming, P.J.C., 1961. Carbonate sedimentation on the Heron Island reef, Great Barrier Reef. J. Sediment. Petrol. 31, 215– 230. Maync, W., 1952. Critical taxonomic study and nomenclatural revision of the family Lituolidae based upon the prototype of the family, Lituola nautiloidea Lamarck, 1804. Contrib. Cushman Foundation for Foraminiferal Res. 3, 35–5 3. Maync, W., 1953. Hemicyclammina sigali, n.gen. n.sp. from the Cenomanian of Algeria. Contribution. Cushman Foundation for Foraminiferal Research 4, 148– 150. Maync, W., 1959. Deux nouvelles espèces Crétacées du genre Pseudocyclammina (Foraminifères). Revue de Micropaléontologie, 1, 179– 189. Maync, W., 1966. Final remarks on C.  D. Redmond’s new lituolid foraminifera from Saudi Arabia. Revue de Micropaléontologie, 9, p. 56. Maync, W., 1972. Lituonella mesojurassica n.  sp. from the Mytilus Dogger of the Swiss Prealps. Revista Española de Micropaleontologīa 4, 251–2 66. References 655 McConnaughey, T.A., 1989. 13C and 18O isotopic disequilibrium in biological carbonates, I.  Patterns. Geochimica et Cosmochimica Acta 53, 151– 162. McConnaughey, T.A., Whelan, J.F., 1997. Calcification generates protons for nutrient and bicabonate uptake. Earth Sci. Rev. 42, 95– 117. McElwain, J.C., Beerling, D.J., Woodward, F.I., 1999. Fossil plants and global warming at the Triassic– Jurassic boundary. Science 285, 1386–1 390. McKee, E.D., Oriel, S.S., Ketner, K.B., MacLachlan, M.E., Goldsmith, J.W., MacLachlan, J.C., Mudge, M.R., 1959. Paleotectonic maps of the Triassic System. United States Geological Survey, Miscellaneous Geologic Investigations Map, Nevada, I- 300, 1– 33. McKinney, M.L., McNamara, K.J., 1991. Heterochrony, The Evolution of Ontogeny. Plenum, New York, 437p. McLaren, D.J., Goodfellow, W.D., 1990. Geological and biological consequences of giant impacts. Annu. Rev. Earth Planet. Sci. 18, 123–1 71. McMenamin, M.A.S., McMenamin, D.L.S., 1990. The Emergence of Animals, The Cambrian Break- through. Columbia University Press, New York. McMillan, I., 2000. Cainozoic planktonic and larger foraminifera distributions around Southern Africa and their implica- tions for past changes of oceanic water temperatures. Suid- Afrikaanse Tydskrif vir Wetenskap 82, 66–6 9. Meinhold, G., BouDagher-F adel, M. K., 2010. Geochemistry and biostratigraphy of Eocene sediments from Samothraki Island, NE Greece. Neues Jahrbuch für Geologie und Paläontologie, 256, 17–3 8. Meinhold, G., Kostopoulos, D., Reischmann, T., Frei, D., BouDagher-F adel, M. K., 2009. Geochemistry, provenance and stratigraphic age of metasedimentary rocks from the eastern Vardar suture zone, northern Greece. Palaeogeography, Palaeoclimatology, Palaeoecology, 277, 199– 225. doi:10.1016/ j.palaeo.2009.04.005. Mello e Sousa, S. H., Fairchild, T. R., and Tibana, P., 2003. Cenozoic biostratigraphy of larger foraminifera from the Foz Do Amazonas basin, Brazil, Micropaleontology, 49, 253–266. Meric- , E., C- oruh, T., 1998. Neosivasella sungurlui, a new genus and species from the Upper Paleocene of Southeast Turkey. Micropaleontology 44, 187– 194. Mikhalevich, V.  I., 1980. Sistematika i evolyuciya foraminifer v svete novyikh dannyikh po ih citologii i ultrastruk- ture [Systematics and evolution of the Foraminifera in view of the new data on their cytology and ultrastructure]. In:  Materialy VIII mikropaleontologicheskogo sovestchaniya “Systematika I  morphologiya mikroorganismov”. 21–2 4 October, Baku: ELM: 77. Baku: Nauka (In Russian.). Mikhalevich, V.  I., 2000. Tip Foraminifera d’Orbigny, 1826– Foraminifery [The phylum Foraminifera d’Orbigny, 1826– Foraminifers]. In: Alimov, A.  F. (ed.), Protisty:  Rukovodstvo po Zoologii, pt. 1, 533–6 23. St. Petersburg:  Nauka Publishers. (In Russian, with English summary 611– 616.). Mikhalevich, V. I., 2004. On the heterogeneity of the former Textulariina (Foraminifera). In: Bubik, M. and Kaminski, M.  A., Eds., Proceedings of the Sixth International Workshop on Agglutinated Foraminifera, Krakow:  Grzybowski Foundation. Special Publication 8, 317– 349. Mikhalevich, V. I., 2009. Taxonomic position of the superorder Fusulinoida Fursenko in the Foraminifera system. 43, 117–128. Mikhalevich, V. I., 2013. New insight into the systematics and evolution of the foraminifera. Micropaleontology, 59, 493–5 27. Miller, K.G., Mountain, G.S., Wright, J.D., Browning, J.V., 2011. A 180-m illion- year record of sea level and ice volume var- iations from continental margin and deep- sea isotopic records. Oceanography, 24, 40– 53. Mohtat Aghai, P., Vachard D., Krainer K., 2009. Transported Foraminifera in the Palaeozoic deep red Nodular Limestones; Exemplified by Latest Permian Neoendothyra in the Zal section (JULFA AREA, NW IRAN). Revista Espanola de Micropaleontologia, Madrid, 41, 197– 213. Montanari, A., Koeberl, C., Hilgen, F., Coccioni, R., 2007. Hothouse, icehouse, and impacts, the Late Eocene Earth Monte Co` nero (Ancona), Italy. GSA Today 17, 60– 61. Moore, R.C., Thompson, M.L., 1949. Main divisions of Pennsylvanian Period and System. Bull. Am. Assoc. Petrol. Geologist. 33, 275– 302. Morhange, C., Espic, K., BouDagher-F adel, M. K., Doumet-S erhal, C. L., 2005. Les paléoenvironnements du port Nord de Sidon, tentative de synthèse. Bulletin d’Archéologie et d’Architecture Libanaises, BAAL, Hors- Série, 135–1 44. Murray, J.W., 1973. Distribution and Ecology of Living Benthic Foraminiferids. Heinemann, London, 274p. Murray, J.W., 1991. Ecology and distribution of benthic foraminifera. In:  Lee, J.J., Anderson, O.R. (Eds), Biology of Foraminifera. Academic Press, New York, pp. 221– 224. Murray, J.W., 2006. Ecology and Applications of Benthic Foraminifera. Cambridge University Press, Cambridge, 426p. Murray, J.W., 2007. Biodiversity of living benthic foraminifera, How many species are there? Mar. Micropaleontol. 64, 163– 176. Najman, Y., Appel, E., Boudagher- Fadel, M., Bown, P., Carter, A., Garzanti, E., Godin, L., Han, J., Liebke, U., Oliver, G,, Parrish, R., Vezzoli, G., 2010a. Timing of India- Asia collision: Geological, biostratigraphic, and palaeomagnetic constraints. Journal of Geophysical Research: Solid Earth, 115 (12). doi:10.1029/ 2010JB007673. Najman, Y., Appel, E., Boudagher-F adel, M., Bown, P., Carter, A., Garzanti, E., Godin, L., Han, J., Liebke, U., Oliver, G., Parrish, R. Vezzoli, G., 2010b. The Age of India-A sia Collision: Biostratigraphic, Sedimentological and Paleomagnetic Constraints. U.S. Geological Survey, Open- File Report 2010– 1099, 1 p. 656 References Najman, Y., Bickle, M., BouDagher-F adel, M. K., & et al., 2008. The Paleogene record of Himalayan erosion: Bengal Basin, Bangladesh. Earth and, Planetary Science Letters, 273, 1– 14. doi:10.1016/j .epsl.2008.04.028. Najman, Y., Jenks, D., Godin, L., Boudagher- Fadel, M., Millar I, Garzanti E, Horstwood M, Bracciali L., 2016. The Tethyan Himalayan detrital record shows that India–A sia terminal collision occurred by 54 Ma in the Western Himalaya. Earth and Planetary Science Letters, 459, 301– 310. doi:10.1016/j .epsl.2016.11.036. Nemkov, G.I., 1962. Remarques sur la palueouecologie des Nummulites. Vopr Micropaleontol SSSR 6, 64–7 2. Neumann, M., 1958. Revision des Orbitoidides du Crétacé et de l’Éocene en Aquitaine Occidentale, Mémoire de la Société géologique de France, 83. 1– 174. Newell, N.D., 1967. Revolutions in the history of life. In: Albritton, C.C. (Ed.), Uniformity and Simplicity. Geol. Soc. Am., special paper, 89, 63– 91. Newell, N.D., 1971. An outline history of tropical organic reefs. Am. Mus. Novit. 2465, 1– 37. Nikolaev, A.I., 2005. Foraminifers and zonal stratigraphy of Bashkirian Stage in the east of Timan–P echora province. Bulletin of Paleontological and Collections of VNIGRI, 120 pp. (In Russian). Nikolaev, S.I., Berney, C., Fahrni, J.F., Bolivar, I., Polet, S., Mylnikov, A.P., Aleshin, V.V., Petrov, N.B., Pawlowski, J., 2004. The twilight of Heliozoa and rise of Rhizaria, an emerging supergroup of amoeboid eukaryotes. PNAS 101, 8066–8 071. Noujaim Clark, G., and BouDagher-F adel, M.K., 2001. The larger benthic foraminifera and stratigraphy of the upper juras- sic/l ower cretaceous of Central Lebanon. Revue de Micropaléontologie, 44, 215– 232. Noujaim Clark, G., BouDagher-F adel, M.K., 2004. Larger Benthic Foraminifera and calcareous algae of the Upper Kesrouane Limestone Formation (Middle/U pper Jurassic) in Central Lebanon, Stratigraphy, Sedimentology and Regional Synopsis. Revue de Paléobiologie, 23, 475– 504. Noujaim Clark, G., BouDagher-F adel, M. K. (2002). Larger foraminiferal assemblages and stratigraphy of the late Jurassic Bhannes complex, Central Lebanon | Assemblages des grands foraminifèrers et stratigraphie du Jurassique supérieur du complexe de Bhannes au Liban Central. Revue de Paleobiologie, 21 (2), 679– 695. Noujaim Clark, G., BouDagher- Fadel, M.  K., 2005. Larger Benthic Foraminifera of the Upper Kesrouane Limestone Formation (Middle/ Upper Jurassic) in Central Lebanon, Stratigraphy, Sedimentology and Regional Synopsis. Revue de Paléobiologie, 23, 475– 504. O’Dogherty, L., Sandoval, J., Vera, J.A., 2000. Ammonite faunal turnover tracing sea level changes during the Jurassic (Betic Cordillera, southern Spain). J. Geol. Soc., Lond. 157, 723– 736. Oberhauser, R., 1956. Bericht über mikropaläontologische Untersuchungen im Herbst 1955. – Verh. Geol. B.-A ., Wien, 118– 119. O’Herne, L., 1972. Secondary chamberlets in Cycloclypeus. — Scripta Geol., 7, 1– 35. O’Herne, L., van der Vlerk, LI.M., 1971. Geological age determinations on a biometrical basis. (Comparison of eight parameters). Bollettino della Societa Paleontologica Italiana 10, 3– 18. Olsen, P.E., Kent, D.V., Sues, H.- D., Koeberl, C., Huber, H., Montanari, A., Rainforth, E.C., Fowell, S.J., Szajna, M.J., Hartline, B.W., 2002. Ascent of dinosaurs linked to an iridium anomaly at the Triassic–J urassic boundary. Science 296, 1305– 1307. Olsen, P.E., Koeberl, C., Montanari, A., Fowell, S., Et-T ouhami, M., Kent, D.V., 2002. Continental Triassic–J urassic bound- ary in central Pangea, Recent progress and discussion of an Ir anomaly. Geol. Soc. Am. special paper, 356, 505–5 22. Oravecz- Scheffer, A., 1987. Triassic foraminifers of the Transdanubian central range. Geol. Hungary 50, 3– 331. Orchard, M.J., 2007, Conodont diversity and evolution through the latest Permian and Early Triassic upheavals: Palaeogeography, Palaeoclimatology, Palaeoecology, 252, 93–117, doi:10.1016/j.palaeo.2006.11.037. Osawa Y., Fujita K., Umezawa Y., Kayanne H., Ide Y., Nagaoka T., Miyajima T., Yamano H., 2010. Human impacts on large benthic foraminifers near a densely populated area of Majuro Atoll, Marshall Islands. Mar Pollut Bull 60,1279–1 287. Ozawa, T., Watanabe, K., Kobayashi, D.F., 1992. Morphologic evolution in some schwagerinid and schubertellid lineages and the definition of the Carboniferous–P ermian boundary. In: Takayanagi, Y., Saito, T.  (Eds), Studies in Benthic Foraminifera, Benthos 90, Sendai, 1990. Tokai University Press, Tokyo, pp. 389– 401. Özcan, E., Less, G., Báldi-B eke, M., Kollányi, K., Kertés, B., 2006. Biometric analysis of middle and upper Eocene Discocyclinidae and Orbitoclypeidae (Foraminifera) from Turkey and updated orthophragmine zonation in the Western Tethys. Micropaleontology, 52, 485– 520. Özgen- Erdem, N., 2002. Sirelella safranboluensis n. gen., n. sp., a Foraminifer from the Lutetian of the Safranbolu Area (Northern Turkey). Micropaleontology, 48, 79– 86. Paleontol. Soc. Japan, special papers, 36, 239. Palfy, J., Smith, P., 2000. Synchrony between early Jurassic extinction, oceanic anoxic event, and the Karoo–F errar flood basalt volcanism. Geology 28, 747– 750. Palmieri, V., 1983. Biostratigraphic appraisal of Permian foraminifera from the Denison Trough–B owen Basin (central Queensland). Geological Society of Australia, Queensland Division, Permian Geology of Queensland 139–1 54. Pawlowski, J., 2000. Introduction to the molecular systematics of foraminifera. Micropaleontology 46, 1–1 2. Pawlowski, J., Bolivar, I., Fahrmi, J.F., Cavalier- Smith, T., Gouy, M., 1996. Early origin of foraminifera suggested by SSU rRNA gene sequences. Molecular Biology and Evolution, 13, 445– 450. References 657 Pawlowski, J., Bolivar, I., Fahrmi, J.F., Gouy, M., 1994. Phylogenetic position of foraminifera inferred from LSU rRNA gene sequences. Molecular Biology and Evolution, 11, 928– 938. Pawlowski J., Burki F., 2009. Untangling the phylogeny of amoeboid protists. J Eukaryot Microbiol, 56,16–2 5. Pawlowski, J., Holzmann, M., Berney, C., Fahrni, J., Gooday, A.J., Cedhagen, T., Habura, A., Bowser, S.S., 2003. The evo- lution of early foraminifera. Proc. Nat. Acad. Sci. USA 100, 11494– 11498. Payne, J.L., Lehrmann, D.J., Wei, J., Orchard, M.J., Schrag, D.P., Knoll, A.H., 2004. Large perturbations of the carbon cycle during recovery from the end-Permian extinction. Science 305, 506–509. Payne, J.L., Summers, M., Rego, B.L., Altiner, D., Wei, J., Yu, M., Lehrmann, D.J., 2011. Early and Middle Triassic trends in diversity, evenness, and size of foraminifers on a carbonate platform in south China: implications for tempo and mode of biotic recovery from the end-Permian mass extinction. Paleobiology, 37, 409–425. Pearson, P.N., van Dongen, B.E., Nicholas, C.J., Pancost, R.D., Schouten, S., Singano, J.M., Wade, B.S., 2007. Stable warm tropical climate through the Eocene Epoch. Geology 35, 211– 214. Pêcheux, M.J.-F ., 1984. Le Senonien superieur-T ertiaire du Chiapas (S.E. Mexique) et ses Macroforaminiferes. These 3e Cycle, Univ. Nice, p 154. Pêcheux, M.J.- F., 1988. Le passage Miliolidae-A lveolinidae, mécanisme d’évolution et polymorphisme: le cas de la micro- faune du Chiapas (Mexique). (Résumé). in “Benthos’ 1986”. Revue de Paléobiology,Volume Spécial, 2, 699–7 02. Pêcheux, M.J.- F., 1995. Ecomorphology of a Recent large foraminifer, Operculina ammonoides. Geobios, 28, 529–5 66. Pelissié B., Peybernès B., Rey J., 1984. Les grands foraminifères benthiques du Jurassique moyen/ supérieur du sud- ouest de la France (Aquitaine, Causses, Pyrénées). Intérêt biostratigraphique, paléoécologique et paléobiogéographique. In: Benthos ‘83, 2nd int. Symp. Benthic Foraminifera (Pau, April, 1983), 479– 489 Percival, L.M.E., Ruhl, M., Hesselbo, S.P., Jenkyns, H.C., Mather, T.A., Whiteside, J.H., Mercury evidence for pulsed vol- canism during the end-Triassic mass extinction. Proceedings of the National Academy of Sciences of the United States of America (PNAS), 114, 30, 7929–7934 (2017) doi: 10.1073/pnas.1705378114. Pignatti, J.S., 1998. The philosophy of larger foraminiferal biozonation – a discussion. In: Hottinger, L., Drobne, K. (Eds), Paleogene Shallow Benthos of the Tethys, Slovenian Academy of Sciences and Arts, Ljubljana, Vol. 2, pp. 15–2 0. Pille, L.  2008. Foraminifères et  algues calcaires du Mississippien supérieur (Viséen supérieur-S erpukhovien):  rôles bio- stratigraphique, paléoécologique et paléogéographique aux échelles locale, régionale et mondiale. PhD thesis, Université de Lille. Pinard, S., Mamet, B., 1998. Taxonomie des petits foraminifères du Carbonifère supérieur- Permien inférieur du bassin de Sverdrup, Arctique Canadien. Palaeontographica Canadiana, 15, 1– 253. Platon, E. and B. K. Sen Gupta. 2001. Benthic foraminiferal communities in oxygen- depleted environments of the Louisiana continental shelf., 147–1 64 in N.  N. Rabalais and R.  E. Turner, eds. Coastal Hypoxia:  Consequences for Living Resources and Ecosystems. Coastal and Estuarine Studies 58. Pokorny, J., 1958. Grundziige der zoologischen mikropaläontologie. Berlin, 582p. Pollitt, D. A., Anthonissen, E., Saller, A. H., Boudagher-F adel, M. K., Dickson, J. A. D., 2012. Abrupt early Eocene global climatic change as a control of carbonate facies and diagenesis:  A new record of the Palaeocene-E ocene Thermal Maximum in the Umm er Radhuma Formation, Saudi Arabia and Kuwait. Terra Nova, 24, 487– 498. Pomar, L., Baceta J., Hallock, P., Mateu-Vicens, G., Basso, D., 2017. Reef building and carbonate production modes in the west-central Tethys during the Cenozoic. Marine and Petroleum Geology 83, 261–304 Pomar, L., Hallock, P., 2008. Carbonate factories:  A conundrum in sedimentary geology. Earth- Sciences Reviews, 87, 134–1 69. Porta Della G., Villa, E., Kenter, J.A.M., 2005. Facies distribution of fusulinida in a Bashkirian– Moscovian (Pennsylvanian) carbonate platform top (Cantabrian Mountains, NW Spain), Journal of Foraminiferal Research 35, 344–3 67. Preat, A., Kasimi, R., 1995. Sédimentation de rampe mixte silico-c arbonatée des couches de transition eiféliennes- givétiennes franco- belges. Première partie: microfaciès et modèle sédimentaire. Bull. Centres Rech. Explor.-P rod. Elf- Aquitaine, 19/ 2, 329– 375. Preece, R.C., Kaminski, M.A., Dignes, T.W., 2000. Popovia johnrolandi n. sp., a new smaller agglutinated foraminifera from northern Venezuela, a biostratigraphic example of the second law of thermodynamics. In:  Hart, M.B., Kaminski, M.A., Smart, C.W. (Eds), Proceedings of the Fifth International Workshop on Agglutinated Foraminifera. Grzybowski Foundation Special Publication, 7, 403– 410. Purton, L., Brasier, M., 1999. Giant protist Nummulites and its Eocene environment: Life span and habitat insights from delta 18O and delta 13C data from Nummulites and Venericadia, Hampshire Basin UK. Geology, 27, 711– 714. Racey, A., 1995. Lithostratigraphy and larger foraminiferal (nummulitid) biostratigraphy of the Tertiary of northern Oman. Micropaleontology 41(Suppl.), 1–1 23. Racey, A., 2001. A review of Eocene nummulite accumulations, structure, formation and reservoir potential. J. Petr. Geol. 24, 79– 100. Radfords, S., 1976, Recent foraminifera from Tobago Island, West Indies. Revista Espanola de Micropaleontologia, 8, 193–2 18. Rahaghi, A., 1978. Paleogene biostratigraphy of some parts of Iran. Nat. Iran. Oil Co., Geol. Lab. Publ. 7, 1– 82. Rahaghi, A., 1980. Tertiary faunal assemblage of Qum- Kashan, Sabzewar and Jahrum areas. Nat. Iran. Oil Co., Geol. Lab. Publ. 8, 1–6 4. 658 References Raju, D.S.N., 1974. Study of Indian miogypsinidae. Utrecht Micropaleontological Bulletins, 9, p. 148. Raju, D.S.N., Meijer, M., 1977. Miogypsinidae from the Cuanza basin (Angola). Akademie Van Wetenschappen. Amsterdam, Proc., Ser. B 80, 241– 255. Ramalho, M.M., 1969. Quelques observations sur les Lituolidae (Foraminifera) du Malm portugais. Bol. Soc. Geol. Port. 17, 123– 127. Ramalho, M.M., 1971. Contribution á létude micropaléontologique et stratigraphique du Jurassique supéricur el ñu Crétacé inférieur des environs de Lisbonne (Portugal). Mem Seii Geol. Portugal, NS., 19, pp 212. Rampino, M.R., Stothers, R.B., 1988. Flood basalt volcanism during the past 250 million years. Science 241, 663–6 68. Rauser-C hernoussova, D.M., 1963. Resheriya Vgorogo Kollokvyma po sistematike Endotiroidnich foraminifer, organieo- vannogo Koordinatsionnoi komissiei po mikropaleontologii a Moskva a aprele. Otdelenie Geologo- Geograficherskich Nauk, Geologicheskii Institut Akademiya Nauk SSSR, Voprosy Paleontologii 7, 223– 227. Rauser- Chernousova, D.M., Bensh, F.R., Vdovenko, M.V., Gibshman, N.B., Leven, E.Ya., Lipina, O.A., Reitlinger, E.A., Solovieva, M.N., and Chedia, I.O., 1996. Guidebook on the Systematics of Foraminifers of Paleozoic. Academy of Sciences of Russia. Nauka Publishing House (In Russian). Rauser-C hernousova D.M., Fursenko, A.V., 1959. Determination of foraminifers from the oil-p roducing regions of the USSR. Glavnaya Redaktsya Gorno−Toplivnoy Literatury, Leningrad, Moskva, 315 pp. (in Russian) Rauser−Chernousova, D.M., Gryslova, N.D., Kireeva, G.D., Leontovich, G.E. [Leontovič, G.E.], Safonova, T.P., and Chernova, E.I. Černova, E.I.] 1951. Srednekamennougol’nye fuzulinidy Russkoy Platformy i sopredel’nyh oblastej. Spravočnik−oprede−litel’. Geologičeskij Institut, Akademiâ Nauk SSSR, Izda−tel’stvo Akademii Nauk SSSR, Moskva., 380 pp. Redmond, C.D., 1964. The Foraminiferal family Pfenderinidae in the Jurassic of Saudi Arabia. Micropaleontology 10, 251–2 63. Redmond, C. D., 1965. Three New Genera of Foraminifera from the Jurassic of Saudi Arabia. Micropaleontology, 11, 133– 140. Rego, B.L., Wang, S.C., Altiner, D., Payne, J.L., 2012. Within- and among-genus components of size evolution during mass extinction, recovery, and background intervals: a case study of Late Permian through Late Triassic foraminifera. Paleobiology, 38, 627–643. Reichel, M., 1936. Etude sur les Alveolines. Mem. Suisses Paleont. 57, 1– 147. Reichel, M., 1964. Alveolinidae. In: Moore, R.C. (Ed.), Treatise on Invertebrate Paleontology, C (Protista 2). Geological Society of America/ University of Kansas Press, Boulder/L awrence, 503–5 10. Reichel, M., 1984. Le crible apertural de Rhapydionina liburnica Stache du Maastrichtien de Vremski- Britof, Yugoslavie. In: Oertli, H.J. (Ed.), Benthos ‘83, Elf- Aqitaine, Paris, France, pp. 525– 532. Reichow, M.K., Saunders, A.D., White, R.V., Pringle, M.S., Al’Mukhamedov, A.I., Medvedev, A.I., Kirda, N.P., 2002. 40Ar/ 39Ar dates from the West Siberian Basin, Siberian flood basalt province doubled. Science 296, 1846– 1849. Reiss, Z., 1963. Reclassification of perforate foraminifera. State of Israel, Ministry of Development, Geol. Surv. Bull. 35, 1–1 11. Reiss, Z., Hottinger, L., 1984. The Gulf of Aqaba:  Ecological Micropaleontology. Ecological Studies, Springer- Verlag, Berlin, Vol. 50, pp. 1– 354. Renema, W., 2002. Larger foraminifera and their distribution patterns on the Spermonde shelf, South Sulawesi. Scripita Geologica 124, 1–2 63. Renema, W., 2005. The genus Planorbulinella (Foraminiferida) in Indonesia. Scripta Geol. 129, 137– 143. Renema, W., 2007. Fauna development of larger benthic foraminifera in the Cenozoic of Southeast Asia. In: W. Renema (Ed.), Biogeography, Time and Place: Distributions, Barriers and Islands, Springer, Dordrecht, 179– 215. Renema, W., Troelstra, S.R., 2001. Larger foraminifera distribution on a mesotrophic carbonate shelf in SW Sulawesi (Indonesia). Palaeogeografie, Paleoclimatologie, Paleoecologie 175, 125– 147. Rettori, R., 1995. Foraminiferi del Trias inferiore e medio della Tetide, Revisione tassonomica, stratigrafia ed interpretazi- one filogenetica. Université de Genève, Publications du Département de Géologie et Paléontologie, 18, pp 147. Rettori, R., Zaninetti, L., Martini, R., Vachard, D., 1993. Piallina tethydis gen. et sp. nov. (Foraminiferida) from the Triassic (Carnian) of the Kocaeli Peninsula, Turkey. J. Micropalaeontol. 12, 170– 174. Rey, J., 1997. A Liassic isolated platform controlled by tectonics, South Iberian margin, southeast Spain. Geol. Mag. 134, 235–2 47. Reytlinger, E.A., 1965. Evolution of foraminifers in the latest Permian and earliest Triassic epochs of the territory of the Transcaucasus. Akademiya Nauk SSSR, Otdelenie Nauk o Zemle, Geologicheskii Institut, Sistematika i Filogeniya Foraminifer i Ostrakod, Voprosy Mikropaleontologii 9, 45– 70. Rhumbler, L., 1909. Die Foraminiferen (Thalamophoren) der Plankton-E xpedition, 1. Teil, In: Hensen, V. (Ed.), Ergebnisse der Plankton- Expedition der Humboldt- Stittung, Leipzig, Berlin, 3, 1– 331. Richardson, S.L., 2001. Endosymbiont change as a key innovation in the adaptive radiation of Soritida (foraminifera). Paléobiology 27, 262– 289. Riding, R., 1977. Skeletal stromatolites. In: Flu¨ gel, E. (Ed.), Fossil Algae. Springer, Berlin, pp. 57– 60. Robinson, E., 1996. The occurrence of the foraminiferal genus Halkyardia in the Caribbean region. Caribb. J. Sci. 32, 72– 77. References 659 Rigaud, S., R Martini, R., Vachard, D., 2015. Early evolution and new classification of the Order Robertinida (Foraminifera). Journal of Foraminiferal Research, 45, 3– 28 Rodland, D.L., Bottjer, D.J., 2001. Biotic recovery from the end-P ermian mass extinction, behavior of the inarticulate bra- chiopod Lingula as a disaster taxon. Palaios 16, 95– 101. Rogers, A.D., 2000. The role of the oceanic oxygen minima in generating biodiversity in the deep sea. Deep- Sea Res. II 47, 119–1 48. Rögl, F., and Steininger, F. F., 1983. Vom Zerfall der Tethys zu Mediterran und Paratethys. Ann. Naturhist. Mus. Wien, 85A, 135–1 63. Rosen, B.R., 1988. Progress, problems and patterns in the biogeography of reef corals and other tropical marine organisms. Helgoländer Meeresuntersuchungen, 42, 269– 301. Ross, C. A., 1965. Fusulinids from the Cyathophyllum Limestone, Central Vestspitsbergen. Contributions from the Cushman Foundation for Foraminiferal Research, 16, 74– 86. Ross, C.A., 1967. Eoparafusulina from the Neal Ranch Formation (Lower Permian), West Texas. Journal of Paleontology, 41, 943– 946. Ross, C.A., 1969. Palaeoecology of Late Pennsylvanian Fusulinids (Foraminiferada), Texas. C. R., International Congress on Carboniferous Stratigraphy and Geology (Sheffield). Maastricht, E. Van Aelst, 4, 1429– 1440. Ross, C.A., 1973. Carboniferous Foraminiferida. In: Hallam, A., (ed.), Atlas of Palaeobiogeography. Elsevier, Amsterdam, 127– 132. Ross, C.A., 1992. Paleobiogeography of Fusulinacean Foraminifera. Studies in Benthic Foraminifera. Proceedings of the Fourth International Symposium on Benthic Foraminifera, Sendai, 1990, Tokyo, Tokai University Press, pp. 23–3 1. Ross, C.A., Ross, J.R.P., 1991. Paleozoic foraminifera. BioSystems 25, 39– 51. Roth, J.M., Droxler, A.W., Kameo, K., 2000. The Caribbean carbonate crash at the Middle to Late Miocene transition, linkage to the establishment of the modern global ocean conveyor. In: Leckie, R.M., Sigurdsson, H., Acton, G.D., Draper, G. (Eds), Proceedings of the Ocean Drilling Program, Scientific Results, College Station, TX (Ocean Drilling Program), Vol. 165, pp. 249– 273. Röttger, R., 1971. Die Bedeutung der Symbiose von Heterostegina depressa (Foraminifera, Nummulitidae) für hohe Siedlungsdichte und Karbonatproduction. Sonderdruck aus Verhandlungsbericht der Deutschen Zoologischen Gesellschaf, 65. Jahresversammlung, Gustav Fischer Verlag, 42– 47. Röttger, R., 1983. A complicated Protozoon, Heterostegina depressa: Test structures and their function. German Research Reports of the DFG, 2/ 83, 11– 13. Röttger, R., 1984. Ökologie der Groβforaminiferen. Film C 1497 des IWF, Göttingen 1983. Publikation von R. Röttger, Publ. Wissenschaftliche Film, Sekt. Biol., Ser. 16, Nr. 20/ C 1497, pp 20. Röttger, R., 1990. Biology of larger foraminifera: Present status of the hypothesis of trimorphism and ontogeny of the gam- ont of Heterostegina depressa. In: Studies in Benthic Foraminifera, Benthos ‘90, Tokai University Press, Sendai, 43– 54. Röttger, R., and Krüger, R., 1990. Observations on the biology of Calcarinidae (Foraminiferida). Marine Biology, 106, 419– 425. Röttger, R., Spindler, M., Schmal, J., Richwien, M., Fladung, M., 1984. Functions of the canal system in the rotaliid fora- minifera, Heterostegina depressa. Nature 309, 789– 791. Ruggiero, M.A., Gordon, D.P., Orrell, T.M., Bailly, N., Bourgoin, T., Brusca, R.C., et al., 2015. A Higher Level Classification of All Living Organisms. PLoS ONE, 10, e0119248. doi:10.1371/ journal.pone.0119248 Ruhl, M., Bonis, N.R., Reichart, G.J., Damste, J.S.S., and Kurschner, W.M., 2011, Atmospheric carbon injection linked to end-Triassic mass extinction: Science, 333, 430–434, doi:10.1126/science.1204255. Runnegar, B., 1982. A molecular- clock date for the origin of the animal phyla. Lethaia 15, 199– 205. Rutten, M.G., 1941. A synopsis of the Orbitoididae. Geologie en Mijnbouw 3, 34– 62. Salaj J., Borza, K., and Samuel, O., 1983. Triassic foraminifers of the west Carpathians. Geologický Ústav Dionýza Štúra, Bratislava, pp 213. Salaj, J., Trifonova, E., Gheorghian, D., 1988. A biostratigraphic zonation based on benthic foraminifera in the Triassic deposits of the Carpatho- Balkans. Revue de Paléobiologie Spec. 2, 153– 159. Salmeron, U. P., 1972. Mutación entre los generos Pararotalia y Miogypsinoides. Rev. Inst. Mex. Pet., 4, 5 - 12 Samanta, B.K., 1967. A revised classification of the family discocyclinidae galloway. Contrib. Cushman Foundation for Foraminiferal Res. 18, 164– 167. Sandberg, C.A., Morrow, J.R., Ziegler, W., 2002. Late Devonian sea- level changes, catastrophic events, and mass extinctions. In Koeberl, C.; MacLeod, K.G. Catastrophic Events and Mass Extinctions: Impacts and Beyond: Boulder, Colorado. Special Paper. 356. Geological Society of America, 473– 487. Saraswati, P.K., 2004. Ontogenetic isotopic variation in foraminifera – implications for palaeo proxy. Curr. Sci. 86, 858– 860. Saraswati, P.K., Seto, K., Nomura, R., 2003. Oxygen and carbon isotopic variation in co-e xisting larger foraminifera from a reef flat at Akajima, Okinawa, Japan. Marine Micropaleontology 50, 339– 349. Sartoni, S., Crescenti, U., 1962. Ricerche biostratigraphiche nel Mesozoica dell’Apennino meridionale. Giornale di Geologia 29, 159–3 02. Sartorio, D., Venturini, S., 1988. Southern Tethys biofacies. Agip, Milan, 235p. 660 References Schaltegger, U., Schoene, B., Bartolini, A., Guex, J., Ovtcharova, M., 2007. Precise ages for the Triassic/ Jurassic boundary and Hettangian recovery from Northern Peru. Goldschmidt Conference 2007. Geochimica et Cosmochimica Acta, Suppl., 71/ 158, A884. Schaub, H., 1981. Nummulites et Assilines de la Tethys paléogene. Taxonomy, phylogenèse et biostratigraphie. Schweizerische Paläontologisches Abhandlungen, 104– 106, pp 236. Schaudinn F., 1895. Über den Dimorphismus der Foraminiferen, Sitz. Ges. Naturforsch. Freunde. Berlin, 87– 97. Scheibner, C., Speijer, R.P., Marzouk, A., 2005. Larger foraminiferal turnover during the Paleocene/ Eocene thermal maxi- mum and paleoclimatic control on the evolution of platform ecosystems. Geology 33, 493– 496. Scheibner, C., Speijer, R.P., Marzouk, A.M., 2006. Turnover of larger foraminifera during the Paleocene– Eocene thermal maximum and paleoclimatic control on the evolution of platform ecosystems. Geology 33, 493–4 96. Scherreiks, R., Bosence, D., BouDagher- Fadel, M., 2006. The Tectono- Sedimentary Evolution of the Pelagonian Carbonate- Platform- Complex. Volumina Jurassica, IV, 64–66. Scherreiks, R., Bosence, D., BouDagher- Fadel, M. K., Melendez, G., Baumgartner, P. O., 2010. Evolution of the Pelagonian carbonate platform complex and the adjacent oceanic realm in response to plate tectonic forcing (Late Triassic to Late Jurassic), Evvoia, Greece. International Journal of Earth Sciences, 99, 1317– 1344. Scherreiks, R., Meléndez, G., BouDagher-F adel, M., Fermeli, G., Bosence, D.  2015a. Stratigraphy and tectonics of a time-t ransgressive ophiolite obduction onto the eastern, margin of the Pelagonian platform from Late Bathonian until Valanginian time, exemplified in northern Evvoia, Greece. International Journal of Earth Sciences, 104, 321. doi:10.1007/s 00531- 014-1 123-5 . Scherreiks, R., Meléndez, G., BouDagher-F adel, M., Fermeli, G., Bosence, D., 2015b. Erratum to: Stratigraphy and tec- tonics of a time- transgressive ophiolite obduction onto the eastern margin of the Pelagonian platform from Late Bathonian until Valanginian time, exemplified in northern Evvoia, Greece. International Journal of Earth Sciences, 104, 321. Scherreiks, R., Melendez, G., Boudagher-Fadel, M. K., Fermeli, G., Bosence, D.. 2016. HYPERLINK "http://iris.ucl.ac.uk/ iris/publication/1058545/1" The Callovian unconformity and the ophiolite obduction onto the Pelagonian carbonate platform of the internal Hellenides. Proceedings of the 14th International Congress, Thessaloniki, May 2016. Bulletin of the Geological Society of Greece, L, 144–152. Schlagintweit, F., Velić, I., Sokač, B. 2013. Robustoconus tisljari n. gen., n. sp., a new larger benthic foraminifer from the Middle Jurassic (Early Bajocian) of the Adriatic Carbonate Platform of Croatia. Geologia Croatica, 66, 16–2 8. Schlagintweit F., Bucur, I., 2017. Banatia aninensis n. gen., n. sp., a new complex larger benthic foraminifer from the upper Barremian of Romania. Cretaceous Research, 75, 23– 30. Schlagintweit F., Rashidi, K., 2017. Zagrosella Rigaudii n.  gen., n sp. A  new biokovinoidean foraminifer from the Maastrichtian of Iran, Acta Palaeontologica Romaniae, 3, 3– 13. Schlagintweit, F., and Wilmsen, M., 2014, Orbitolinid biostratigraphy of the top Taft Formation (Lower Cretaceous of the Yazd Block, Central Iran): Cretaceous Research, 4, 125– 133. Schlagintweit, F., Rashidi, K., and Barani, 2016. Tarburina zagrosiana n. gen., n. sp., a new larger benthic porcelaneous for- aminifer from the late Maastrichtian of Iran. Journal of Micropalaeontology https:// doi.org/ 10.1144/j mpaleo2016-0 19. Schmitz, B., Speijer, R., Aubry, M.-P ., 1996. Latest Paleocene benthic extinction event on the southern Tethyan shelf (Egypt): Foraminiferal stable isotopic (d13C, d18O) records. Geology 24, 347– 350. Schoene, B., Guex, J., Bartolini, A., Schaltegger, U., Blackburn, T.J., 2010. Correlating the end-Triassic mass extinction and flood basalt volcanism at the 100 ka level. Geology, 38, 387–390. Schroeder, R., 1962, Orbitolinen des Cenomans S¨ud-westeuropas: Paliiontologische Zeitschrift, 36, 171–202. Schroeder, R., 1964, Orbitoliniden-Biostratigraphie des Urgons nordostlich von Teruel (Spanien): Neues Jahrbuch fur Geologie und Paleontologie, Monatshefte, 462–474. Schroeder, R., 1975, General evolutionary trends in orbitolinas: Revista espanola de Micropaleontologia, Numero especial, 117–128. Schroeder,R.,Van Buchem, F. S. P., Cherchi, A., Baghbani, D.,Vincent, B., Immenhauser, A., Granier, B., 2010, Revised orbitolinid biostratigraphic zonation for the Barremian – Aptian of the eastern Arabian Plate and implications for regional stratigraphic correlations: GeoArabia Special Publication 4, 1, 49–96. Schulte, P., Speijer, R.P., Mai, H., Kontny, A., 2006. The Cretaceous– Paleogene (K–P ) boundary at Brazos, Texas: Sequence stratigraphy, depositional events and the Chicxulub impact. Sediment. Geol. 184, 77– 109. Scotchman, J.  I., Bown, P., Pickering, K.  T., BouDagher- Fadel, M., Bayliss, N.  J., Robinson, S.  A., 2014. A  new age model for the middle Eocene deep-m arine Ainsa Basin, Spanish Pyrenees. Earth-S cience Reviews. doi:10.1016/ j.earscirev.2014.11.006. Seiglie, A., Grove, K., Rivera, J.A., 1976. Revision of some Caribbean Archaiasinae, new genera, species and subspecies. Eclogae geologicae Helvetiae 70, 855– 883. Seiglie, G.A., Baker, M.B., 1983. Some West African Cenozoic agglutinated foraminifers with inner structures: Taxonomy, age and evolution. Micropaleontology 29, 391–4 03. Seiglie, G.A., Fleisher, R.L., Baker, M.B., 1986. Alveovalulnidae, n. fam., and Neogene diversification of agglutinated fora- minifers with inner structures. Micropaleontology 32, 169–1 81. References 661 Sen Gupta, B. K., 1999. Systematics of modern Foraminifera. 7– 36 in B. K. Sen Gupta, ed. Modern Foraminifera. Kluwer Academic Publishers, Dordrecht. Sepkoski J.J., Jr., 1990. The taxonomic structure of periodic extinction. In:  Sharpton, V.L., Ward, P.D. (Eds), Global Catastrophies in Earth History. Geological Society of America, special paper, 247, 33– 44. Sepkoski, J.J., Jr., 1986. Phanerozoic overview of mass extinction. In:  Raup, D.M., Jablonski, D.  (Eds), Patterns and Processes in the History of Life. Springer- Verlag, Berlin, pp. 277–2 95. Sepkoski, J.J., Jr., 1989. Periodicity in extinction and the problem of catastrophism in the history of life. Geol. Soc. Lond. J. 146, 7–1 9. Septfontaine, M., 1977. Niveaux a Foraminiferes (Pfenderininae et Valvulininae) dans le Dogger des Prealpes medianes du Chablais occidental (Haute- Savoie, France). Eclogae geologicae Helvetiae 70, 599– 625. Septfontaine, M., 1981. Les foraminiferes imperfores des milieux de plate-f orme au mesozoique, determination practique, interpretation phylogenetique et utilisation biostratigraphique. Rev. Micropal. 23, 169– 203. Septfontaine, M., 1984. Biozonation (à l’aide des foraminifères imperforés) de la plate- forme interne carbonatée Liasique du Haut Atlas (Maroc). Revue de Micropaléontologie, 27, 209– 229. Septfontaine, M.,1988. Vers une classification évolutive des Lituolides (Foraminifères) Jurassiques en milieu de plate-f orme carbonatée. Revue de Paléobiologie, Volume Spéciale 2, 229– 256. Serra- Kiel, J., Hottinger, L., Caus, E., Drobne, K., Ferrández, C., Jaurhi, A.K., Less, G., Pavlovec, R., Pignatti, J., Samso, J.M., Schaub, H., Sirel, E., Strougo, A., Tambareau, Y., Tosquella, J., Zakrevskaya, E., 1998. Larger foraminiferal bio- stratigraphy of the Tethyan Paleocene and Eocene. Bulletin de la Société Géologique de France 169, 281–2 99. Severin, K.P., Lipps, J.H., 1989. The weight-v olume relationship of the test of Alveolinella quoyi:  Implications for the taphonomy of large fusiform foraminifera. Lethaia 22, 1– 12. Sharaf, E., BouDagher- Fadel, M. K., Simo (Toni), J. A., & Carroll, A. R., 2014. A revision of the biostratigraphy and strontium isotope dating of Oligocene- Miocene outcrops in East Java, Indonesia. Berita Sedimentology, 30, 44– 81. Sharaf, E., BouDagher- Fadel, M. K., Simo, T. A., & Carroll, A. R., 2006. Biostratigraphy and strontium isotope dating of Oligocene- Miocene carbonates and siliciclastics, east Java, Indonesia. Stratigraphy, 2, 239–2 58. Sheppard, C.R.C., 2003. Predicted recurrences of mass coral mortality in the Indian Ocean. Nature 425, 294–2 97. Simmons, M.D., BouDagher- Fadel, M.K., Banner, F.T., Whittaker, J.E., 1997. The Jurassic Favusellacea, the earli- est Globigerinina. In: BouDagher- Fadel, M.K., Banner, F.T., Whittaker, J.E. (Eds), Early Evolutionary History of Planktonic Foraminifera. British Micropalaeontological Society Publica- tion Series, Chapman and Hall, pp. 17– 52. Simmons, M.D., Whittaker, J.E., Jones, R.W., 2000. Orbitolinids from Cretaceous sediments of the Middle East  – a revision of the F.R.S. Henson and associated collection. In:  Hart, M.B., Smart, C.W. (Eds), Proceedings of the Fifth International Workshop on Agglutinated Foraminifera. Gybrowski Foundation special publication, Vol. 7, pp. 411–4 37. Sinitsyna, Z.A., Sinitsyn, I.I., 1987. Biostratigrafiâ baškirskogo ârusa v stratotipe. 72 pp. Baškirskoe otdelenie Akademii Nauk SSR, Baškirsij Intitut geologii, Ufa. Sirel, E., 1988. Anatoliella, a new foraminiferal genus and a new species of Dictyokathina from the Paleocene of the Van area (East Turkey). Revue de Paléobiologie, 7, 447– 493. Sirel, E., 1993. Kastamonina abanica n. gen. n. sp., a complex lituolid (Foraminiferida) from the Upper Jurassic Limestone of the Kastamonu Area (North Turkey). Geologia Croatica, 46, 1– 7. Sirel, E., 1997. Karsella, a new complex orbitolinid (Foraminiferida) from the thanetian limestone of the Van Region (East Turkey). Micropaleontology 43, 206– 210. Sirel, E., 1999. Four new genera (Haymanella, Kayseriella, Elazigella and Orduella) and one new species of Sirel, E., 2004. Türkije’nin Mesozoyik ve Senozoyik Yeni Bentik Foraminiferleri. Jeoloji Mühendisleri Odasi Yayinlari 84 (Chamber of Geological Engineers of Turkey, publication 84), Ankara, Emegin Bilimsel Sentezi, özel Sayi 1, pp. 1–2 19. Sirel, E., Acar, S., 1982. Praebullalveolina, a new foraminiferal genus from the Upper Eocene from the Afyon and Çanakkale region (west of Turkey). Eclogae Geologicae Helvetiae 75, 821– 839. Sirel, E., Acar, S., 1993. Malatyna, a new foraminiferal genus from the Lutetian of Malatya region (East Turkey). Geologia Croatica, 46, 181–1 88. Sirotti, A., 1982. Phylogenetic classification of Lepidocyclinidae:  A proposal. Bollettino della Societa Paleontologica Italiana 21, 99– 112. Skipp, B., 1969. Foraminifera. In: McKee, E.D., Gutschick, R.C. (Eds), History of the Redwall Limestone of Northern Arizona. Geological Society of America Memoir, 114, 173– 255. Smith, A.G., Pickering, K.T., 2003. Oceanic gateways as a critical factor to initiate icehouse earth. Journal of the Geological Society London, 160, 337– 340. Smout, A.H., 1954. Lower tertiary foraminifera of the Qatar peninsula. Brit. Mus. (Nat. Hist.), Lond. 96 p. Smythe, D.K., Russell, M.J., Skuce, A.G., 1995. Intracontinental rifting from major late carboniferous quartz- dolorite dyke swarm of NW Europe. Scott. Journal of Geology 31, 151– 162. Song, H., WignalL, P. B., Chen, Z. Q., Tong, J., Bond, D. P. G., Lai, X., ZhaO, X., JiaNG, H., Yan, C., Niu, Z., Chen, J., Yang, H. & Wang, Y., 2011. Recovery tempo and pattern of marine ecosystems after the end-Permian mass extinction. Geology, 39, 739-632. Spray, J.G., Kelley, S.P., Rowley, D.B., 1998. Evidence for a late Triassic multiple impact event on earth. Nature 392, 171– 173. 662 References Stache, G., 1913. Uber Rhipidionina St. und Rhapydionina St. Jahrbuch der Geologischen Reichsanstalt 62, 659–6 80. Stanley, G.D. (Ed.), 2001. Ancient Reef Ecosystems, their Evolution, Paleoecology and Importance in Earth History. Kluwer Academic/ Plenum Press, New York, 299p. Stanley, G.D., 2003. The evolution of modern corals and their early history. Earth Sci. Rev. 60, 195– 225. Stanley SM. 2009. Evidence from ammonoids and conodonts for multiple Early Triassic mass extinctions. Proc Natl Acad Sci USA. 106, 15264–15267. Stefaniuk, L., Morhange, C., Saghieh-B eydoun, M., Frost, H., BouDagher-F adel, M. K., Bourcier, M., Noujaim-C lark, G.  (2005). Localisation et étude paléoenvironnementale des ports antiques de Byblos. Bulletin d’Archéologie et d’Architecture Libanaises, BAAL, Hors- Série 2, 19– 41. Stevens, G.R., 1971. Relationship of isotopic temperatures and faunal realms to Jurassic- Cretaceous palaeogeography, par- ticularly of the south west Pacific: J. Roy. Soc. New Zealand, v. 1, p. 145. Storey, M., Duncan, R.A., Swisher, C.C., 2007. Paleocene– Eocene thermal maximum and the opening of the northeast Atlantic. Science 316, 587– 589. Strathearn, G.E., 1986. Homotrema rubrum, symbiosis identified by chemical and isotopic analyses. Palaios 1, 48– 54. Streel, M., 1986. Miospore contribution to the Upper Famennian–S trunian event stratigraphy. Annales de la Socie´ te´ ge´ ologique de Belgique 109, 75– 92. Stüben, D., Kramar, U., Harting, M., Stinnesbeck, W., Keller, G., 2005. High-r esolution geochemical record of Cretaceous– Tertiary boundary sections in Mexico: New constraints on the K/T and Chicxulub events. Geochimica et Cosmochimica Acta 69, 2559– 2579. Sun, G., Hu, X., Sinclair, HD, BouDagher- Fadel, MK, Wang, J., 2015. Late Cretaceous evolution of the Coqen Basin (Lhasa terrane) and implications for early topographic growth on the Tibetan Plateau. Geological Society of America Bulletin. DOI: 10.1130 / B31137.1. Sweet, W.C., Yang, Z., Dickins, J.M., Yin, H., 1992. Permo-T riassic events in the eastern Tethys – An overview. In: Sweet, W.C., Zunyi, Y., Dickins, J.M., Hongfu, Y. (Eds), Permo-T riassic Events in the Eastern Tethys. Cambridge University Press, Cambridge, pp. 1– 8. Talent, J.A., 1988. Organic reef- building: Episodes of extinction and symbiosis. Senckenbergiana Lethaea 69, 315– 368. Tan Sin Hok, 1932. On the genus Cycloclypeus, Pt.1, and an appendix on the heterostegines of Tjimanggoe, S. Bantam, Java. Wetenschappelijke Mededeelingen van de Dienst van de Mijnbouw in Nederlandsch-O ost- Indië, 19, 1– 194. Tan Sin Hok, 1936. Zur Kenntniss der Miogypsiniden. Ingenieur Nederland. - Indië, Mijnbouw Geologie, 4, 45–6 1. Tanner, L.H., Kyte, F.T., 2005. Anomalous iridium enrichment in sediments at the Triassic– Jurassic boundary, Blomidon formation, Fundy Basin, Canada. Earth Planet. Sci. Lett. 240, 634– 641. Tanner, L.H., Lucas, S.G., Chapman, M.G., 2004. Assessing the record and causes of Late Triassic extinctions. Earth Sci. Rev. 65, 103– 139. Tappan, H., 1971. Microplankton, ecological succession and evolution. North American Paleontological Convention Chicago, 1969, Proceedings H, Chicago, pp. 1058– 1103. Tappan, H., Loeblich, A., 1988. Foraminiferal evolution, diversification, and extinction. Journal of Paleontology, 62, 695–7 14. Teodorovich, G.I. [Teodorovič, G.I.] 1949. The carbonate facies of upper Carboniferous–l ower Permian of the Uralo−Volga area [in Russian]. Materialy k poznaniu geologičeskogo stroeniâ SSSR. Novaâ seriâ 13, 1– 304. Ter Kuile, B., 1991. Mechanisms for calcification and carbon cycling in algal symbiont-b earing foraminifera. In: Lee, J.J., Anderson, O.R. (Eds), Biology of Foraminifera. Academic Press, New York, pp. 73– 89. Thomas, E., Brinkhuis, H., Huber, M., Ro¨ hl, U., 2006. An ocean view of the early Cenozoic world. Oceanography 19, 94– 103. Thompson, M.  L., 1964, Fusulinacea, in Loeblich, A.  R., and Tappan, H., Treatise on Invertebrate Paleontology, Part C, Protista 2; Sarcodina chiefly ‘‘Thecamoebians’’ and Foraminiferida: Geological Society America and University Kansas Press, 1, C358–C 436. Todd, R., 1960. Some observations on the distribution of Calcarina and Baculogypsina in the Pacific Science Reports Tohoku University, Ser. 2 (Geol.), spec. 4, 100–107. Tong, J., Shi, G.R., 2000. Evolution of the Permian and Triassic foraminifera in South China. In: Yin, H., Dickins, J.M., Shi, G.R., Tong, J., (Eds), Permian– Triassic Evolution of Tethys and Western Circum- Pacific, Developments in Palaeontology and Stratigraphy. Elsevier, Amsterdam, 18, 291– 307. Toomey D.F., Winland H.D., 1973. Rock and biotic facies associated with Middle Pennsylvanian (Desmoinesian) algal buildup, Nena Lucia Weld, Nolan County, Texas. AAPG Bull, 57, 1053– 1074. Hallock, P., Triantaphyllou, M., Dimiza, M., Koukousioura O., 2011. An invasive foraminifer in coastal ecosystems of the Eastern Mediterranean: Implications for understanding larger foraminiferal-dominated biofacies in the Cenozoic. In: GSA Annual Meeting in Minneapolis, Paper No. 231-10, Geological Society of America Abstracts with Programs, 43, 556 p. Trifonova, E., 1978. The foraminifera zones and subzones of the Triassic in Bulgaria: Scythian and Anisian. Geol. Balcanica 8, 85–1 04. References 663 Trifonova, E., 1984. Correlation of Triassic foraminifers from Bulgaria and some localities in Europe, Caucasus and Turkey. Geol. Balcanica, 13, 3– 24. Trifonova, E., 1992. Taxonomy of Bulgarian triassic foraminifera. I. Families Psammosphaeridae to Nodosinellidae. Geol. Balcanica, 22, 3– 50. Trifonova, E., 1993. Taxonomy of Bulgarian triassic foraminifera, II. Families Endothyriidae to Ophthamiidae. Geol. Balcanica, 23, 19–6 6. Trifonova, E., 1994. Taxonomy of Bulgarian triassic foraminifera. III. Families Spiroloculiniidae to Oberhausereliidae. Geol. Balcannica, 24, 21–7 0. Tucker, M.E., Benton, M.J., 1982. Triassic environments, climates and reptile evolution. Palaeogeography, Palaeoclimatology, Palaeoecology, 40, 361– 379. Twitchett, R.J., 2006. The palaeoclimatology, palaeoecology and palaeoenvironmental analysis of mass extinction events. Paleogeography, Paleoclimatology, Paleoecology, 232, 190– 213. Umbgrove, J.H.F., 1928. Het genus Pellatispira in het Indo- pacifische gebeid. Wetenschappelijke Mededeelingen van de Dienst van de Mijnbouw in Nederlandsch- Oost- Indië, 10, 43– 71. Urbanek, A., 1993. Biotic crisis in the history of the Upper Silurian graptolites:  A palaeobiological model. Hist. Biol. 7, 29–5 0. Vachard, D., 1973. Remarques sur les Foraminifères des calcaires griottes sensu lato (Frasnien inférieur-T ournaisien inférieur) du versant méridional de la Montagne Noire (Aude- Hérault). Comptes Rendus Sommaires de la Société géologique de France 1, 116– 118. Vachard, D., 1974. Contribution à l’étude stratigraphique et micropaléontologique (algues et foraminifères) du Dévonien- Carbonifère inférieur de la partie orientale du versant méridional de la Montagne Noire (Hérault, France). PhD Thesis, Université Pierre et Marie Curie (Paris VI), 1– 408 (unpublished). Vachard, D., Krainer, K., 2001. Smaller foraminifers of the Upper Carboniferous Auernig Group, Carnic Alps (Austria/ Italy). Rivista Italiana di Paleontologia e Stratigrafia, 107, 147– 168. Vachard D., Massa, D., and Strank, A., 1993. Le Carbonifère du sondage A1– 37 (Cyrénaique, Libye), analyse bio- stratigraphique, conséquences paléogeographiques. Revue de Micropaléontologie, 36,165– 186 Vachard, D., Martini, R., Rettori, R., Zaninetti, L., 1994. Nouvelle classification des foraminifères endothyroïdes du Trias. Geobios, Lyon 27/ 5, 543– 557. Vachard, D., Munnecke, A., Servais, T., 2004. New SEM observations of keriothecal walls implications for the evolution of the fusulinida. Journal of Foraminiferal Research, 34, 232– 242. Vachard D., Pille L., Gaillot, J. 2010. Palaeozoic Foraminifera: Systematics, palaeoecology and responses to global changes. Revue de micropaléontologie 53, 209– 254. van der Vlerk, I.M., 1922. Studiën over nummulinidae en alveolinidae. Haar voorkomen op Soembawa en haar betekenis voor de geologie van oost- Azië en Australië. Wetenschappelijke Mededeelingen van de Dienst van de Mijnbouw in Nederlandsch- Oost- Indie, 6, 329– 468. van der Vlerk, I.M., 1929. Groote foraminiferen van N.O. Borneo. Wetenschappelijke Mededeelingen van de Dienst van de Mijnbouw in Nederlandsch- Oost-I ndië, 9, 1– 45. van der Vlerk, I.M., 1959. Problems and principles of Tertiary and Quaternary stratigraphy. Quar. J. Geol. Soc. London, 105, 49– 64. van der Vlerk, I.M., 1963. Biometric research on Lepidocyclina. Micropaleontology 9, 425–4 26. van der Vlerk, I.M., 1966. Miogypsinoides, Miogypsina, Lepidocyclina, et Cycloclypeus de Larat (Moluques). Eclogae Geologicae Helvetiae 59, 421– 430. van der Vlerk, I.M., Bannink, D.D., 1969. Biometrical investigations on Operculina, Proceedings Koninklijke Nederlandse Akadamie Wetenschappen Amsterdam, B, Vol. 72, pp. 169– 174. van der Vlerk, I.M., Gloor, H., 1968. Evolution of an embryo. Genetica 39, 45– 63. van der Vlerk, I.M., Umbgrove, J.H., 1927. Tertiary Gidsforaminifera van Nederladisch Oost- Indie. Wetenschappelijke Mededeelingen van de Dienst van de Mijnbouw in Nederlandsch- Oost- Indië, 6, 3– 35. van Ginkel, A.C., Villa, E., 1999. Late fusulinellid– early schwagerinid Foraminifera; relationships and occurrences in the Las Llacerias Section (Moscovian/ Kasimovian), Cantabrian Mountains, Spain. Journal of Foraminiferal Research, 29, 263– 290. van Gorsel, J.T., 1975. Evolutionary trends and stratigraphic significance of the Late Cretaceous helicorbitoides– lepidorbitoides lineage. Utrech Micropaleontol. Bull. 12, 1– 19. van Gorsel, J.T., 1978. Late Cretaceous orbitoidal foraminifera. In:  Hedley, R.  H., Adams, C.G. (Eds), Foraminifera, Academic Press, London, 3, 1– 109. van Vessem, E.J., 1978. Study of lepidocyclinidae from Southeast Asia, particularly from Java and Borneo. Utrecht Micropaleontological Bulletins, 19, 1– 163. Vasseur, P., 1974. The overhangs, tunnels and dark reef galleries of Tule´ ar (Madagascar) and their sessile invertebrate com- munities. Proceedings of the Second International Coral Reef Symposium, Brisbane 2, 143–1 59. Vaughan, T.W., 1945. American Paleocene and Eocene larger foraminifera. Mem. Geol. Soc. Amer. 9, 1– 175. Vaughan, T.W., Cole, W.S., 1941. Preliminary report on the cretaceous and tertiary larger foraminifera of Trinidad, British West Indies. Geol. Soc. Amer., special paper, 30, 1– 137. 664 References Vdovenko, M.V., Rauzer- Chernousova, D.M., Reitlinger, E.A., Sabirov, A.A. 1993. Spravochnik po sistematike melkikh foraminifer Paleozoya (za isklyucheniem endotiroidey I permskikh mnogokamernykh lagenoidey. [A reference- book on the systematics of Paleozoic Foraminifera (excluding endothyrids and Permian multichambered lagenids)] Izdat. Nauka, Moscow, 1– 125. Vecchio, E., Hottinger, L., 2007. Agglutinated conical foraminifera from the lower Middle Eocene of the Trentinara forma- tion (southern Italy). Facies 53, 509– 533. Vera, J. A., 1988. Evolución de los sistemas de depósito en el margen ibérico de la Cordillera Bética. Revista de la Sociedad Geológica de España, 1, 373–3 91. Vermeij, G.J., 2004. Ecological avalanche and the two kinds of extinction. Evol. Ecol. Res. 6, 315– 337. Vilas, L., Masse, J.- P., Arias, C., 1995. Orbitolina episodes in carbonate platform evolution, the early Aptian model from SE Spain. Paleogeography, Paleoclimatology, Paleoecology, 119, 35–45. von der Heydt, A., Dijkstra, H.A., 2006. Effect of ocean gateways on the global ocean circulation in the late Oligocene and early Miocene. Paleoceanography 21, p. PA1011. Wade, B.S., Pearson, P.N., Berggren, W.A., Pälike, H., 2011. Review and revision of Cenozoic tropical planktonic foraminif- eral biostratigraphy and calibration to the geomagnetic polarity and astronomical time scale. Earth-Science Reviews, 104, 111–142. Wahlman, G.P., 2002. Upper Carboniferous-L ower Permian (Bashkirian-K ungurian) mounds and reefs. In W. Kiessling, E.  Flugel, and J.  Golonka [eds.], Phanerozoic Reef Patterns, 271‒338. SEPM (Society for Sedimentary Geology) Special Publication, 72. Tulsa, Oklahoma, USA. Wahlman, G.P., Konovalova, and M.V., 2002. Upper Carboniferous- Lower Permian Kozhim carbonate bank, Subpolar Ural Mountains, Russia: outcrop analog for some Timan Pechora Basin Reservoirs. In W. A. Zempolich, and H. E. Cook [eds.], Carbonate Reservoirs and Field Analogs of the CIS, 223‒245. SEPM (Society for Sedimentary Geology) Special Publication 74. Tulsa, Oklahoma, USA. Walliser, O.H., 1995. Global events and event Stratigraphy in the Phanerozoic. In: Walliser, O.H. (Ed.), Global Events and Event Stratigraphy. Springer, Berlin, 7– 19. Walliser, O.H., 2003. Sterben und Neubeginn im Spiegel der Palaeofauna. Die Bedeutung der globalen Faunenschnitte fu¨ r die Stammesgeschichte. In: Hansch, W. (Ed.), Katastrophen in der Erdgeschichte. Heilbronn (Sta¨ dtische Museen), Wendezeiten des Lebens, museo, Heilbronn (Sta¨ dtische Museen), 19, 60– 69. Walter, B., 1989. Au Valanginien supe´ rieur, une crise de la faune des bryozoaires, indication d’un important refroidisse- ment dans le Jura. Paleogeography, Paleoclimatology, Paleoecology, 74, 255– 263. Wang, J., Hu, X.., BouDagher- Fadel, M., Wu, F.Y., Sun, G.Y., 2015. Early Eocene sedimentary recycling in the Kailas area, southwestern Tibet:  Implications for the initial India–A sia collision. Sedimentary Geology, 315, 1–1 3. doi:10.1016/ j.sedgeo.2014.10.009. Ward, P.D., 2006. Out of Thin Air. Joseph Henry Press, Washington, DC, 282p. Ward, P.D., Brownlee, D., 2000. Rare Earth:  Why Complex Life is Uncommon in the Universe. Copernicus Books, New York, 335p. Ward, P.D., Haggart, J.W., Carter, E.S., et  al., 2001. Sudden productivity collapse associated with the Triassic– Jurassic boundary mass extinction. Science 292, 1148– 1151. Wardlaw, B.R., Davydov, V., Gradstein, F.M., 2004. The Permian period. In: Gradstein, F.M., Ogg, J.G., Smith, A.G. (Eds), A Geologic Time Scale. Cambridge University Press, Cambridge, 249– 270. Wedekind, R., 1937. Einfuhrung in die Grundlagen der historischen Geologie, II. Band. Microbiostratigraphie, Die korallen- und Fora- miniferenzeit. Ferdinand Enke, Stuttgart, 136 p. Wefer, G., Berger, W.H., 1980. Stable isotopes in benthic foraminifera: Seasonal variation in large tropical species. Science 209, 803– 805. Weidlich, O., 2007. Permian reef and shelf carbonates of the Arabian platform and Neo- Tethys as recorders of climatic and oceanographic changes. Geol. Soc. London, special publications, 275, 229– 253. White, L. T., Hall, R., Armstrong, R. A., Barber, A. J., BouDagher-F adel, M. K., Baxter A, Wakita K, Manning C, Soesilo, J.  (2017). The geological history of the Latimojong region of western Sulawesi. Journal of Asian Earth Sciences. doi:10.1016/ j.jseaes.2017.02.005. Whiteside, J.H., Olsen, P.E., Eglinton, T., Brookfield, M.E., Sambrotto, R.N., 2010. Compound-specific carbon isotopes from Earth's largest flood basalt eruptions directly linked to the end-Triassic mass extinction. Proceedings National Academy of Science USA. doi:10.1073/pnas.1001706107. Wignall, P.B., 2001. Sedimentology of the Triassic–J urassic boundary beds in Pinhay Bay (Devon, SW England). Proc. Geol. Assoc. 112, 349– 360. Wignall, P.B., Newton, R.J., Little, C.T.S., 2005. The timing of paleoenvironmental change and cause- and- effect relation- ships during the early Jurassic mass extinction in Europe. Am. J. Sci. 305, 1014– 1032. Wignall, P.B., Twitchett, R.J., 1996. Oceanic anoxia and the end Permian mass extinction. Science 272, 1155–1158. Wildenborg, A., 1991. Evolutionary aspects of the miogypsinids in the Oligo– Miocene carbonates near Mineo (Sicily). Utrecht Micropaleontological Bulletins, 41, 1– 140. Wilson, J. L., 1975. Carbonate Facies in Geologic History. Springer Verlag, Berlin. References 665 Woods, A.D., 2005. Paleoceanographic and paleoclimatic context of early Triassic time. Palevol. Com. Rend. Acad. Siences Paris 4, 395– 404. Wray, C.G., Martin, R., Langer, M., DeSalle, R., Lee, J.J., Lipps, J.H., 1995. Origin of the foraminifera. Proc. Natl. Acad. Sci. USA 92, 141– 145. Wyn, G., Hughes, G., 2004. Middle to Upper Jurassic Saudi Arabian carbonate petroleum reservoirs: biostratigraphy, micro- palaeontology and palaeoenvironments. GeoArabia- Manama, 9, 79– 114. Yan Song, R., Black, G., Lipps, J.H., 1994. Morphological optimization in the largest living foraminifera: Implications from finite element analysis. Paleobiology 20, 14– 26. Yordanova, E.K., Hohenegger, J., 2002. Taphonomy of larger foraminifera: Relationships between living individuals and empty tests on flat reef slopes (Sesoko Island, Japan). Facies, 46, 169– 204. Zachos, J.C., Pagani, M., Sloan, L., Thomas, E., Billups, K., 2001. Trends, rhythms, and aberrations in global climate: 65 Ma to present. Science, 292, 686– 693. Zachos, J.C., Ro¨ hl, U., Schellenberg, S.A., Sluijs, A., Hodell, D.A., Kelly, D.C., Thomas, E., Nicolo, M., Raffi, I., Lourens, L.J., McCarren, H., Kroon, D., 2005. Rapid acidification of the ocean during the Paleocene– Eocene Thermal Maximum. Science, 308, 1611– 1616. Zachos, J.C., Wara, M.W., Bohaty, S., Delaney, M.L., Petrizzo, M.R., Brill, A., Bralower, T.J., Premoli- Silva, I., 2003. A  transient rise in tropical sea surface temperature during the Paleocene– Eocene thermal maximum. Science, 302, 1551– 1554. Zaninetti, L. 1976. Les Foraminifères du Trias. Essai de synthèse et corrélation entre les domaines mésogéens européen et asiatique. Rivista Italiana di Paleontologia e Stratigrafia, 82, 1– 258 Zaninetti, L., Rettori, R., He, Y., Martini, R., 1991. Paratriasina He, 1980 (Foraminiferida, Trias medio della Cina), morfo- logia, tassonomia, filogenesi. Revue de Paléobiology, 10, 301– 308. Zaninetti, L., Rettori, R., Martini, R., 1994. Paulbronnimanninae Rettori and Zaninetti, 1993 (Foraminiferida, Ammodiscidae) and other Anisian foraminifers from the Piz Da Peres section (Valdaora-O lang, Pusteria Valley, dolo- mites, Ne Italy). Rivista Italiana di Paleontologia e Stratigrafia 100, 339– 350. Zempolich, W.G., 1993. The drowning succession in Jurassic carbonates of the Venetian Alps, Italy: A record of supercon- tinent breakup, gradual eustatic rise, and eutrophication of shallow- water environments. In: Loucks, R.G., Sarg, J.F. (Eds), Carbonate Sequence Stratigraphy: Recent Developments and Applications. Memoirs-A merican Association of Petroleum Geologists, 57, 63– 105. Zharkov, M.A., Murdmaa, I.O., Filatova, N.I., 1998. Paleogeography of the Berriasian– Barremian of the Early Cretaceous. Stratigr. Geol. Correl. 6, 47– 69. Zhuang, S.Q., 1999. A study of stalatotheca in spirotheca of fusulinids and some new genera and species with stalatotheca. Acta Micropaleontol. Sinica 6, 357– 371. Subject Index Aalenian–B ajocian regional anoxic event 257 Alveolina cf. stipes 508 Abadehella 61, 107 Alveolina compressa 320, 400 Abadehella tarazi 61, 131 Alveolina corbarica 441, 450 Abadehellidae 61, 100 Alveolina cucumiformis 505 Aberisphaera gambanica 514 Alveolina dachelensis 399 Abrardia 301, 367 Alveolina dainellii 441, 450 Abrardia mosae 367 Alveolina ellipsoidalis 397, 441, 450 Abriolina 167, 177, 198 Alveolina elliptica 21, 450, 502, 505, 508 Abriolina mediterranea 167, 198 Alveolina elliptica nuttalli 450, 505, 508 Abriolinidae 167 Alveolina elliptica var. flosculina 502 Abu Dhabi 261, 278, 358, 371, 377, 379, 380 Alveolina elongata 441, 450 Accordiella 297, 306, 330, 377 Alveolina globosa 397, 450, 505 Accordiella conica 306, 377 Alveolina indicatrix 397 Acervulina 558, 559 Alveolina katicae 502 Acervulina inhaerens 558 Alveolina leupoldi 508, 511 Acervulina (Ladoronia) vermicularis 559 Alveolina minervensis 397, 505 Acervulinidae 424, 425, 470, 558 Alveolina montipara 87 Acervulinoidea 453, 490–4 91, 569, 575 Alveolina moussoulensis 441, 450, 508 Acropora millepora 619 Alveolina munieri 441, 450 Actinocyclina 414, 415, 456, 532 Alveolina oblonga 398, 505 Actinocyclina radians 532 Alveolina ovum 323 Actinoporella podolica 350 Alveolina palermitana 508 Actinosiphon 430, 444, 479, 523 Alveolina pasticillata 397, 450 Actinosiphon semmesi 430, 523 Alveolina prorrecta 441, 450 Adrahentina 320 Alveolina quoyi 544 Adrahentina iberica 320 Alveolina schwageri 441, 450 Adriatic Sea 196 Alveolina stipes 441 Aeolisaccus 54 Alveolina subpyrenaica 397, 505, 508 Aeolisaccus dunningtoni 54 Alveolina trempina 441, 450 Afghanistan 179, 505 Alveolina violae 441, 450 Africa 348, 622 Alveolina vredenburgi 441, 450, 505 Agathammina 163, 173, 196 Alveolinaleupoldi 505 Agathammina judicariensis 163 Alveolinella 17, 18, 26, 29, 110, 443, 544, 545, 564, 565, Agathammina?austroalpina 196 566, 585, 595, 599, 609 Agglutinella 452 Alveolinella bontangensis 545 Aktinorbitoides 310, 317, 343 Alveolinella fennemai 564, 609 Aktinorbitoides browni 317 Alveolinella praequoyi 564, 565, 566, 595, 609 Alabama 631 Alveolinella quoii 585 Alanlordia 558, 559, 575, 585, 612 Alveolinella quoyi 545, 564, 565, 566, 609 Alanlordia banyakensis 575, 612 Alveolinidae 323, 347, 396, 446, 448, 449, 544 Alanlordia niasensis 558, 559, 575, 612 Alveolinids 20, 493, 599 Albian– Cenomanian crisis 353 Alveolinoidea 322, 335, 446, 447, 451, 565 algal symbionts, in modern larger foraminifera 101 Alveolophragmium venezuelanum 561 Algeria 273, 276, 364 Alveosepta 204, 219, 231, 234, 235, 245, 249, 257, 276, Allogromiida 45, 161–1 62 280, 292 schematic morphological evolution 49 Alveosepta jaccardi 204, 231, 234, 249, 276 Alpinophragmium 165, 198, 201 Alveospeta 241, 252 Alpinophragmium perforatum 165, 198, 201 Alveovalvulina 560 Alps 157, 348, 352 Alveovalvulinidae 560, 581 Alveolina 20, 21, 22, 26, 30, 33, 87, 320, 323, 396, 397, 398, Alyeovalvulina suteri 560 399, 400, 441, 444, 446, 447, 448, 449, 450, 486, 494, Alzonella 22, 216, 217, 218, 219, 235, 240, 241, 252, 276 502, 505, 508, 511, 517, 544 Alzonella cuvillier 216, 276 alveoles 31, 33, 36 Alzonorbitopsella 216, 221, 232, 261, 278 Alveolina aramaea 502 Alzonorbitopsella arabia 216, 232, 233, 261, 278 Alveolina avellana 397 Americogypsina 435, 438, 541, 575 Alveolina bulloides 399 Americogypsina americana 541 668 Subject Index Americogypsina braziliana 435, 541 Archaecyclus 307, 310, 342, 364 Americogypsina koutsoukosi 541 Archaecyclus cenomaniana 307 Amijiella 215, 217, 232, 235, 240, 241, 244, 248, 264, 280 Archaecyclus midorientalis 364 Amijiella amiji 232, 244, 264, 280 Archaediscida 7, 45, 47, 52, 104 Amijiella slingeri 232 Archaediscidae 49 Amijiellinae 216, 291 Archaediscoidea 100, 104 Ammobaculites 26, 62, 67, 164, 170, 196, 218, 220, 241, archaediscidae 52–5 3 245, 292, 330, 364 lasiodiscidae 53– 54 Ammobaculites gr. edgelli 364 Archaediscus 18, 52, 53, 105, 130, 132, 134, 136 Ammobaculites powersi 67 chamber arrangement of 18 Ammobaculites sarbaicus 62 Archaediscus baschkiricus 53 Ammobaculites sarbaicus subsp. beschevensis 62 Archaediscus complanatus 134 Ammobaculites subgoodlandensis 220 Archaediscus electus 134 Ammobaculites wirzi 170 Archaediscus inflatus 132, 134 Ammobaculites? Dinantii 62 Archaediscus karreri 52, 134 Ammobaculites? pygmaeus 62 Archaediscus spirillinoides 53 Ammochilostoma? triloculina 72 Archaesphaera 48, 49 Ammodiscidae 163, 177, 179, 184, 185, 187, 188 Archaesphaera minima 48 Ammomarginulininae 164, 218, 292 Archaesphaeridae 48, 100 Ammonia batava 40 Archaias 17, 18, 22, 26, 29, 37, 404, 405, 447, 452, 514, Ammonites 283 517, 547, 548, 549, 566, 567, 568, 598, 609, 612 Amphisorus 25, 29, 37, 404, 405, 444, 447, 452, 549, 566, Archaias aduncus 514 567, 596, 598 Archaias angulatus 405, 547, 609 Amphisorus hemprichii 405 Archaias kirkukensis 517, 612 Amphisorus martini 444, 566 Archaias spirans 405, 547 Amphistegina 12, 13, 25, 28, 29, 30, 37, 41, 42, 413, 426, Archiacina 403, 447, 514 441, 443, 444, 455, 471, 476, 487, 526, 536, 553, 564, Archiacina armorica 514 565, 586, 587, 588, 598, 599, 603, 615, 617, 631 Arenobulimina 210, 296, 297, 298, 333, 392, 393, 445 Amphistegina cumingii 413, 553 Arenobulimina labirynthica 296 Amphistegina gibbosa 599 Arenovidalina 170, 172, 177, 180, 195 trimorphic life cycle of 12 Arenovidalina chialingchiangensis 195 Amphistegina lessonii 41, 599, 631 Arenovidalina chialingehian- gensis 170 Amphistegina lobifera 586 Arenovidalinidae 170 Amphistegina quoyii 426 Arnaudiella 310, 312 Amphistegina radiata 42 Arnaudiella grossouvrei 312 Amphisteginidae 426 asexual reproduction Amphoratheka 91 of Hererostegina 12 Amphoratheka iniqua 91 schizonts 14 Anatoliella 392, 394, 502 Asia 348 Anatoliella ozalpiensis 394, 502 Asselian Anchispirocyclina 217, 233, 234, 235, 240, 241, 249, 252, schwagerinidae evolution 102 280, 331, 332 tectonic closure of East European Basin 119 Anchispirocyclina henbesti 217 of Tethyan province 119 Anchispirocyclina lusitanica 234, 240, 280 Assilina 19, 35, 408, 412, 441, 456, 458, 463, 486, 498, 499, Anchispirocyclina neumanni 234 508, 520, 523, 526, 529, 532 Anchispirocyclina praelusitanica 233, 234, 280 Assilina cuvillieri 520 Andersenolina 204, 206, 227, 232, 233, 259 Assilina daviesi 408, 529 Andersenolina alpina 206, 232, 259 Assilina depressa 412 Andersenolina elongata 204, 206, 232, 259 Assilina granulata 526 Andersenolina perconigi 206 Assilina granulosa var. chumbiensis 523 Andrejella 66 Assilina leymeriei 508, 529 Andrejella laxiformis 66 Assilina mamillata 526 Androsina 547 Assilina pomeroli 441 Androsina lucasi 547 Assilina prisca 441 Androsinopsis 547 Assilina subdaviesi 523 Androsinopsis radians 547 Assilina sublaminosa 508 Angola 350 Asterigerinoidea 453, 471, 475, 476, 477, 569, 575 Angotia 422, 437, 438, 478, 479 Asteroachaediscus 53, 136, 137 Angotia aquitanica 437 Asteroachaediscus karreri 137 Angulodiscus 174, 176, 177 Asteroachaediscus pressulus 136 Angulodiscus communis 174 Asterocyclina 416, 456, 469, 470, 488, 532, 534 Annulocibicides 455 Asterocyclina stella 534 Annulosorites 547 Asterocyclina stellata 534 Antarctica 114, 255, 588, 589 Asterophragmina 414, 456, 469, 532 Antigua 631, 637 Asterophragmina pagoda 532 Aphistegina lobifera 41 Asterorbis 310, 317, 344, 347 Archaealveolina 322, 323, 337, 385 Asterorbis rooki 317 Archaealveolina reicheli 385 Astrolepidina 420, 428, 444, 471, 476 Subject Index 669 Ataxella 215 rotaliides 339– 344 Ataxella occitanica 232, 233, 264 textulariides 34, 36, 45, 161, 203, 234, 248, 329, 387, Ataxophragmiidae 3, 163, 296, 297, 347 392, 442, 445, 543, 563, 581, 595 Ataxophragmioidea 329, 332, 333, 334, 581 of fusulinides 14– 36, 45– 49, 73–1 22 Ataxophragmium 296, 297, 298, 333, 353 of Jurassic larger benthic foraminifera 226–2 45 Ataxophragmium (Opertum) incognitum 296 involutinoidea 243– 245 Athecocyclina 414, 468 lituoloidea 240– 242 Atlantic Ocean 348, 352, 354, 498, 601, 604 pfenderinoidea 234– 239 Auloconus 174, 177, 201, 206 textularioidea 245– 246 Auloconus permodiscoides 201 of Neogene larger benthic foraminifera 562– 582 Aulosina 175 letter stages for Neogene of SE Asia 564– 565 Aulotortidae 174 miliolides 565–5 68 Aulotortus 174, 176, 177, 181, 184, 195, 197, 198, 201, 207 rotaliides 568–5 81 Aulotortus friedli 197 textulariines 581– 582 Aulotortus praegaschei 181, 184 of palaeozoic lagenides 105 Aulotortus pragsoides 184, 201 of Paleogene larger benthic foraminifera 439– 484 Aulotortus sinuosus 184, 195, 201 letter stages of Southeast Asia and provincial Auroria 51, 130 biostratigraphy 442–4 45 Auroria singularis 51 miliolides 446–4 52 Auroriidae 51 rotaliides 452– 484 Australia 37, 114, 352, 472, 500, 588, 595, 597, 609, textulariides 445–4 46 612, 631 of shelf limestones 9 Austria 157, 264 of Triassic larger benthic foraminifera 175– 180 Austrocolomia 166, 167, 168, 177 biostratigraphy 175– 177 Austrocolomia marschalli 167 fusulinides 179 Austrotrillina 20, 21, 31, 226, 402, 443, 452, 478, 486, 489, involutides 180– 181 505, 514, 564, 565, 566, 583, 588, 595, 609, 619 lagenides 177– 179 Austrotrillina asmariensis 505, 565, 566, 609 miliolides 179–1 80 Austrotrillina brunni 609 Palaeozoic–T riassic lagenides 177–1 79 Austrotrillina eocaenica 452 Biparietata 93 Austrotrillina howchini 20, 564, 565, 566, 609, 619 Biparietata ampula 93 Austrotrillina paucialveolata 452, 514, 565 Bipertorbis 339, 340 Austrotrillina striata 564, 565, 609 Biplanispira 35, 407, 410, 456, 467, 519, 523, 532 Austrotrillinidae 402 Biplanispira mirabilis 410, 519 Ayalaina 324 Biseriammina 58 Biseriammina uralica 58 Bacinella irregularis 361 Biseriamminidae 49, 58, 100 Baculogypsina 26, 37, 42, 555, 556, 571, 586, 598, 622 Biseriella 58, 131 Baculogypsina baculatus 622 Biseriella parva 131 Baculogypsina floresiana 556 Bisphaera 51 Baculogypsina sphaerulata 42, 622 Bisphaera malevkensis 51 Baculogypsinoides 488, 556 Biwaella 80, 102 Baculogypsinoides spinosus 556 Biwaella omiensis 80 Balkhania 294, 367 Bogushella 61 Balkhania balkhanica 294, 367 Bolkarina 418 Banatia 293 Bolkarina aksarayi 418 Banatia aninensis 293, 361 Boninella 436, 628 Banffella 66 Boninella boninensis 436 Bangladesh 526, 532, 538 Boninella sp. 628 Barattolites 393, 502 Borelis 26, 88, 397, 400, 441, 447, 449, 489, 508, 545, 565, Barattolites trentinarensis 393, 502 566, 583, 599, 609, 619 Barbados 538, 619 Borelis curdica 609 Beedeina 82, 83, 113, 147 Borelis haueri 508, 609 Belgium 383, 538 Borelis matleyi 400 Bermuda 585 Borelis melo 397, 609 Biapertobis 478 Borelis melonoides 396 Biapertorbis 480 Borelis princeps 88 Biarritzina 423, 491, 492, 536, 622 Borelis pulchra 609 Bibradya 72 Borelis pygmaeus 441, 449, 489, 508, 566, 609 Bibradya inflata 72 Boreloides 22, 426 Bigenerina geyeri 58 Boreloides cubensis 426 Bigenerina sumatrana 58 Boreloididae 426 Biokovinidae 223, 289 Borneo 408, 508, 520, 521, 523, 526, 535, 536, 538, 541, Biokovinoidea 223, 245, 288 564, 609, 615, 619, 622, 625, 628, 631, 634, 636, 637 biostratigraphy 9, 96, 175, 226, 328, 439, 442–4 43, 562, Borodinia 559, 575, 619 593, 605, 606 Borodinia septetrionalis 559 of Cretaceous larger benthic foraminifera 328– 344 Borodinia sp. 619 miliolides 335– 339 Bosniella 223, 232, 235, 241, 280 670 Subject Index Bosniella oenensis 223, 232, 280 Cancellina 90, 98 Bostia 217, 235, 240, 241, 252, 280 Capitanian, in diversity of fusulinines during 121 Bostia irregularis 217, 280 Carbonella 61, 62, 139 Boultonia 80, 119 Carbonella spectabilis 62 Boultonia willsi 80 Carboniferous 30 Bozorgniella 553 extinction of foraminifera since 33 Bozorgniella qumiensis 553 number of fusulinide genera through 109 Bradya tergestina 327 parathutamminoidea and robuloidoidea ranges in 100 Bradyina 67, 68, 72, 74, 96, 108–1 09, 117, 137, 139, Carboniferous–P ermian 31 142, 143 boundary 102–1 03 Bradyina nautiliformis 72 Pseudoschwagerina 102 Bradyina rotula 139, 142 Carbonoschwagerina 86, 102 Bradyinidae 72, 96, 100, 108– 109 Caribbean- Colombian Cretaceous Igneous Province wall structure in 96 (CCIP) 353 Brazil 408, 520, 529, 535, 536, 541, 577, 602, 622, 628, Carpenteria 423, 424, 454, 455, 491, 492, 555, 619, 622 634, 636 Carpenteria balaniformis 423, 555 Brevaxina 89, 103 Carpenteria hamiltonensis 424 Broeckina 324, 325 Carpenteria proteiformis 424, 619 Broeckina (Pastrikella) balcanica 325 Carpenteriinae 423, 454, 555 Broeckinella 294, 367 Cassianopapillaria 167 Broeckinella arabica 294, 367 Caudriella 419, 420, 476, 491 Broken Ridge volcanic event 354 Caudriella ospinae 420 Brunsia 55, 131 Caudrina 463 Brunsiella 55, 128, 139, 142 Cayeuxia 229 Brunsiella buskensis 139, 142 Cayeuxia piae 248 Brunsiina 61, 62 Cayeuxia?piae 283 Brunsiina uralica 61 Cellanthus 439, 615 Buccicrenata 220, 232, 233, 234, 241, 264, 330, 331, Cellanthus craticulatus 615 332, 364 Central Atlantic Magmatic Province (CAMP) 191 Buccicrenata hedbergi 220, 364 Central Oman, fusulinine assemblages from Buccicrenata libyca 220 Sakmarian of 120 Buccicrenata primitiva 220, 234, 264 Cercozoan- foraminiferan 5 Buccicrenata subgoodlandensis 220 Cercozoans 4 Bulgaria 188 polyubiquitin genes 5 Bulimina variabilis 296 Ceriopora globulus 425 Bullalveolina 20, 399, 447, 449, 508 Chablaisia 210, 212, 236, 238, 267 Bullalveolina bulloides 508 Chablaisia chablaisensis 238, 267 Burma 134, 136, 145, 146, 147, 151, 193, 194, 195, 196, Chaetetes depressus 159 198, 201, 532 Chalaroschwagerina 86, 119, 149 Chalaroschwagerina stachei 149 calcareous benthic assemblages 176 Chalaroschwagerinainflata 86 calcareous granular fusulinines 176 Chapmanina 23, 422, 437, 444, 479 Calcarina 37, 42, 416, 439, 556, 571, 586, 587, 598, Chapmanina gassinensis 437 615, 622 Chapmaninidae 422, 437, 478, 556 Calcarina gaudichaudii 42, 571 Charentia 223, 231, 289, 290, 364 Calcarina hispida 622 Charentia cuvillieri 364 Calcarina tetraëdra 439 Charentia cuyillieri 289 Calcarina? stellata 416 Charentiidae 223, 289 Calcarinidae 319, 439, 555, 571 Chenella 76 Calcarinids 598 Chenia 79 Calcifolium 134, 159 Chenia kwangsiensis 79 Calcifolium okense 159 Chernobaculites 62 Calcifolium punctatum 134 Chernyshinella 62 California 121, 137, 534 Chernyshinella exelikta 139 Caligellidae 49, 55, 93, 100 Chernyshinella tumulosa 64 Calveziconus 302, 367, 396 Chernyshinellina 62 Calveziconus lecalvezae 302, 367 China 46, 118, 121, 122, 185, 186, 187, 619 Calvezina 92, 178 Chofatellidae 347 Calvezina ottoman 92 Choffatella 22, 24, 26, 219, 234, 241, 249, 250, 264, 289, Camagueyia 435 290, 294, 330, 331, 332, 345, 364 Camagueyia perplexa 435 Choffatella decipiens 219, 250, 290, 364 Camerina laevigata 412 Choffatella peneropliformis 219 Camerina matleyi 439 Choffatella rugoretis 289 Camerina? dickersoni 314 Choffatella tingitana 234, 264 Cameroon 502 Choffatellidae 218, 294 Campanellula 301, 367 Chordoperculinoides 412, 463, 497 Campanellula capuensis 301, 367 Chordoperculinoides bermudezi 463 Canada 189, 253 Chordoperculinoides sahnii 529 Subject Index 671 Chrysalidina 20, 21, 306, 330, 369, 377 Coskinolina 215, 301, 392, 393, 394, 395, 445, 502, Chrysalidina gradata 306, 377 505, 595 Chrysalidinidae 224, 225, 297, 305, 345, 390, 581 Coskinolina balsilliei 395, 502 Chrysalidininae 225, 306, 390 Coskinolina cf. douvillei 395, 502 Chrysalinidae 329, 330 Coskinolina (Coskinon) rajkae 393 Chrysothurammina 49, 130 Coskinolina elongata 393 Chrysothurammina tenuis 49 Coskinolina liburnica 393 Chrysothuramminidae 48, 100 Coskinolina (Meyendorffina) minoricensis 215 Chubbina 321, 322, 338, 361, 364, 369 Coskinolina sunnilandensis 301 Chubbina cardenasensis 361, 369 Coskinolinidae 305, 393 Chubbina jamaicensis 321, 361, 364 Coskinolinoidea 35, 445 Chusenella 84, 86, 99 Coskinolinoides 300, 301, 334, 374, 396 Chusenella ishanensis 86 Coskinolinoides jamaicensis 396 Cibicidella 455 Coskinolinoides texanus 301, 374 Cibicides 342, 453, 454, 455, 470, 471 Coskinolinopsis primaevus 242 Cisalveolina 322, 323, 337, 385 Coskinon 392, 393, 445, 502 Cisalveolina fallar 323 Costayella 62 Cisalveolina lehneri 385 Coxites 288, 358 Cladocoropsis mirabilis 249, 283 Coxites zubairensis 288, 358 Climacammina 57, 105, 107, 132, 133 Coxitinae 288 Climacammina antiqua 133 Crespinina 422, 438, 478, 479 Clypeina jurassica 249, 283 Crespinina kingscotensis 438 Clypeorbis 310, 313, 347 Cretaceous larger benthic foraminifera Codonofusiella 80, 81, 98, 103 biostratigraphy and phylogenetic evolution 328– 344 Codonofusiella (Lantschichites) maslennikovi 81 miliolides of Cretaceous 335– 339 Codonofusiella paradoxica 80 rotaliides of Cretaceous 339– 344 coiling in fusulinines 95 textulariides of Cretaceous 329– 335 Colaniella 49, 93 morphology and taxonomy of 285– 328 Colaniellidae 49, 93, 100 palaeoecology of 344– 348 Colanielloidea 100 palaeogeographic distribution of 348–3 56 Coleiconus 392, 393, 394, 395, 445, 502 Cretaceous- Paleogene crisis 387 Coleiconus christianaensis 395, 502 Cretaceous–P aleogene (K– P) event 355 Columella carpenteriaeformis 423 Cribellopsis 301, 367 Condrustella 62 Cribellopsis neoelongata 367 Conicokurnubia 212, 213, 233, 234, 236, 238, 270 Cribranopsis 66 Conicokurnubia orbitoliniformis 213, 238, 270 Cribranopsis fossa 66 Conicopfenderina 210, 211, 212, 232, 233, 238, 267 Cribrogenerina 57, 58, 105, 107 Conicopfenderina mesojurassica 232, 238, 267 Cribrosphaeroides (Parphia) robusta 51 Conicorbitolina 303, 334, 335, 336, 374 Cribrospira 66 Conicorbitolina cf. cuvillieri 336 Cribrospira panderi 66 Conicorbitolina cobarica 374 Cribrospira pansa 71 Conicorbitolina sp. A 336 Cribrostomum 57, 58, 132 Conicospirillina 175 Cribrostomum inflatum 132 Conilites 62 Cribrostomum textulariforme 58 Conofusiella 84 Cryptoseptida 166, 167, 168, 177 Conoglobigerina 257 Cryptoseptida anatoliensis 167 Conoglobigerna 252 Ctenorbitoides 310, 315, 317 Conorbitoides 310, 315, 316 Ctenorbitoides cardwelli 317 Conorbitoides cristalensis 316 Cuba 343, 381, 383, 454, 502 Conorbitoides kozaryi 316 Cubanina 389, 502 Conulites americana 394 Cubanina alavensis 389, 502 Conulites cooki 433 Cuneolina 297, 298, 299, 333, 334, 371 Cope’s rule 606 Cuneolina cylindrica 371 corals 25, 26, 109, 122, 125, 159, 184, 192, 247, 344, Cuneolina hensoni 371 345, 350, 352, 485, 486, 490, 492, 562, 585, 587, Cuneolina parva 299, 371 591, 594 Cuneolina pavonia 298, 371 Cornuspira 95, 180, 188 Cuneolinella 560 Cornuspira mahajeri 188 Cuneolinella lewisi 560 Cornuspira schlumbergeri 95 Cuneolinidae 298, 347 Cornuspiridae 95, 173 Cuniculi, lineages in 103 Cornuspiroidea 188, 565 Cuniculinella 86, 119 Coronipora 177 Cuniculinella tumida 86 Corrigotubella 66 Cushmania 392, 394, 396, 502 Corrigotubella posneri 66 Cushmania americana 502 Coscinoconus 207 Cuvillierina 431, 432, 538 Coscinoconus alpinus 206 Cuvillierina eocenica 431 Coscinoconus chouberti 207 Cuvillierina vallensis 432 Coscinophragmatoidea 165, 205, 388, 390 Cyclammina 20, 21, 219, 389, 391, 502, 560, 561 672 Subject Index Cyclammina cancellata 391, 560, 561 Daviesininae 432 Cyclammina jaccardi 219 Daxia 223, 245, 288, 289, 361 Cyclammina lituus 219 Daxia cenomana 361 Cyclammina uhligi 389 Daxia minima 361 Cyclamminidae 220, 345, 391, 560, 581 Daxia orbignyi 223, 289 Cycledomia 325 Debarina 288, 290, 364 Cyclocibicides 455 Debarina hahounerensis 290, 364 Cycloclypeidae 414, 464, 552 Deckerella 57, 58, 132 Cycloclypeus 15, 25, 408, 414, 441, 443, 444, 453, 456, 458, Deckerella clavata 58 464, 465, 466, 468, 469, 478, 486, 487, 488, 489, 521, Deckerella quadrata 132 523, 543, 552, 553, 564, 565, 570, 571, 584, 586, 588, Deckerellina 57, 58, 132 593, 595, 598, 615, 617 Deckerellina istiensis 58 Cycloclypeus carpenteri 15, 408, 414, 465, 487, 552, 570, Demirina 289 585, 615 Demirina meridionalis 289, 361 Cycloclypeus cf. guembelianus 465, 466 Dendritina 403, 447, 514, 609 Cycloclypeus cf. orbitoitades 612 Dendritina arbuscula 403 Cycloclypeus droogeri 570 Dendritina cf. rangi 514 Cycloclypeus eidae 408, 466, 570, 615 Dendritina rangi 609 Cycloclypeus guembelianus 570 Devonian– Carboniferous boundary 113 Cycloclypeus indo- pacificus 570 Dhrumella 217, 264 Cycloclypeus indopacificus 466, 552, 615, 617 Dhrumella evoluta 217, 264 Cycloclypeus (Katacycloclypeus) martini 552 Dictyoconella 302, 367 Cycloclypeus koolhoveni 465, 466 Dictyoconella complanata 302, 367 Cycloclypeus mediterraneus 570 Dictyoconella minima 367 Cycloclypeus neglectus var. stellatus 553 Dictyoconinae 301, 345, 394 Cycloclypeus opernoorthi 570 Dictyoconoides 318, 433, 434, 435, 478, 479, 480, 538 Cycloclypeus oppenoorthi 465, 466 Dictyoconoides cooki 538 Cycloclypeus pillaria 615 Dictyoconoides haimei 434 Cycloclypeus posteidae 466, 570 Dictyoconoides kohaticus 434, 538 Cycloclypeus postindopacific 466 Dictyoconus 23, 213, 300, 301, 392, 394, 395, 445, 446, Cyclolina armorica 403 502, 538, 581, 595 Cyclolina dufrenoyi 324 Dictyoconus walnutensis 374 Cyclolinidae 306 Dictyoconus cuvillieri 301 Cycloloculina 454, 455 Dictyoconus egyptiensis 395 Cyclopsina steinmanni 306 Dictyoconus indicus 502 Cyclopsinella 306, 306–3 07, 307 Dictyoconus mosae 301 Cyclopsinellinae 306, 307 Dictyokathina 318, 433, 441, 444, 478, 479, 538 Cycloputeolina 547 Dictyokathina simplex 433, 538 Cyclorbiculina 17, 18, 22, 37, 404, 405, 447, 452, 549, 566, Dicyclina 20, 299, 333, 334, 371, 374 567, 568, 598, 609, 612 Dicyclina qatarensis 371, 374 Cyclorbiculina compressa 612 Dicyclina schlumbergeri 299 Cyclorbiculinoides 405, 447, 452, 514 Dicyclinidae 299 Cyclorbiculinoides jamaicensis 405, 514 Diplosphaera inaequalis 48 Cyclorbitopsella 216, 221, 232, 235, 240, 241, 243, 282 Diplosphaerina 48, 50, 128 Cyclorbitopsella tibetica 221, 232, 243, 282 Diplosphaerina inaequalis 128 Cylammina 364 Diplosphaerina sphaerica 128 Cylindrocolaniella 94 disaster forms 34, 114, 176, 178, 180, 253 Cymbalopora 421, 422, 423, 454, 455, 491 Discocyclina 26, 33, 387, 414, 415, 416, 441, 444, 456, 468, Cymbalopora radiata 421, 423 469, 470, 473, 486, 488, 489, 490, 498, 508, 517, 523, Cymbaloporetta 455 532, 534, 535 Cymbaloporidae 307, 421, 422 Discocyclina anconensis 416 Cymbriaella 215, 235, 240, 241, 273 Discocyclina californica 534 Cymbriaella lorigae 215, 273 Discocyclina dispansa 468, 532 Cyprus 601, 612, 628, 634 Discocyclina perpusilla 415 Czechoslovakia 197 Discocyclina peruviana 532 Discocyclina prattii 468 Dagmarita 58 Discocyclina ranikotensis 532 Dagmarita chanakchiensis 58 Discocyclina sella 534 Dainella 73 Discocyclina sheppardi 468, 532 Dainella? efremoyi 68 Discocyclininae 414, 468, 469 Daixina 86, 119 Discogypsina 424, 470, 471, 491, 536, 612, 628 Daixina ruzhencevi 86 Discogypsina discus 470, 536, 628 Darjella 93 Discogypsina vesicularis 424, 470, 612 Darjella monilis 93 Discolites concentricus 406 Daviesiconus 392, 395, 502 Discospirina 29, 546, 604, 612 Daviesina 432, 441, 444, 479, 508, 538 Discospirina italica 612 Daviesina khatiyahi 432, 538 Discospirinidae 546 Daviesina langhami 508, 538 Distichoplax biserialis 517 Subject Index 673 Dohaia 307, 379 Elphidium sp. 541 Dohaia planata 307, 379 Elphidium williamsoni Doliolina compressa 89 megalospheric form of 40 Doliolina ovalis 89 pseudopod and cytoplasm of 39 Dorothia 297 Elvis taxa 179 Draffania 51, 128, 130 Emeishan Large Igneous Province 124 Draffania biloba 51, 128, 130 end Barremian-e arly Aptian extinctions 351 Dubai 636 End Permian extinction 175, 176, 179, 180, 182, 185 Dukhania 297, 306, 330, 377 End Permian mass extinction, fusulinines 122– 124 Dukhania conica 306, 377 Endochernella 63 Dunbarinella 87 Endospiroplectammina 65 Dunbarinella ervinensis 87, 137, 146, 154, 156 Endostaffella 66 Dunbarula 80 Endoteba 54, 66, 169, 170, 177, 179, 185 Dunbarula mathieui 80 Endoteba controversa 54, 66, 169, 179 Duostomina 165, 177, 185, 194 Endoteba ex gr. bithynica 179 Duostomina alta 185 Endoteba wirzi 185 Duostomina astrofimbriata 185 Endotebanella 169, 170, 177, 179 Duostomina biconvexa 165 Endotebidae 54, 66, 169, 170, 179 Duostominoidea 184 Endothyra 18, 62, 64, 66, 67, 68, 69, 70, 71, 72, 73, 78, 96, Duotaxis 196, 201, 209, 229, 232, 247, 253, 267 134, 142, 143, 170 Duotaxis birmanica 196, 201 Endothyra aljutovica 70 Duotaxis metula 209, 229, 232, 267 Endothyra ammonoides 73 Dutch New Guinea 541, 615, 617, 634 Endothyra bella 68 Dutkevichites 80, 84 Endothyra bowmani 66, 142, 143 Dutkevichites darvasica 80 Endothyra communis 68 Dutkevitchites 102, 103 Endothyra convexa 71 Dyocibicides 454, 455 Endothyra costifera 71 Endothyra discoidea 72 Earlandia 26, 54, 91, 106, 128, 130, 166, 169, 170, 178, 179, Endothyra eostaffelloides 72 195, 229 Endothyra glomiformis 62 Earlandia elegans 128 Endothyra kobeitusana 71 Earlandia perparva 54, 169 Endothyra latispiralis 69, 70 Earlandia tintinniformis 195 Endothyra mirifica 70 Earlandia vulgaris 128, 130 Endothyra omphalota 70 Earlandiida 7, 34, 45, 47, 54, 93, 161, 166, 168, 178, 179 Endothyra panderi 69 Earlandiidae 49, 54, 100, 169 Endothyra parakosyensis 70 Earlandinita 49, 93 Endothyra parya 66 Earlandioidae 49 Endothyra pauciseptata 72 Earlandioidea 93, 100, 161, 178 Endothyra primaeva 64 Early Cimmerian orogeny 250 Endothyra prisca 71 Early Toarcian transgression 254– 255 Endothyra (Rectoendothyra) donbassica 71 East Africa 580, 598 Endothyra staffelliformis 78 East Asia 185, 187, 593 Endothyra tuberculata 72 East European Basin Endothyra? banffensis 66 Asselian tectonic closure of 115 Endothyra? chomatica 73 Carboniferous– Permian boundary 101 Endothyra? Krainica 64 distribution of fusulinines in 114 Endothyranella 67, 138, 169, 196 Montiparus montiparus in 97 Endothyranella lwcaeliensis 169 East Indies 442, 480 Endothyranopsis 67, 138, 142, 143, 144 Eclusia 307, 379 Endothyranopsis aff. E. macrus 142 Eclusia moutyi 307, 379 Endothyranopsis crassa 138, 142, 143, 144 Edomia 325, 326, 379 Endothyranopsis pechorica 142 Edomia iranica 325 Endothyridae 49, 66, 72, 100, 170, 179 Edomia reicheli 326, 379 Endothyrids, during Visean 114–1 15 Egypt 358, 371, 408, 514, 526, 538 Endothyrina? gracilis 70 Elazigella 403, 514 Endothyrida 7, 34, 45, 47, 65, 108, 114, 162, 169, 179, 203 Elazigella altineri 403, 514 Endothyroidea 65– 72, 100, 161, 169, 177, 185 Elbanaia 63 bradyinidae 72 Elenella 50, 130 endothyridae 66– 72 Elergella 66 evolution during Palaeozoic 67 Elergella simakoyi 66 Endotriada 26, 170, 177, 179 Elphidiella multiscissurata 358 Endotriada tyrrhenica 170 Elphidiidae 3, 439 Endotriadella 170, 177, 179 Elphidium 26, 39, 40, 108, 439, 479, 541, 586, 615 Endotriadidae 169 Elphidium craticulatus 615 England 128, 130, 131, 132, 133, 134, 136, 138, 139, 140, Elphidium excavatum 142, 143, 144, 147, 159, 253, 364, 371, 529 equatorial section of 39, 40 Eoannularia 419, 454, 455, 491, 538 nuclear envelope of nucleus of 39 Eoannularia eocenica 419, 538 674 Subject Index Eoannularidae 419 Eulepidina undosa 631, 637 Eoconuloides 426, 427, 472, 475, 476, 477 Eulepidina undosa laramblaensis 631, 637 Eoconuloides lopeztrigoi 475, 477 Eulinderina 427, 471, 472, 473, 475, 476, 477, 478, Eoconuloides wellsi 426 536, 541 Eocristellaria 92 Eulinderina guayabalensis 475, 477, 478 Eocristellaria permica 92 Eulinderina sp. 473, 536, 541 Eoendothyra 68 Europe 37, 114, 117, 119, 185, 186, 187, 188, 236, 247, Eoendothyranopsis 68 248, 249, 254, 255, 257, 258, 321, 347, 348, 352, 442, Eofabiania 422 444, 480, 580, 600, 601 Eofabiania grahami 422 Euxinita 68 Eoforschia 63 Everticyclammina 20, 21, 26, 164, 220, 229, 231, 232, 234, Eofusulina 83, 84 235, 241, 243, 245, 248, 249, 257, 264, 280, 290, 330, Eofusulina (Paraeofusulina) trianguliformis 84 331, 332, 344, 361 Eolasiodiscus 53 Everticyclammina eccentrica 361 Eolasiodiscus donbassicus 53 Everticyclammina greigi 231, 280, 331 Eoophthalmidium 171, 193 Everticyclammina hensoni 220 Eoophthalmidium tricki 193 Everticyclammina kelleri 280, 330 Eoparafusulina 87, 103, 110 Everticyclammina praekelleri 234, 264 in Early Permian 103, 110 Everticyclammina praevirguliana 220, 229, 233, 243, palaeoecology of 111 280, 330 Eoparastaffella 76 biozone 229, 232, 234 Eoparastaffella simplex 116 Everticyclammina praevirgulina 164 Eopolydiexodina 87, 99 Everticyclammina virguliana 231, 234, 245, 280 Eoquasiendothyra 68 Everticyclamminidae 219, 220, 291, 347 Eorupertia 213, 267, 423, 491, 536, 538 evolute test 19 Eorupertia incrassata var. laevis 538 evolutionary diversity 4, 45 Eorupertia neocomiensis 213, 267 exoskeletal elements 20 Eorupertia sp. 536 extinction Eoschubertella 81, 118 events affecting palaeontology of larger benthic Eostaffella 73, 76, 98, 102, 109, 117, 130, 142, 145 foraminifera 35– 37 Eostaffella arcuata 138 of foraminifera since Carboniferous 33 Eostaffella mediocris 78 of fusulinines during Guadalupian 103 Eostaffella mosquensis 134, 144 Eostaffella ornata 142 Fabiania 20, 422, 491, 538 Eostaffella radiata 142 Fabiania cassis 538 Eostaffella (Seminovella) elegantula 73 Fabularia 399, 400, 401, 447, 451, 514 Eostaffelloides 77 Fabularia discolites 399 Eostaffelloides orientalis 77 Fabularia discolithus 514 Eotextularia 65 Fabularia hanzawai 514 Eotournayella 61, 62, 63, 108 Fabulariidae 320, 323, 338, 399, 401, 446, 449, 451 Eotuberitina 51, 128 Fallotella 392, 394, 396, 444, 502 Eotuberitina cornuta 128 Fallotella alayensis 396 Eotuberitina fornicata 128 Fallotia 325, 326, 338 Eotuberitina reitlingerae 51, 128 Fallotia jacquoti 325 Eovolutina 51, 130 Falsurgonina 300, 301 Eovolutina elementa 51 Falsurgonina pileola 301 Eovolutinidae 51 Far East 37, 38, 188, 387, 428, 437, 442, 444, 452, 480, Eowedekindellina 83 483, 569, 570, 571, 573, 578, 585, 606, 625 Eowedekindellina fusiformis 83 Fasciolites 396 Epiannularia 419 Ferayina 422, 438, 479 Epiannularia pollonaisae 419 Ferayina coralliformis 438 Ethiopia 498, 499, 601 Feurtillia 219, 245, 294, 361 Eukaryotes, phylogeny of 4 Feurtillia frequens 219, 294, 361 Eulepidina 33, 387, 420, 428, 429, 441, 444, 471, 475, 476, Fiji 615, 617, 622 477, 478, 523, 526, 535, 536, 541, 564, 582, 593, 600, Fissoelphidium 319, 339, 340, 358 619, 631, 634, 637 Fissoelphidium operculiferum 319, 358 Eulepidina andrewsiana 634 Florida 523, 529, 554, 634 Eulepidina badjirraensis 634, 637 Flosculinella 17, 18, 443, 449, 545, 564, 565, 566, 583, Eulepidina dilatata 429, 634 595, 609 Eulepidina dilatate 535 Flosculinella bontangensis 449, 565, 566, 583, 595, 609 Eulepidina eodilatata 637 Flosculinella globulosa 609 Eulepidina ephippioides 429, 535, 541, 619 Flosculinella reicheli 565, 583 Eulepidina favosa 477, 478, 637 foraminifera Eulepidina formosa 637 classification of 4 Eulepidina formosides 634 cretaceous agglutinated 21 Eulepidina inaequalis 634 larger benthic see larger benthic foraminifera Eulepidina papuaensis 523 living and fossil 1 Eulepidina parkinsonia 637 planktonic 6 Subject Index 675 polyubiquitin genes 5 schematic features of advanced forms of 85 shell of 21 stratigraphic occurrence of stages 84 foraminiferal test 33 of Triassic 179 forereef 441, 605 views of schematic 22 Forschia 63 Fusulinoidea 73–7 6 Forschia cf. subangulata 139 evolution 96, 101 Forschiella 63 fusulinidae 75, 82 Forschiella prisca 63 loeblichiidae 73 France 9, 189, 204, 264, 267, 270, 273, 276, 280, 290, 361, morphological features of 74 364, 367, 371, 374, 377, 379, 380, 383, 384, 385, 398, Neoschwagerinidae 90 408, 415, 429, 468, 505, 508, 511, 514, 517, 523, 526, Ozawainellidae 76 529, 532, 534, 536, 538, 622, 625, 628 schematic palaeoecological distribution of 111 Franklinian corridor 117, 121 schubertellidae 74– 75, 77, 80 Freixialina 223, 233, 234, 245, 264, 289 schwagerinidae 85–8 9 Freixialina atlasica 234 shells, structural features 98 Freixialina planispiralis 223, 233, 234, 264 staffellidae 79 French Guyana 191 verbeekinidae 85, 89 Friexialina 288 wall structures 74 Frondicularia woodwardi 94 development in Late Carboniferous and Fusarchaias 548, 567 Permian 96, 98 Fusarchaias bermudezi 548 Fusiella 81, 110, 145 Galea tollmanni 173 Fusiella typica 81 Galeanella 173, 196 fusiform fusulinides 33 Gallowaiina 81 fusiform test Gallowaiina meitiensis 81 and convergence of similar shapes of test 32 Gandinella 163, 177, 193 Flosculinella 17, 18 Gandinella apenninica 163 Fusulina and Alveolina 29– 30 Gandinella falsofriedli 193 Fusulina 79, 82, 83, 98, 102, 110, 113, 117, 119, 157 Geinitzella (Lunucammina) permiana 94 assemblages in Wolfcamp beds 137 Geinitzina 94, 166, 168, 178 schubertellidae and schwagerinidae arose from 96, 97 Geinitzinidae 49, 94, 100 Fusulina cylindrica 83, 137, 150 Geinitzinita 167 Fusulina gracilis 87 Geinitzinita oberhauseri 167 Fusulina longissima 84 Geinitzinoidea 100 Fusulina (Neoschwagerina) primigena 90 Gendrotella 288, 289, 361 Fusulina obsoleta 84 Gendrotella rugoretis 361 Fusulina prima 145 Georgella dytica 63 Fusulina sphaerica 79 Germany 38, 196 Fusulina triangula 83 Gibraltar 203, 229, 267, 270, 273, 280, 283, 596, 597 Fusulina verbeeki 89 Gigasbia 54 Fusulinella 16, 77, 79, 83, 84, 85, 97, 99, 102, 110, 113, 117, Gigasbia gigas 54 118, 119 Globigerina 26, 532 Fusulinella angulata 77 Globispiroplectammina 58 Fusulinella bocki 83 Globispiroplectammina mameti 58 Fusulinella euthusepta 85 Globivalvulina 59 Fusulinella girtyi 83 Globivalvulina parva 58 Fusulinella struvii 79 Globochernella 69 fusulinid test 98 Globochernella braibanti 69 Fusulinida 7, 14, 34, 45, 46, 47, 73, 73– 76, 74, 82, 96, 96– Globoendothyra 69 97, 98, 114, 161, 175, 178 Globoendothyra pseudoglobulus 69 morphology and taxonomy of Palaeozoic 48– 96 Globoreticulina 399, 447, 449, 508 schematic morphological evolution 49 Globoreticulina iranica 399, 508 Fusulinidae 8, 75, 82, 97, 98, 99, 101 Globotetrataxis 59 evolved from ozawainellidae in Middle Globotextulariidae 389, 581 Carboniferous 99 Globuligerina 26, 252, 257 Late Mississippian 101 Glomalveolina 399, 441, 446, 447, 448, 449, 450, 502, fusulinides 179 505, 514 biostratigraphy and phylogenetic evolution of 96–1 03 Glomalveolina dachelensis 514 calcareous granular 32 Glomalveolina delicatissima 502 coiling in 95 Glomalveolina levis 441, 450 End Permian mass extinction 122– 124 Glomalveolina primaeva 441, 450 genera as function of time 112 Glomodiscus 53 genera through Carboniferous and Permian 114 Glomodiscus biarmicus 53 Marginopora symbionts in keriothecal walls of 110 Glomospira 55, 163, 177, 194, 195, 196, 229, 247 microgranular walls 15 Glomospira ammodiscoidea 55 palaeoecology of 104–1 07 Glomospira cf. tenuifistula 194 palaeogeographic distribution of 111– 125 Glomospira meandrospiroides 194, 195 rhizopods 24 Glomospira umbilicata 163 676 Subject Index Glomospiranella 61, 62, 63 Haurania (Platyhaurania) subcompressa 215 Glomospiranella Asiatica 63 Haurania sp. 232 Glomospiranella finitima 63 Hauraniidae 214, 215, 291 Glomospirella 134, 163, 175, 177, 181, 193, 196, Hauraniinae 214 198, 201 Haymanella 403, 447, 452, 514 Glomospirella friedli 175 Haymanella paleocenica 403, 514 Glomospirella irregularis 193, 196, 198, 201 Haynesina 39 Glomospirella paseudopulchra 134 Helenalveolina 323, 338 Glomospiroides 63 Helenalveolina tappanae 323 Glomospiroides fursenkoi 63 Helicolepidina 427, 471, 472, 475, 476, 477, 541, 600 Glomotrocholina 53 Helicolepidina nortoni 475 Glomotrocholina pojarkovi 53 Helicolepidina spiralis 475, 477, 541, 600 Glyphostomella 72, 111 Helicolepidina spiralis trinitatis 477 Goesella miocenica 560 Helicolepidina trinitatis 475 Gondwana 102, 112, 113, 114, 119, 182, 183, 348, 352 Helicolepidininae 426, 471 Granuliferella 69 Helicolepidinoides 472, 476 Granuliferella granulosa 69 Helicorbitoides 310, 314, 343 Granuliferelloides 69 Helicostegina 427, 471, 472, 473, 475, 476, 477, 541 Granuliferelloides jasperensis 69 Helicostegina dimorpha 427, 475, 477 Greece 121, 197, 198, 283, 379, 429, 541, 634, 637 Helicostegina gyralis 473, 475, 477, 541 Grillina 166, 167, 168, 177, 179, 197 Helicostegina paucipira 475 Griphoporella sp. 184, 193 Helicostegina paucispira 477 Grovesella 81, 102 Helicostegina polygyralis 475, 477 Grovesella tabasensis 81 Helicostegina soldadensis 427, 475, 477 Grzybowskia 411, 456, 465, 520 Helicosteginoides 603 Grzybowskia multifida 520 Helicosteginopsis 427, 471, 476, 541 Guadalupian 35, 57, 78, 80, 81, 87, 88, 90, 92, 93, 95, Helicosteginopsis soldadensis 541 97, 98, 100, 103, 106, 112, 121, 125, 151, 156, 166, Hellenocyclina 310, 311 172, 179 Hellenocyclina beotica 311 Guinea 191, 545, 580, 585, 609, 615, 619, 625, 631, Hellenocyclinidae 311 634, 636 Hemiarchaediscus 53, 136 Gulf of Aqaba 487, 587 Hemiarchaediscus angulatus 136 Gulf States 251 Hemiarchaediscus compressus 136 Gunteria 422, 491 Hemiarchaediscus planus 53 Gunteria floridana 422 Hemicyclammina 290, 291, 292, 331, 332, 364 Guppyella 560 Hemicyclammina sigali 291, 292, 364 Gutnicella 215, 232, 235, 240, 241, 273, 276 Hemicyclammina whitei 331, 364 Gutnicella bizonorum 232, 276 Hemifusulina 83, 110 Gutnicella cayeuxi 232, 233, 273 Hemifusulina bocki 83 Gutnicella minoricensis 232, 276 Hemigordiidae 95 Gutnicella oxitanica 276 Hemigordiopsidae 95, 173 Gypsina 421, 424, 455, 470, 471, 491, 536, 559, 619 Hemigordiopsis 95, 172 Gypsina marianensis 421 Hemigordiopsis renzi 95 Gypsina mastelensis 470 Hemigordius 53, 95, 172, 180 Gypsina vesicularis var. squamiformis 559 Hemigordius ulmeri 53 Gyroconulina 293, 358 Heterillina 402, 514 Gyroconulina columellifera 293, 358 Heterillina guespellensis 402 Heterillina hensoni 514 Haddonia 390, 502, 587 Heterocyclina 458, 553, 587, 598 Haddonia torresiensis 390, 502 Heterocyclina tuberculata 587 Haddoniidae 390 Heterospira mirabilis 407 Halenia 65 Heterostegina 2, 4, 12, 14, 19, 24, 26, 30, 41, 42, 408, 411, Halenia legrandi 65 412, 413, 414, 441, 444, 458, 463, 464, 465, 486, 487, Halimeda 443, 565 508, 520, 521, 552, 553, 564, 569, 570, 584, 586, 588, Halkyardia 422, 423, 441, 444, 454, 455, 491, 538 598, 604, 628 Halophila 586 Heterostegina depressa 2, 4, 6, 7, 12, 14, 30, 41, 42, 411, Haplophragmella 66, 67, 69, 96 456, 570 Haplophragmina 69 asexual reproduction of 12 Haplophragmina kashkirica 69 shell 4 Haplophragmioidea 288 Heterostegina gracilis 441 Haplophragmoides 288 Heterostegina (Heterostegina) 408, 411, 456, 520 Haplophragmoides sp. 259 Heterostegina (Hetrostegina) 521 Hauerina occidentalis 402 Heterostegina luberculata 553 Hauerinidae 402 Heterostegina operculinoides 413 Haurania 26, 215, 216, 217, 229, 232, 234, 235, 240, 241, Heterostegina sensu lato 441, 444 244, 248, 264, 273, 276 Heterostegina sensu stricto 465, 569 Haurania amiji 217 Heterostegina (Vlerkina) 408, 412, 456, 458, 465, 508, 520, Haurania deserta 215, 232, 233, 244, 264, 273, 276 521, 564, 569, 628 Subject Index 677 Heterostegina (Vlerkina) borneensis 408, 412, 441, 508, Isolepidina 475 520, 521, 564, 628 Israel 251, 259, 264, 270, 280, 358, 379 Heterostegina (Vlerkinella) 412 Italy 152, 154, 158, 179, 193, 196, 197, 198, 203, 208, Heterostegininae 411, 463, 464, 570 259, 273, 283, 361, 367, 371, 377, 385, 413, 502, 521, Heterosteginoides 557, 571 529, 553 Heterostgina (Vlerkina) borneensis 619 Ivanovellidae 50, 100 Hexagonocyclina 415, 469, 534 Hexagonocyclina cristensis 534 Jamaica 361, 364, 395, 454, 502, 508, 511, 514, 517, 532 Historbitoides 310, 316 Janischewskina 72 Holkeria 69 Janischewskina typica 72 Holland 383, 384, 631 Japan 37, 42, 154, 179, 249, 273, 580, 585 Homotrema 559, 575, 585, 619 Jarvisella 502, 560 Homotrema rubra 575, 585, 619 Jarvisella karamatensis 502 Homotrema rubrum 575, 619 Jaryisella karamatensis 560 Homotrematidae 425, 559 Java 345, 350, 351, 352, 353, 442, 517, 521, 523, 535, 541, Hottingerina 403, 514 609, 612, 615, 617, 619, 622, 628, 631, 637 Hottingerina lukasi 403, 514 Jurassic larger benthic foraminifera 203–2 58 Hottingerita 292, 331, 332, 358 biostratigraphy and phylogenetic evolution 226–2 45 Hottingerita complanata 358 involutinoidea of Jurassic 243– 245 Howchinella 94, 178 lituoloidea of Jurassic 240– 242 Howchinia 49, 54, 104, 106, 128, 133, 140 pfenderinoidea of Jurassic 234–2 39 Howchinia bradyana 106, 133, 140 textularioidea of Jurassic 245– 246 Howchinia nuda 140 morphology and taxonomy of 204– 226 Hoynella 173, 197 palaeoecology of 247–2 49 Hoynella gr. sinensis 197 palaeogeographic distribution 250– 258 Hoynella sinensis 173 Jutland Flood Basalt event 122 Hoynellidae 173 Hubeiella 78 K- P event 485– 486 Hubeiella simplex 78 K-s election 606 Hypermorphosis 606 K-s trategists 27, 30 K-s trategy 485 Ichtyolariidae 92, 167 Kahlerina 80 Idalina 338, 401, 451 Kahlerina pachytheca 80 Ijdranella 217, 235, 240, 241, 280 Kamurana 173, 188 Ijdranella altasica 280 Kamurana bronnimanni 173, 188 Ijdranella atlasica 217 Kamurana chatalovi 188 Ilerdorbinae 307 Kanakaia 549 India 114, 348, 352, 385, 408, 434, 486, 502, 505, 508, 517, Kanakaia marianensis 549 519, 526, 529, 532, 534, 538, 612 Kangvarella 78 Indonesia 321, 417, 472, 480, 488, 514, 519, 520, 521, 609, Kangvarella irregularis 78 612, 615, 622, 628 Karaburunia 171, 193 involute test 19 Karaburunia rendeli 171, 193 involutides of Triassic 180 Karaisella 223, 280 Involutina 26, 67, 175, 176, 177, 181, 185, 201, 206, 207, Karaisella uzbekistanica 223, 280 208, 227, 229, 247, 253, 259 Karoo– Ferrar flood basalt event 255– 256 Involutina communis 201 Karsella 392, 396, 514 Involutina conica 175, 206, 207 Karsella hottingeri 396, 514 Involutina crassa 67 Kastamonina 217, 234, 235, 240, 241, 273 Involutina jonesi 175, 206 Kastamonina abanica 217, 273 Involutina liassica 208, 229, 259 Katacycloclypeus 552, 564, 571, 586, 588, 595, 615, Involutina tenuis 201 617, 619 Involutina tumida 201 Katacycloclypeus annulatus 564, 571, 615, 617, 619 Involutinida 45, 161– 162, 174– 175, 177, 180, 184– 185, Katacycloclypeus martini 571, 617 188, 203, 204– 205, 286– 287 Katacycloclypeus sp. 552, 564 evolutionary lineages of 176 Kathina 318, 433, 479, 538 Involutinidae 175, 206 Kathina delseata 318 Involutinoidea 34, 184, 205, 227, 228, 234, 247 Kathina delseota 433, 538 of Jurassic 243– 245 Kathina major 538 Iran 131, 251, 273, 296, 358, 361, 364, 367, 369, 371, Kathininae 433 374, 377, 381, 383, 432, 452, 505, 508, 514, 529, 580, Kayseriella 402 609, 634 Kayseriella decastroi 402 Iranica slingeri 244, 264 Kenya 417, 609, 619 Iraq 147, 251, 264, 273, 280, 357, 358, 364, 367, 371, 374, Keramosphaera 551, 585 377, 380, 381, 383, 451, 502, 505, 508, 514, 517, 538, Keramosphaera murrayi 551 541, 609, 612 Keramosphaeridae 327, 549 Iraqia 302, 374 Keramosphaerina 327 Iraqia simplex 302, 374 Kerguelen plateau 352 Irunites 559 “keriotheca” 15 678 Subject Index Kilianina 26, 212, 213, 214, 232, 234, 237, 239, 252, 264, stressful environmental conditions 30, 31 267, 273 evolutionary history of 30 Kilianina blancheti 213, 232, 239, 267, 273 convergent evolutionary trends 32 Kilianina lata 233, 234, 239, 264 morphological evolution 31 Kilianina prebancheti 233 stolon system 33 Kilianina preblancheti 232 exoskeletal alveoles and endoskeleton 20 Kilianina rahonensis 239 extinction events affecting 35–3 7 Kimmeridgian crisis 249 factors affecting shape of 15 Klubonibelia 69 geological range of suborders of 10 Klubonibelia immanis 69 habitat 9 Korobkovella 424, 491 palaeontology of 34 Koskinobigenerina 56 study of living and extant forms 8 Koskinobigenerina breviseptata 56 subdivision of 14, 15 Koskinotextularia 56, 132 suborders 6 Koskinotextularia cribriformis 56 symbionts 7, 8, 12, 20, 23, 24, 26, 28, 33 Kurdistan, Polydiexodina from 147 three-l ayered 8 Kurnubia 211, 212, 213, 232, 233, 234, 236, 237, 238, trimorphic life cycle of 12– 14 259, 270 wall structures of 14, 15 Kurnubia jurassica 232, 237, 238, 259, 270 Larrazetia 325, 549 Kurnubia palastiniensis 213, 232, 233, 238, 270 Lasiodiscidae 49, 53, 100 Kurnubia wellingsi 237, 238, 270 Lasiodiscus 49, 54 Kurnubiinae 213, 238, 293 Lasiodiscus granifer 54 Kuwait 273, 357, 364 Lasiotrochus 54 Kwantoella 81 Lasiotrochus tatoiensis 54 Kwantoella fujimotoi 81 Late Carboniferous fusulinine at Tethyan realm during 119 Labyrinthina 221, 222, 234, 235, 240, 241, 270 schubertellidae and schwagerinidae evolved from Labyrinthina mirabilis 222, 270 fusulina 97, 98 Labyrinthininae 222 schwagerinidae of 97, 98 Lacazina 22, 24, 320, 322, 338, 346, 385, 400, 401, 447, Triticites 85 451, 514, 538 Late Devonian Lacazina sp. 385, 514, 538 palaeogeographic and tectonic reconstruction of 113 Lacazina wichmanni 400 Late Mississippian Lacazinella 400, 401, 441, 447, 451, 514 fusulinidae 101 Lacazinella wichmanni 514 Late Pennsylvanian Ladoronia 559 palaeogeographic and tectonic reconstruction of 119 Laevipeneroplis 404, 547 Late Permian rocks 179 Laffieina bibensis 441 Late Tournaisian Laffiteina vanbellini 432 evolution of tournayellidae 61 Laffitteina 319, 383, 431, 535, 538 Latiendothyra 69 Laffitteina bibensis 319, 431 Latiendothyranopsis 70 Laffitteina vanbelleni 383, 535, 538 Laurasia 182, 183 Laffitteininae 431 Laxoendothyra 70 Laffteina vallensis 431 Laxoseptabrunsiina 63 Lagenida 7, 14, 45, 47, 49, 90–9 4, 105, 161, 165, 167, 175, Laxoseptabrunsiina valuzierensis 63 178, 204 Lebanon 194, 220, 259, 264, 267, 270, 273, 276, 283, 377, biostratigraphy and phylogenetic evolution 106 508, 612, 615 schematic morphological evolution 49 Lenticulites complanatus 413 lagenides, phylogenetic development of 166 Lepidocyclina 8, 9, 11, 22, 23, 26, 387, 420, 427, 428, 430, lagenides of Palaeozoic–T riassic 177–1 79 437, 443, 444, 471, 472, 473, 474, 475, 476, 477, 478, Lamarmorella 326 486, 521, 523, 535, 558, 564, 565, 572, 575, 577, 578, Lamarmorella sarda 326 579, 580, 581, 582, 583, 588, 593, 600, 601, 602, 615, Lamelliconus 174, 177, 181, 184, 188, 198 619, 625, 628, 631, 634, 636 Lamelliconus cordevolicus 184 pillars in 22 Lamelliconus gr. biconvexus-v entroplanus 184 Lepidocyclina acuta 631 Lamelliconus multispirus 184, 198 Lepidocyclina angulosa quadrata 625 Lamelliconus procerus 184 Lepidocyclina asterodisca 428 Lantschichites 81 Lepidocyclina banneri 625 larger benthic foraminifera 6 Lepidocyclina canellei 631 as biostratigraphic markers 11 Lepidocyclina delicata 625 canal system 24 Lepidocyclina gigantea 580, 619 chamberlet cycle growth 23 Lepidocyclina (Helicolepidina) spiralis 427 dimorphic forms 12– 13 Lepidocyclina isolepidinoides 631 ecology of 25 Lepidocyclina (Lepidocyclina) 428, 444, 477, 478, 497, abundance and diversity 30 523, 535, 575, 576, 577, 578, 579, 600, 601, 602, 628, food particles 25 634, 636 life spans 27 Lepidocyclina (Lepidocyclina) ariana 477 living distribution patterns of 28 Lepidocyclina (Lepidocyclina) boetonensis 634 Subject Index 679 Lepidocyclina (Lepidocyclina) canellei 535, 577, 601, 636 Lepidocyclininae 420, 427, 471, 551 Lepidocyclina (Lepidocyclina) isolepidinoides 581, 628, 636 Lepidorbitoides 11, 310, 311, 313, 314, 341, 342, 343, Lepidocyclina (Lepidocyclina) mauretanica 477 347, 358 Lepidocyclina (Lepidocyclina) ocalana 523 Lepidorbitoides minor 358 Lepidocyclina (Lepidocyclina) praemarginata 578 Lepidorbitoides sp. 358 Lepidocyclina (Lepidocyclina) proteiformis 477, 478 Lepidorbitoididae 310, 312, 430 Lepidocyclina (Lepidocyclina) pustulosa 477, 535, 600 Lepidosemicyclina 557, 558, 564, 573, 603, 619, 622, 625 Lepidocyclina (Lepidocyclina) rdouvillei 600 Lepidosemicyclina (Nephrolepidina) sumatrensis 619 Lepidocyclina (Lepidocyclina) yurnagunensis 577, 578, 601 Lepidosemicyclina sp. 564, 619, 625 Lepidocyclina mantelli 631 Lepidosemicyclina thecideaeformis 622 Lepidocyclina murrayana 615, 631, 634 Letter Stages, subdivision of Indo- Pacific Lepidocyclina (Nephrolepidina) 428, 444, 473–4 75, 478, Cenozoic 442–4 45 497, 523, 535, 541, 576, 577, 578, 579, 580, 581, 582, Levantinella 222, 235, 237, 241, 264, 273 600, 601, 602, 615, 619, 625, 628, 631, 634, 636 Levantinella egyptiensis 264, 273 Lepidocyclina (Nephrolepidina) aff. epigona 631 Levantinellinae 222 Lepidocyclina (Nephrolepidina) angulosa 473, 579, 581, Libya 358, 361, 364, 384, 511, 521, 529, 532 625, 634 Liebusella 389, 502, 595 Lepidocyclina (Nephrolepidina) aquitaniae 579, 631 Liebusella soldanii 502 Lepidocyclina (Nephrolepidina) batesfordensis 631 Liebusellidae 581 Lepidocyclina (Nephrolepidina) bikiniensis 577, 634 Lilliput effect 30, 185, 253, 387 Lepidocyclina (Nephrolepidina) bikiniensis Linderina 419, 420, 454, 455, 491, 508, 538 pumilipapilla 631 Linderina brugesi 420, 538 Lepidocyclina (Nephrolepidina) braziliana 477, 577, Linderina buranensis 538 601, 634 Linderina burgesi 508 Lepidocyclina (Nephrolepidina) brouweri 625, 628, Linderina floridensis 538 631, 634 Linderinidae 419, 420 Lepidocyclina (Nephrolepidina) chaperi 535 Lipinella 59 Lepidocyclina (Nephrolepidina) ferreroi 489, 580, 612, 615, Lipinella notata 59 619, 631 Lituola 164, 218, 221, 240, 241, 243, 288, 293, 331, 332, Lepidocyclina (Nephrolepidina) gibbosa 634 364, 389, 587 Lepidocyclina (Nephrolepidina) inflata 631 Lituola compressa 221 Lepidocyclina (Nephrolepidina) marginata 634 Lituola gr camerata 331 Lepidocyclina (Nephrolepidina) martini 581, 631 Lituola nautiloidea 389 Lepidocyclina (Nephrolepidina) morgani 474, 581, 625 Lituola nautiloidea var. soldanii 389 Lepidocyclina (Nephrolepidina) nephrolepidinoides 628 Lituola nautiloides 364 Lepidocyclina (Nephrolepidina) pilifera 636 Lituola obscura 364 Lepidocyclina (Nephrolepidina) praemarginata 541, 581, Lituolidae 3, 164, 218, 292, 391 600, 601 lituolids 35 Lepidocyclina (Nephrolepidina) praetournoueri 579 Lituolinae 164, 218, 292 Lepidocyclina (Nephrolepidina) rutteni 579, 580, 581, 631 Lituolipora 222, 224, 282 Lepidocyclina (Nephrolepidina) rutteni quadrata 579, Lituolipora polymorpha 224, 282 580, 631 Lituoliporidae 224 Lepidocyclina (Nephrolepidina) soebandii 631 Lituolites nautiloidea 164, 218, 293 Lepidocyclina (Nephrolepidina) sondaica 625 Lituoloidea 214, 227, 228, 234, 235, 240, 241, 252, Lepidocyclina (Nephrolepidina) sp. aff. gibbosa 634 288, 329 Lepidocyclina (Nephrolepidina) sumatrensis 473, 581, 619, Lituoloidea of Jurassic 240–2 42 634, 636 Lituonella 210, 390, 393, 394, 595 Lepidocyclina (Nephrolepidina) suwanneensis 625 Lituonella makarskae 390 Lepidocyclina (Nephrolepidina) tournoueri 478, 579, 581, Lituonella mesojurassica 210 582, 634 Lituonella roberti 393 Lepidocyclina (Nephrolepidina) transiens 579 Lituonelloides 305, 364 Lepidocyclina (Nephrolepidina) veracruziana 523 Lituonelloides compressus 305, 364 Lepidocyclina (Nephrolepidina) yurnagunensis 477 Lituosepta 221, 229, 232, 233, 235, 240, 241, 248, 280 Lepidocyclina (Nephrolepidina)? rutteni 475 Lituosepta recoarensis 221, 229, 232, 233, 280 Lepidocyclina ngampelensis 634 Lituosepta recoarensis spp. biozone 232 Lepidocyclina omphalus 631 Lituotuba? gravata 55 Lepidocyclina (Polylepidina) chiapasensis 427 Lituotubella 63 Lepidocyclina (polylepidnal) punjabensis 430 Lituotubella glomospiroides 63 Lepidocyclina sensu lato 428, 565, 575, 577, 600, 602 Lockhara condi 441 Lepidocyclina sensu stricto 476 Lockhartia 318, 434, 441, 444, 478, 479, 480, 538 Lepidocyclina s.l. 478, 580 Lockhartia cf. newboldi 538 Lepidocyclina sp. 8, 23, 428, 444, 521, 619, 631 Lockhartia conditi 538 Lepidocyclina (Tribliolepidina) 579, 580 Lockhartia haimei 434, 538 Lepidocyclina verbeeki 631 Lockhartia sp. 538 Lepidocyclina volucris 634 Lockhartiinae 433 Lepidocyclinidae 424, 426, 453, 471, 473, 475, 551, 578, Loeblichia 66, 67, 68, 73, 96, 97, 114, 136, 137, 143 579, 580, 581, 599 Loeblichiidae 73, 96, 97, 98, 226 Lepidocyclinids 36, 497 Loftusia 15, 216, 295, 296, 330, 331, 332, 361, 367, 369 680 Subject Index Loftusia coxi 367 Mediopsis 70 Loftusia harrison 367 Medocia 433, 481 Loftusia morgani 296 Medocia blayensis 433 Loftusia persica 295, 296, 367, 369 Meganeura 117 Loftusia sp. 15, 367 Meghalaya 517 Loftusiidae 294 Meghalayana 439, 517 Loftusioidea 218, 228, 252 Meghalayana indica 439, 517 Lokhartia diversa 538 Melatolla 70 Lopingian Melatolla whitfieldensis 70 evolutionary diversity of schubertellidae during 103 Mesobitolina parva 336 fusulinine biota under stress during 121 Mesoendothyra 221, 223, 229, 264, 292 Louisettita 59 Mesoendothyra complanata 292 Louisettita elegantissima 59 Mesoendothyra izjumiana 221 Lucammina jonesi 133 Mesoendothyra izumiana 264 Lucasella 215 Mesoendothyridae 163, 220 Lugtonia 93 Mesoendothyrinae 221 Lupertosinnia 287 Mesorbitolina 303, 304, 334, 335, 336, 374, 377 Lupertosinnia pallinii 287, 361 Mesorbitolina aperta 303, 336 Lysella 73 Mesorbitolina birmanica 336 Lysella gadukensis 73 Mesorbitolina delicata 377 Mesorbitolina kurdica 377 Madagascar event 353 Mesorbitolina libanica 377 Madreporites lenticularis 304 Mesorbitolina lotzei 336 Malatyna 399, 449, 514 Mesorbitolina subconcava 336 Malatyna drobneae 399, 514 Mesorbitolina texana 303, 336, 374, 377 Malatyninae 449 Mesoschubertella 81, 102 Malaya 514 Mesoschubertella thompsoni 81 Malaysia 514, 517 Mesozoic non- perforate discoidal shells 25 Mangashtia 222, 306, 307, 358, 369 Mesozoic–C enozoic lagenides 178 Mangashtia viennoti 306, 358, 369 Mexico 110, 356, 369, 385, 454, 523, 534, 536, 541, 597 Mangashtia? egyptiensis 222 Meyendorffina 204, 215, 222, 233, 235, 236, 240, 241, 244, Mardinella 405, 447 252, 270, 276 Margaritella ospinae 420 Meyendorffina bathonica 215, 222, 236, 244, 270, 276 Marginara 50 Miarchaias 548 Marginaridae 50 Miarchaias meander 548 Marginopora 17, 18, 23, 26, 28, 37, 109, 110, 404, 452, 548, Middle Carboniferous 564, 566, 567, 568, 583, 588, 596, 598, 609, 612 fusulinidae evolved from Ozawainellidae in 101 symbionts in keriothecal walls of fusulinines 106 Middle Cretaceous 36, 285, 361 Marginopora sp. 612 Middle East 37, 234, 236, 249, 258, 442, 451, 452, 480, Marginopora vertebralis 28, 548, 564, 566, 568, 609 486, 575, 606 marine biota, of fusulinine limestones 109 Middle Kasimovian 102 Marssonella 297 Middle Miocene 35, 36 Marssonella keijzeri 391 migration route, of Profusulinella 121 Martiguesia 245, 291, 331, 332, 361 Mikhailovella 70, 139 Martiguesia cyclamminiformis 291, 361 Milahaina tortuosa 264 Maslinella 424, 454, 491 Miliola 401, 451 Maslinella chapmani 424 Miliolida 95, 107, 161, 162, 170– 173, 175, 179, 188, Matanzia 389, 502 203– 204, 226, 319– 328, 347, 388, 396–4 07, Matanzia bermudezi 389, 502 544–5 51 Mayncina 222, 288, 289 Miliolidae 347 Mayncina termieri 222 miliolides 34, 36 Mayncinidae 223, 289 of Cretaceous 335– 339 Meandropsina 324, 325, 326, 338, 358, 451 of Neogene 565–5 68 Meandropsina larrazeti 325 of Paleogene 44652 Meandropsina vidali 325, 358 of Triassic 107, 179– 180 Meandropsina? rutteni 324 Milioliporidae 173 Meandropsinidae 324 Miliolites secalicus Say 88 Meandrospina 326, 358 Milioloidea 226, 247, 446, 452, 565 Meandrospina vidali 358 Millepora miniacea 559 Meandrospira 171, 177, 180, 184, 188, 197 Millerella 77, 97, 101, 113, 130, 137, 142, 144, 226 Meandrospira dinarica 184, 197 Millerella designata 130, 144 Meandrospira pusilla 180, 188, 197 Millerella marblensis 77 Meandrospira washitensis 171 Millerella tortula 137, 142 Meandrospiranella 171, 177, 180, 197 Miniacina 559, 575, 585, 619, 622 Meandrospiranella samueli 171, 197 Miniacina miniacea 619, 622 Meandrospiridae 171 Minojapanella 81, 84, 119 Mediocris 78 Minojapanella elongata 81 Subject Index 681 Miogypsina 11, 20, 23, 26, 29, 31, 436, 437, 438, 441, 443, Miosorites americanus 548 444, 473, 478, 479, 480, 481, 482, 483, 488, 489, 541, Mirifica 70 557, 558, 564, 572, 573, 575, 581, 583, 584, 586, 601, Miscellanea 355, 417, 418, 441, 444, 456, 494, 519 603, 604, 617, 619, 622, 625, 628 Miscellanea globularis 441, 444 internal pores in 23 Miscellanea iranica 418 Miogypsina africana 622 Miscellanea meandrina 441, 519 Miogypsina antillea 573, 622 Miscellanea miscella 417, 519 Miogypsina basraensis 573 Miscellanea miscella var. dukhani 519 Miogypsina bifida 29, 489, 625 Miscellanea stampi 519 Miogypsina borneensis 437, 572, 625 Miscellaneidae 417 Miogypsina cushmani 482, 573, 622, 625 Miscellaneinae 417 Miogypsina globulina 573, 625 Miscellanites 418 Miogypsina gunteri 541, 573, 601, 625 Miscellanitinae 418 Miogypsina indonensis 437 Misellina 89, 98, 103 Miogypsina indonesiensis 483, 572, 622, 625 Misellinidae 89 Miogypsina intermedia 573, 625 Mistinia ziganensis 61 Miogypsina kotoi 625 molecular phylogenetic trees 4 Miogypsina mcmillania 622, 628 Monodiexodina 87 Miogypsina mediterranea 625 in Early to Middle Permian 103 Miogypsina Mexicana 437 Monogenerina 57, 58 Miogypsina (Miogypsina) primitiva 436, 557 Monogenerina atava 58 Miogypsina niasiensis 622 Monolepidorbis 309, 310, 312, 343, 346, 380 Miogypsina orientalis 625 Monolepidorbis douvillei 380 Miogypsina sabahensis 622 Monolepidorbis sanctae- pelagiae 309 Miogypsina samuelia 622 Monotaxinoides 54 Miogypsina s.l. 581, 603 Monotaxinoides transitorius 54 Miogypsina southernia 622 Monotaxis exilis 60 Miogypsina subiensis 625 Montiparus 84, 87, 102, 113, 137 Miogypsina tani 564, 573, 619 Montiparus montiparus 137 Miogypsina triangulata 541, 628 marker species of Middle Kasimovian 102 Miogypsindis 583, 584 Montsechiana 291, 358 Miogypsinella 436, 437, 438, 441, 444, 479, 480, 481, 482, Montsechiana aff. Montsechiensis 358 483, 488, 541, 564, 572, 574, 575, 584, 619, 628 Montsechiana martiguae 291 Miogypsinella bornea 628 Montseciella 301 Miogypsinella borodinensis 436, 564, 628 Montseciella arabica 303, 369, 374 Miogypsinella cf. borodinensis 541 Moravamminida 7, 45, 47, 52 Miogypsinella ubaghsi 488, 574, 619, 628 Moravamminidae 52 Miogypsinidae 435, 438, 453, 478, 480, 481, 488, 556, 572, Moravammininoidea 49 573, 575, 584, 599, 603 Moravamminoidea 107 miogypsinids 35, 36 Morocco 203, 204, 229, 231, 248, 264, 267, 270, 276, 280, common occurrence of 29 282, 283, 625 elongated test of 17, 18 morphology and taxonomy Miogypsinita 437, 444, 479, 558, 603 of Cretaceous larger benthic foraminifera 285–3 28 Miogypsinodella 436, 438, 483, 557, 572, 574, 575, of Jurassic larger benthic foraminifera 204– 226 584, 622 of Neogene larger benthic foraminfera 543– 562 Miogypsinodella corsicana 622 of Paleogene larger benthic foraminifera 388–4 39 Miogypsinodella primitiva 572, 622 of Triassic larger benthic foraminifera 161–1 81 Miogypsinoides 31, 436, 438, 441, 444, 473, 479, 480, 481, Moscovian 482, 483, 488, 541, 557, 558, 572, 574, 575, 584, 601, Jutland Flood Basalt event 122 609, 625, 628 Moscovian– Kasimovian boundary 102, 114 Miogypsinoides abunensis 483 Mstinia 61, 63 Miogypsinoides bantamensis 481, 482, 601, 628 Mstinia bulloides 63 Miogypsinoides complanatus 441, 444, 481, 482 Mstinia cf. bulloides 138 Miogypsinoides dehaarti 436, 481, 483, 625, 628 Mstinia fursenkoi 63 Miogypsinoides formosensis 444, 481, 541, 601, 628 “Mstinia” modavensis 62 Miogypsinoides indicus 481 Mstiniella 63 Miogypsinoides mauretanica 628 Multidiscus 173, 181 Miogypsinoides sp. 609 Multilepidina 579 Miogysinoides 601 Multispirina 322, 323, 358 Miogyspsina 593 Multispirina iranensis 323 Miolepidocyclina 29, 436, 437, 438, 444, 541, 557, 558, 572, Murciella 321, 322, 338, 361 573, 575, 603, 622, 625 Murciella cuvillieri 321, 361 Miolepidocyclina braziliana 541 Murgella 327 Miolepidocyclina excentrica 622, 625 Murgella lata 327 Miolepidocyclina mexicana 541, 622 Namibia 350 Miolepidocyclina panamensis 541 Nankinella 79 Miosorites 548, 549 Nanlingella 81, 84, 99 682 Subject Index Nanlingella meridionalis 81 Neoschwagerinidae 75, 85, 90, 97, 98, 99, 101, 103, 121 Nanopora anglica 159 schematic features of 85 Naupliella 304, 379 Neosivasella 430, 479 Naupliella insolita 304, 379 Neosivasella sungurlui 430 Nautiloculina 226, 247, 270, 328 Neostaffella 78, 146 Nautiloculina circularis 270, 328 Neotaberina 406, 447, 451, 514 Nautiloculina cretacea 270, 328 Neotaberina neaniconica 406, 514 Nautiloculina oolithca 270 Neotrocholina 206, 207, 208, 227, 247, 259, 377 Nautiloculina oolithica 226, 328 Neotrocholina friburgensis 208 Nautiloculinidae 226, 328 Neotrocholina lenticularis 206, 377 Nautilus angulatus 405, 547 Neotrocholina minima 377 Nautilus macellus 439 Neotrocholina valdensis 259 Nautilus orbiculus 407, 549 Nephrolepdina radiata 625 Nautilus planatus 403 Nephrolepidina 441, 475, 476, 497, 578, 579, 601, 602, 625, Nautilus spengleri 556 628, 631, 634, 636 Nemkovella 415, 456, 470, 523 Nephrolepidina sumatrensis 441 Nemkovella mcmilliana 523 Netherlands 358, 480 Neoarchaesphaera (Elenella) multispinosa 50 Neurnannites 559 Neobrunsiina 63 Nevillea 63 Neocarpenteria 423, 491 New Zealand 121, 593, 622 Neocarpenteria cubana 423 Nezzazata 287, 358 Neodiscocyclina 416, 456, 538 Nezzazata conica 358 Neodiscocyclina anconensis 538 Nezzazata simplex 287 Neofusulinella 81 Nezzazatidae 287, 347 Neofusulinella praecursor 81 Nezzazatinae 287 Neogene larger benthic foraminfera Nezzazatinella 287, 358 biostratigraphy and phylogenetic evolution 561– 582 Nezzazatinella adhami 287 letter stages for Neogene of SE Asia 564–5 65 Nezzazatinella picardi 358 miliolides of Neogene 565–5 68 Nezzazatoidea 222, 245, 288 rotaliides of Neogene 568–5 81 Nicaragua 514, 517 textulariides of Neogene 581– 582 Ninella 78 evolution suborders and superfamilies 544 Nodochernyshinella 64 Miliolida 65– 568, 543, 544–5 51 Nodosaria 90, 92, 94, 107, 134 morphology and taxonomy 543– 562 Nodosaria gracilis 94 palaeoecology 583– 587 Nodosaria proceraformis 92 palaeogeographic distribution 588– 604 Nodosaria radicula 107, 134 Rotaliida 543, 551– 559, 568–5 81 Nodosinella 49, 57, 93 taxa in Tethyan carbonate shelf 588 Nodosinella concinna 93 Textulariida 33, 559– 562, 581– 582 Nodosinella digitata 93 Nodosinella perelegans 93 Neohemigordius 45, 95, 172, 180 Nodosinellidae 49, 93, 100 Neohemigordius maopingensis 95 Nodosinelloidea 51–5 3 Neoiraqia 304, 358 Earlandinitidae 51 Neoiraqia convexa 304, 358 Nodosinellidae 51– 53 Neokilianina 212, 214, 234, 237, 273 Nodosinelloides 94, 166, 178 Neokilianina rahonensis 234, 273 Nodosinelloides potievskayae 94 Neolacazopsis 400 Nonionina cretacea 325 Neolacazopsis osozawai 400 North Africa 114, 119, 249, 258, 486, 498, 599, 601 Neolepidina 428 North America 103, 114, 115, 117, 118, 119, 121, 185, 187, Neolepidina (Bronnimann) 428 188, 191, 224, 348, 492, 598, 599 Neoplanorbulinella 420, 571 North American Basin 115 Neoplanorbulinella saipanensis 420 North Atlantic 250, 256, 257, 494, 495, 604 Neorbitolinopsis 302, 374 North Atlantic Ocean 250, 604 Neorbitolinopsis conulus 374 North Atlantic Volcanic Provence (NAVP) 494, 495 Neorhipidionina 406, 447, 508, 514 North Pacific Ocean 291, 352 Neorhipidionina macfadyeni 508 Northern Alps 198, 201 Neorhipidionina spiralis 514 Novella 73 Neorhipidionina williamsoni 508 Novella evoluta 73 Neorotalia 318, 319, 435, 436, 438, 441, 479, 481, 482, 483, Nummofallotia 325 498, 541, 574, 575, 601, 604, 628 Nummularia exponens 413 Neorotalia sp. 541 Nummulina globulina 437 Neorotalia tethyana 498, 628 Nummulina pristina 413 Neoschubertella 81 Nummulites 2, 11, 19, 24, 26, 27, 309, 387, 408, 411, Neoschubertella sisophonensis 81 412, 413, 417, 428, 441, 444, 456, 457, 458, 459, 460, Neoschwagerina 82, 90, 98 461, 463, 464, 473, 478, 486, 487, 488, 490, 494, 498, Neoschwagerina aff. craticulifera 154 499, 508, 517, 523, 526, 529, 532, 553, 554, 570, 600, Neoschwagerina sp. 151, 154, 158 601, 606 Neoschwagerina (Yabeina) inouyei 90 canal system 25 Subject Index 683 Nummulites atacicus 441, 461, 508, 517 Omphalocyclinae 309 Nummulites aturicus 529 Omphalocyclus 309, 310, 311, 312, 316, 343, 347, 383, 384 Nummulites beaumonti 461 Omphalocyclus macroporus 309, 383 Nummulites benehamensis 441, 459 Omphalocyclus sp. 311, 383, 384 Nummulites beneharnensis 529 Omphalotis 70 Nummulites boulangeri 441 Ontong Java Plateau 352 Nummulites burdigalensis 441 Ontong Java volcanic event 345, 351, 353 Nummulites cantabricus 441 Operculina 19, 24, 26, 345, 408, 412, 413, 443, 456, 463, Nummulites cf. irregularis 417 464, 486, 487, 489, 497, 508, 517, 521, 526, 532, 553, Nummulites cf. striatus 529 564, 565, 570, 586, 588, 598, 604, 628 Nummulites clypeus 526 Operculina aegyptiaca 408, 526 Nummulites crassus 441 Operculina ammonoides 570 Nummulites deserti 408, 461 Operculina (Assilina) 412 Nummulites djokdjokartae 459 Operculina douvillei 526 Nummulites fabianii 441 Operculina heterostegnoides 553 Nummulites fichteli 408, 413, 441, 461, 478, 526, Operculina subgranosa 526 529, 553 Operculinella 413, 456, 463, 487, 488, 553, 554 Nummulites fichteli- intermedius 408, 478, 526 Operculinella cumingii 413, 553, 554 Nummulites fossulata 508 Operculinella venosus 487, 488 Nummulites gizehensis 408, 460, 526, 529 Operculinoides 413, 604 Nummulites globulus 461, 529 Operculinoides antiguensis 628 Nummulites herbi 441 Operculinoides ocalanus 529 Nummulites hormoensis 461 Operculinoides panamensis 628 Nummulites incrassatus 461 Opertorbitolites 406, 447, 511 Nummulites intermedius 408, 478, 517, 529 Opertorbitolites cf. douvillei 511 Nummulites irregularis 408, 460 Opertorbitolites douvillei 406, 511 Nummulites javanus 460 Opertorbitolites lehmanni 511 Nummulites laevigatus 27, 441, 459, 529 Opertorbitolites sp. 511 Nummulites lamarcki 529 Opertum 296, 297, 298, 333, 371 Nummulites lyelli 441 Opertum orbignyna 333 Nummulites mamilla 529 Opertum sp. 371 Nummulites manfredi 441 Ophthalmidiidae 171 Nummulites mantelli 428 Ophthalmidium 171, 172, 180 Nummulites marginata 428 Orbicula elliptica 222 Nummulites masiraensis 417, 526 Orbiculina compressa 405 Nummulites millecaput 460, 487, 488 Orbiculina peruviana 325 Nummulites minervensis 441 Orbiculina sp. 609 Nummulites nuttalli 413 Orbignyna 298, 333 Nummulites papyraceus 309 Orbis foliaceus 95 Nummulites pengaronensis 461 Orbitammina 222 Nummulites perforatus 529, 532 Orbitoclypeinae 416, 468, 470 Nummulites planulatus 459 Orbitoclypeus 415, 416, 456, 470, 477, 490, 538 Nummulites praecursor 461 Orbitoclypeus himerensis 416 Nummulites robusformis 441 Orbitoclypeus nummuliticus 538 Nummulites rotularius 461 Orbitoclypeus? cristensis 415 Nummulites somaliensis 523 Orbitocyclina 310, 315, 317, 343 Nummulites spira 412 Orbitogypsina 424 Nummulites striatus 461 Orbitogypsina vesivularis 424 Nummulites uroniensis 459 Orbitoides 26, 308, 309, 310, 311, 313, 314, 340, 341, 342, Nummulites vascus 413, 461, 529, 553, 554 343, 346, 347, 379, 380, 381, 384, 415, 428, 430, 437, Nummulites venosus 570 557, 558 Nummulites willcoxi 413 Orbitoides browni 381 nummulitic tests, shapes of 19 Orbitoides cf. tissoti 380 Nummulitidae 411, 457, 458, 463, 553, 570, 571, 585 Orbitoides dilatata 428 Nummulitinae 412, 570 Orbitoides faujasii 380, 381 Nummulitoidea 453, 456, 465, 563, 568, 569 Orbitoides (Lepidocyclina) burdigalensis 437, 558 Nummulitoides margaretae 521 Orbitoides (Lepidosemicyclina) thecideaeformis 557 Nummulostegina padangensis 173 Orbitoides mammillatus 313 Nwabangyi Dolomite Formation 193, 194, 195, 196, Orbitoides media 309 198, 201 Orbitoides medius 381 Obsoletes 84, 113 Orbitoides socialls 314 Occidentoschwagerina alpina 149 Orbitoides strophiolata 415 Oceanic Anoxic Event (OAE) 248 Orbitoides vredenburgi 381 Oculina liasica 171 Orbitoididae 308, 310, 312, 342, 347, 430 Oerculina bermudezi 412 Orbitoidinae 309 Oligoporella sp. 159 Orbitoidoidea 35 Oman 120, 242, 264, 276, 280, 282, 376, 377, 417, 502, Orbitokathina 319, 430 511, 517, 519, 529 Orbitokathina vonderschmitti 319 684 Subject Index Orbitolina 20, 21, 22, 23, 24, 35, 36, 300, 302, 303, 304, Pachyphloia magna 133 334, 335, 371, 374, 376, 377, 396, 445, 555 Pachyphloia ovata 94 Orbitolina cf. duranddelgai 374 Pachyphloia– Howchinella 178 Orbitolina concava 371, 555 Pachyphloides 167 Orbitolina concava var. sphaerulata 555 Pachyphloiidae 94, 100 Orbitolina conulus 302 Pakistan 283, 408, 502, 505, 508, 517, 519, 520, 523, 526, Orbitolina qatarica 303, 374, 377 529, 532, 538, 541 Orbitolina sefini 374 Palaeobigenerina 57, 58, 104, 105 Orbitolina sp. 303, 374 Palaeocene– Eocene Thermal Maximum (PETM) 484, Orbitolina trochus 376 493, 494 Orbitolinella 302 Palaeocyclammina 219, 235, 241, 280 Orbitolinella depressa 302, 374 Palaeocyclammina complanata 219, 280 Orbitolinidae 299, 347, 394 Palaeodasycladus 229, 248, 283 Orbitolinoidea 35, 329, 445, 446, 581 Palaeodasycladus mediterraneus 229, 248 Orbitolinopsis 301, 302, 336, 374 Palaeodasycladus? mediterraneus 283 Orbitolinopsis kiliani 302, 374 palaeoecology Orbitolinopsis sp. A 336 of Cretaceous larger benthic foraminifera 344– 348 Orbitolinopsis? neoelongata 301 of fusulinines 107– 111 Orbitolites 9, 23, 26, 303, 309, 404, 405, 406, 414, 444, 447, of Jurassic larger benthic foraminifera 247–2 49 452, 486, 508, 511, 517, 532, 546, 549 of Neogene larger benthic foraminfera 583– 587 Orbitolites americana 549 of Paleogene larger benthic foraminifera 484–4 92 Orbitolites biplanus 511 of Triassic larger benthic foraminifera 182–1 85 Orbitolites complanatus 406, 508, 511, 517 Palaeofusulina 81, 82, 84, 98, 99, 103 Orbitolites conica 303 Palaeofusulina prisca 81 Orbitolites malabarica 549 palaeogeographic distribution Orbitolites media 309 of Cretaceous larger benthic foraminifera 348–3 56 Orbitolites radians 414 of Jurassic larger benthic foraminifera 250–2 58 Orbitolites shirazensis 405 of Neogene larger benthic foraminfera 588–6 04 Orbitolites sp. 9, 532 of Paleogene larger benthic foraminifera 492– 500 Orbitolites tenuissimus 546 of Triassic larger benthic foraminifera 182–1 85 Orbitopsella 22, 23, 24, 204, 216, 221, 222, 229, palaeogeographic distribution, of fusulinines 111–1 25 232, 233, 235, 240, 241, 242, 243, 248, 280, at Serpukhovian- Bashkirian boundary 117 282, 283 decrease in diversity at the end Sakmarian 120 Orbitopsella africana 243 diversity during Capitanian 121 Orbitopsella “circumvulvata” 232 diversity during Pennsylvanian 119– 120 Orbitopsella dubari 282 Early Carboniferous 113–1 14 Orbitopsella praecursor 232, 242, 243, 282 End Devonian extinction events 114 Orbitopsella primaeva 242, 243, 280, 282 End Permian mass extinction 122–1 24 Orbitopsella spp. biozone 232 partial extinction at the end Moscovian 121 Orbitopsella tibetica 243 recovery of shallow reefal environment 114 Orbitopsella zone 229 palaeogeographic reconstruction Orbitopsellinae 221, 243 early Permian 120 Orbitosiphon 430, 523 Late Devonian 113 Orbitosiphon praepunjabensis 523 Late Pennsylvanian 119 Orbitulites praecursor 221 Visean 118 Orbitulites texanus 304 Palaeolituonella 163, 177, 185, 188 Orbulites concava 304 Palaeolituonella majzoni 163 Orduella 431 Palaeolituonella meridionalis 185, 188 Orduella sphaerica 431 Palaeomayncina 222, 235, 240, 241, 273 Orduellinidae 431 Palaeomayncina ternieri 273 Orientalia 297 palaeontology, of larger foraminifera 34–3 6 Orobias kueichihensis S. 76 Palaeonubecularia cf. uniserialis 128 Orthophragmina advena 416 Palaeonummulites 413, 463, 464 Orthophragmina floridana 415 Palaeonummulites kugleri 529, 554 Orthophragminidae 414, 467, 468, 488 Palaeonummulites nomen oblitum 413, 553 Oryzaria boscii 396 Palaeonummulites pristina 538 Otbitammina elliptica 270 Palaeopfenderina 210, 211, 212, 234, 267 Ovalveolina 322, 323, 337, 338, 385 Palaeopfenderina salernitana 267 Ovalveolina reicheli 323 Palaeopfenderina trochoidea 267 Ovalveolina sp. 385 Palaeospiroplectammina 65, 104, 105, 131 Ozawainella 77, 99, 101, 109, 145 Palaeospiroplectamminidae 65, 100, 104 Ozawainellidae 26, 75, 76, 77, 97, 98, 101, 102, 116 Palaeotextularia 57, 58, 65, 104, 105, 107, 132, 133 fusulinidae evolved from 97, 98 Palaeotextularia angulata 132 schematic features of 77 Palaeotextularia diversa 65 Palaeotextularia longiseptata 132 Pachyphloia 94, 111, 133, 166, 168, 178 Palaeotextularia schellwieni 58 Pachyphloia asymmetrica 133 Palaeotextulariida 7, 45, 47, 56, 57, 104 Pachyphloia depressa 133 Palaeotextulariidae 56, 57, 100, 104, 105 Subject Index 685 Palaeotextularioidea 56– 57, 65, 100 Pararotalia 26, 318, 463, 464, 479, 481, 482, 497 biseriamminidae 58 Pararotaliinae 318, 435 evolution during palaeozoic 65 Paraschwagerina 87, 147 semitextulariidae 56 Paraschwagerina lata 149 Palaeozoic lagenides 104, 105 Paraschwagerina stachei 149 Paleodictyoconus 301, 334, 335, 336, 374 Parasorites 549 Paleodictyoconus cuvillieri 374 Parastaffella (Eoparastaffella) simplex 76 Paleogene larger benthic foraminifera Parastaffella pressa 68 biostratigraphy and phylogenetic evolution 439– 484 Parathikhinella 61 letter stages of Southeast Asia and provincial Parathurammina 48, 50, 128 biostratigraphy 442– 445 Parathurammina aperturata 128 miliolides of Paleogene 446–4 52 Parathurammina dagmarae 48 rotaliides of Paleogene 452–4 84 Parathurammina tamarae 50 textulariides of Paleogene 445– 446 Parathuramminida 7, 45, 47, 48, 96, 114, 162, 178 morphology and taxonomy 388– 439 Parathuramminidae 48, 49, 100 Miliolida 388, 396–4 07 Parathuramminoidea 45 Rotaliida 388, 407– 439 Moravamminoidea evolved from 107 Textulariida 388– 396 morphology and taxonomy of Palaeozoic 48– 55 palaeoecology 484–4 92 Nodosinelloidea evolved from 57 palaeogeographic distribution 492–5 00 ranges in Carboniferous and Permian 100 Paleomiogypsina 435, 436, 438, 441, 444, 479, 481, 482, Paratikhinella 55 541, 574, 575, 628 Paratikhinella cylindrica 133 Paleomiogypsina boninensis 436, 541 Paratikhinellidae 55–5 6, 100 Paleomiogypsina sp. 628 Paratriasina 171, 172, 177, 180 Paleomyogypsina 483 Paratriasina jiangyouensis 171 Paleopfenderininae 210, 238 Paratuberitina 51, 128 Palorbitolina 303, 304, 305, 334, 335, 336, 345, 376, 377 Paravalvulina 214, 225, 232, 246, 273, 377 Palorbitolina cf. Lenticularis 377 Paravalvulina arabica 377 Palorbitolina discoidea 376 Paravalvulina complicata 214, 225, 232 Palorbitolina lenticularis 303, 336, 345, 376 Paravalvulina sp. 273 Palorbitolinoides 303, 305, 336 Paravalvulininae 225, 306 Palorbitolinoides hedini 303, 305, 336 Parphia 51, 130 Palorbitolinoides orbiculata 336 Parurgonina 210, 212, 214, 233, 234, 237, 239, 273 Pamirina 77, 103 Parurgonina caelinensis 233, 234 Pamirina darvasica 77 Parurgonina coelinensis 273 Pangea 182, 250, 348 Parurgonininae 214 Papillaria 167, 177 Parvalamella 175 Papillaria laghii 167 Pastrikella 325 Papua New Guinea 28, 609, 612, 617, 619, 622, 625, 628, Patella (Cymbiola) cassis 422 631, 634, 636 Patellina bradyana 54 Paracaligella 49, 55 Patellina egyptiensis 395 Paracaligella antropovi 55 Paulbronnimannia 163, 177, 198 Paracoskinolina 215, 264, 301, 358, 374 Paulbronnimannia judicariensis 198 Paracoskinolina occitana 264 Pavlovecina 327 Paracoskinolina occitanica 215 Pavonitina 390, 502 Paracoskinolina sunnilandensis 374 Pavonitina styriaca 390, 502 Paradagmarita 59 Pavonitinidae 389, 562, 581 Paradagmarita monodi 59 Pavonitininae 389, 562 Paradainella 70 Pavonitinoidea 581 Paradainella dainelliformis 70 Pavopsammia 390 Paradoxiella 82 Pavopsammia flabellum 390, 502 Paradoxiella pratti 82 Peelella 421 Paradunbarula 82 Peelella boninensis 421 Paradunbarula dallyi 82 Pellatispira 35, 407, 409, 411, 444, 453, 456, 464, 467, Paraendothyra 70 508, 520 Paraendothyra nalivkini 70 Pellatispira douvillei 409 Paraeofusulina 84, 101 Pellatispira fulgeria 409, 520 Parafusulina 84, 85, 87, 97, 99, 101, 103, 109, 146, 154 Pellatispira inflata 409 Parafusulina kaerimizensis 154 Pellatispira sp. 508, 520 Parafusulina schuberti 85 Pellatispirella 439, 479, 523 Parafusulina wordensis 87 Pellatispirella antillea 523 Parafusulinella 84 Pellatispiridae 407, 465 Parafusulinella propria 84 Penarchaias 403, 447 Paraglobivalvulina 59 Peneroplidae 42, 403, 452, 546 Paraglobivalvulina mira 59 Peneroplis 26, 29, 30, 42, 108, 209, 321, 327, 403, 404, 447, Paraglobivalvulina? septulifera 59 451, 514, 547, 567, 568, 612 Paraglobivalvulinoides 59 Peneroplis glynnjonesi 403 Paraplectogyra 70 Peneroplis karreri 327, 547 686 Subject Index Peneroplis liburnica 321 Planoendothyra? kharaulakhensis 70 Peneroplis pertusus var. discoideus 547 Planogypsina 470, 559 Peneroplis proteus 404 Planogypsina squamiformis 470 Peneroplis senoniensis 208 Planolinderina 421, 454, 455, 491, 571 Peneroplis sp. 42, 514, 612 Planolinderina escornebovensis 421 Peneroplis thomasi 612 Planoperculina 553, 586 Periloculina 338, 401 Planorbulina 307, 420, 421, 427, 454, 455, 491 Permian- Triassic extinction 177, 180– 181 Planorbulina mediterranensis 421 Permo-C arboniferous Planorbulina (Planorbulinella) guayabalensis 427 fusulinides generic extinctions and speciations 117 Planorbulina vulgaris var. larvata 421 fusulinides genera through 114 Planorbulina? cenomaniana 307 fusulinoidea wall structures development in 96 Planorbulinella 421, 427, 454, 455, 470, 491, 538, 564, 587, Parathutamminoidea and Robuloidoidea ranges in 100 588, 599, 612, 615, 622, 631 Permodiscus 53 Planorbulinella batangenensis adamsi 615 Permodiscus vetustus 53 Planorbulinella kinabatangenensis 615 Perouvianella 325 Planorbulinella larvata 421, 612, 615, 622 Peru 154, 157, 532, 541 Planorbulinella solida 421, 538, 564, 615 Pfenderella 211, 212, 213, 234, 236, 267 Planorbulinidae 420, 454, 554 Pfenderella arabica 211, 267 Planorbulinoidea 422, 453, 454, 491, 568, 571 Pfenderenidae 293, 297 Planorbulinopsis 545, 554, 571, 585 Pfendericonus 390, 392, 445 Planorbulinopsis parasitica 545, 554, 585 Pfenderina 210, 211, 212, 213, 232, 233, 234, 236, 237, Planostegina 413, 456, 462, 465, 521, 526, 569, 570, 612 238, 267 Planostegina africana 462, 569 Pfenderina butterlini 210 Planostegina operculinoides 570 Pfenderina neocomiensis 238, 267 Platyhaurania 215, 216, 235, 240, 241, 276 Pfenderina salernitana 210, 232, 237, 238 Platyhaurania subcompressa 276 Pfenderina trochoidea 232, 236, 238 Plectogyra 63, 64, 71, 130, 142, 143, 144 Pfenderina? chablaisensis 210 Plectogyra bradyi 144 Pfenderininae 211, 238, 293 Plectogyra cf. geniculata 142 Pfenderinoidea 209, 212, 213, 227, 228, 234, 252 Plectogyra cf. pandorae 144 of Jurassic 234– 239 Plectogyra excellens 144 Phenacophragma 288, 289 Plectogyra irregularis 130, 144 Phenacophragma assurgens 289 Plectogyra (Latiendothyra) quaesita 63 phylogenetic analysis, foraminiferal DNA sequences 4 Plectogyra michoti 63 phylogenetic evolution Plectogyra pauli 64 and biostratigraphy see biostratigraphy Plectogyra phrissa 144 of fusulinides 96–1 03 Plectogyra plectogyra 71 Palaeozoic lagenides 106 Plectogyra tumula 64 phylogeny of eukaryotes 4 Plectogyranopsis 71 Piallina 165, 177, 198 Pliolepidina tobleri 535 Piallina bronnimanni 198 Pohlia 64 Piallina tethydis 165 Pojarkovella 71 Piallinidae 165 Pojarkovella honesta 71 Pilaminella 177 Polydiexodina 85, 86, 87, 97, 99, 103, 147 Pilaminella gemerica 184 from Kurdistan 147 Pilammina 163, 177, 184, 198 Polydiexodina afghanensis 87 Pilammina densa 163, 184, 198 Polydiexodina capitanensis 87 Pilammina grandis 163 Polydiexodina praecursor 147 Pilamminella 163 Polylepidina 427, 471, 472, 473, 476, 477, 478, 532 Pisolina 79 Polylepidina chiapasensis 473, 477, 478 Pisolina excessa 79 Polylepidina chiapasensis subplana 477, 478 Planiinvoluta 173, 201 Polytaxis 59 Planiinvoluta carinata 173 Polytaxis laheei 59 Planiinvoluta? mesotriasica 201 Polytrema cylindrica 559 Planisepta 221, 222, 229, 232, 233, 235, 240, 241, 270 Polytrema cylindricum 425 Planisepta compressa 229, 232, 233, 270 Polytrema planum 424 biozone 232 Porites 585 Planisepta compressa zone 229 Postendothyra 72 Planiseptinae 221 Postendothyra scabra 72 Planispiral-f usiform 24, 34 Praealveolina 322, 323, 324, 337, 384, 385, 446 Planispirina 226 Praealveolina cretacea 385 Planktonic foraminifera 251– 258 Praealveolina simplex 324 Planoarchaediscus 53, 136 Praealveolina tenuis 323, 385 Planoarchaediscus emphaticus 136 Praebullalveolina 399, 449, 514 Planocamerinoides 413, 456, 529 Praebullalveolina afyonica 399, 514 Planoendothyra 70, 139, 142 Praechrysalidina 297, 306, 330, 377 Planoendothyra cf. aljutovica 139 Praechrysalidina infracretacea 306, 377 Planoendothyra mameti 142 Praekurnubia 211, 212, 213, 236, 238, 270 Subject Index 687 Praekurnubia crusei 213, 238, 270 Pseudoendothyra 79, 108, 138, 139, 142, 144 Praeophthalmidium (Eoophthalmidium) tricki 171 Pseudoendothyra composite 139 Praeorbitolina 303, 305, 334, 335, 336 Pseudoendothyra luminosa 138 Praeorbitolina cf. wienandsi 336 Pseudoendothyra struvii 142, 144 Praeorbitolina cormyi 303, 305, 336 Pseudofabularia 400, 441, 447, 514 Praeorbitolina wienandsi 336 Pseudofabularia matleyi 514 Praepeneroplis 338, 451 Pseudofusulina 84, 87, 99, 120, 149 Praereticulinella 294, 361 Pseudofusulina huecoensis 87 Praereticulinella cuvilleri 361 Pseudofusulinella occidentalis 137 Praereticulinella cuvillieri 294 Pseudohauerina 402, 546 Praerhapydionina 26, 406, 441, 447, 451, 511, 612 Pseudohauerina dissidens 546 Praerhapydionina cubana 406 Pseudohauerinella 546 Praerhapydionina delicata 511, 612 Pseudolacazina 321, 400, 401, 447, 451 Praeskinerella 99, 146 Pseudolacazina hottingeri 321, 400 Praesorites orbitolitoides 549 Pseudolepidina 420, 430, 476, 523 Praetaberina 327 Pseudolepidina trimera 430, 523 Praetaberina bingistani 381 Pseudolituonella 305, 364, 392, 445 Priscella 71 Pseudolituonella reicheli 305, 364 Profusulinella 82, 84, 97, 98, 99, 102, 109, 110, 113, 117, Pseudolituotuba 55 118, 137 Pseudolituotubella 64 migration route of 121 Pseudolituotubella multicamerata 64 Profusulinella pararhomboides 84 Pseudolituotubidae 55, 100 Propermodiscus 53, 136 Pseudomarssonella 225, 232, 246, 270, 273, 297, 377 Proporocyclina 415, 416, 468 Pseudomarssonella bipartia 273 Proto-A tlantic Ocean 348 Pseudomarssonella bipartita 232 Protogypsina 425, 517 Pseudomarssonella cf. plicata 377 Protogypsina indica 425, 517 Pseudomarssonella inflata 273 Protonodosaria 92, 166, 178 Pseudomarssonella maxima 225, 232, 270, 273 Protonodosariidae 92 Pseudomarssonella mcclurei 225 Protopeneroplis 209, 227, 247, 264 Pseudomarssonella media 225 Protopeneroplis striata 264 Pseudomarssonella plicata 273 Prototriticites 102 Pseudomphalocyclus 309, 310 Protriticites 84, 99, 113 Pseudomphalocyclus blumenthali 309 Protriticites globulus 84 Pseudonovella 77 Provincialism 11, 595, 599 Pseudonovella irregularis 77 Pseudedomia 321, 322, 338, 357, 385 Pseudonummuloculina 338, 371, 446, 448 Pseudedomia complanata 357 Pseudopfenderina 210, 211, 212, 229, 232, 234, 238, Pseudedomia globularis 321, 357 248, 267 Pseudedomia multistriata 321 Pseudopfenderina butterlini 229, 232, 238, 267 Pseudoammodiscidae 55, 96, 100 Pseudopfenderininae 210 Pseudobolivina 224 Pseudophragmina 414, 415, 416, 456, 468, 469, 511, 536 Pseudobroeckinella 325 Pseudophragmina (Astemphragmina) pagoda 414 Pseudobroeckinella soumoulouensis 325 Pseudophragmina (Athecocyclina) 415 Pseudochoffatella 20, 21, 294, 295, 330, 331, 332, 361, 364 Pseudophragmina (Athecocyclina) cookei 414 Pseudochoffatella cuvillieri 294, 295, 361, 364 Pseudophragmina floridana 511, 536 Pseudochrysalidina 391, 392, 445 Pseudophragmina (Proporocyclina) 415 Pseudochrysalidina floridana 391 Pseudorbitellinae 317 Pseudochubbina 321 Pseudorbitoides 310, 314, 315, 316, 317, 342, 343, 347 Pseudocyclammina 20, 21, 218, 219, 231, 232, 234, 235, Pseudorbitoides longispiralis 314 241, 243, 245, 248, 249, 264, 273, 331, 332, 364, Pseudorbitoides trechmanni 316 371, 374 Pseudorbitoididae 310, 316 Pseudocyclammina bukowiensis 231, 232, 234, 273, 364 Pseudorbitoidinae 316 Pseudocyclammina cylindrica 364 Pseudorbitolina 301, 367 Pseudocyclammina kelleri 273 Pseudorbitolina marthae 301, 367 Pseudocyclammina lituus 231, 234, 264, 273, 364, 371, 374 Pseudorhapydionina 327 Pseudocyclammina maynci 232, 234 Pseudorhipidionina 327 Pseudocyclammina powersi 219 Pseudoschwagerina 85, 86, 87, 97, 103, 111, 119, 120, 156 Pseudocyclammina rugosa 364 East European Basin 102 Pseudocyclammina sp. 232 palaeoecology of 111 Pseudocyclammina sphaeroidalis 231, 234, 264 Pseudoschwagerina aequalis 149 Pseudocyclammina (Streptocyclammina) parvula 218 Pseudoschwagerina cf. P. fusiformis 147 Pseudocyclammina ukrainica 234 Pseudoschwagerina elegans 149 Pseudocyclammina vasconica 231, 273 Pseudoschwagerina morikawai 86 Pseudodoliolina 89, 111 Pseudoschwagerina nitida 149 Pseudodoliolina ozawai 89 Pseudoschwagerina pulchra 149 Pseudodoliolininae 89 Pseudoschwagerina schellwieni 149 Pseudoeggerella 211, 212, 236, 267 Pseudoschwagerina (Zellia) heritschi 89 Pseudoeggerella elongata 211, 267 Pseudosiderolites 310, 314, 343 688 Subject Index Pseudospirocyclina 217, 218, 219, 232, 234, 240, 241, 249, Rectocyclammina 217, 220, 231, 232, 234, 241, 245, 276 258, 264 Rectocyclammina ammobaculitiformis 232, 234 Pseudospirocyclina mauretanica 234, 240, 264 Rectocyclammina cf. chouberti 231 Pseudospirocyclina maynci 217, 240, 264 Rectocyclammina chouberti 220, 234, 245, 276 Pseudospirocyclina muluchensis 264 Rectodictyoconus 305 Pseudospirocyclina smouti 232, 233, 240, 264 Rectodictyoconus giganteus 305 Pseudostaffella 79, 98, 109 Rectoendothyra 71 Pseudostaffella antiqua 117 Rectoseptatournayella 64 Pseudostaffella needhami 79 Rectoseptatournayella stylaensis 64 Pseudostaffella sphaeroidea 78 Rectostipulina 91, 178 Pseudostaffellidae 78 Rectostipulina quadrata 91 Pseudotaberina 549, 550, 566, 588, 612 Rectotournayellina 64 Pseudotaberina malabarica 550, 612 Redmondellina 219, 234, 235, 241, 245, 249, 252, 270, Pseudotaxidae 60, 100 276, 280 Pseudotaxis 60, 107 Redmondellina powersi 234, 270, 276, 280 Pseudotextulariella 297, 298, 333, 334, 371 Redmondina 433 Pseudotextulariella courtionensis 371 Redmondina henningtoni 433 Pseudotextulariella cretosa 371 Redmondininae 433 Pseudotriplasia 562 Redmondoides 225, 232, 246, 273 Pseudotriplasia elongata 562 Redmondoides inflatus 232 Pseudotristix 92, 178 Redmondoides lugeoni 232, 273 Pseudovidalina 45, 95, 180 Redmondoides medius 232, 273 Pseudowanganella 94 Redmondoides rotundatus 273 Pseudowanganella tenuitheca 94 reef ecology 605, 607 Ptychocladia 49, 56 reef gap 182, 192 Ptychocladia agellus 56 reef systems 182, 184 Ptychocladiidae 49, 56, 100 Reichelina 78 Ptychocladioidea 100 Reichelina cribroseptata 78 Punjab 380, 532 Reticulina reicheli 294 Puteolina 404, 452, 568 Reticulinella reicheli 361 Pyramis parva 93 Reticulophragmium 561 Reticulophragmium orbicularis 561 Qatar 237, 261, 267, 276, 280, 303, 358, 361, 364, 367, 369, Rhabdorites 406, 451, 508, 511 371, 374, 377, 379, 380, 382, 385, 434, 502, 519, 523, Rhabdorites malatyaensis 511 529, 538, 541 Rhabdorites urensis 508 Qataria 291, 361 Rhapydionina 321, 322, 327, 338, 406, 451 Qataria dukhani 291, 361 Rhapydionina laurinensis 327 Quasiendothyra 71, 96 Rhapydionina malatyaensis 406 Quasiendothyra urbana 72 Rhapydionina urensis 451 Quasifusulina 15, 84, 113, 145, 146, 151 Rhapydioninidae 321 Quasireichelina 78 Rhipidionina 327, 338, 406, 451 Quasireichelina expansa 78 Rhipidionina casertana 327 Quasirotalia 556, 557 Rhipidionina macfadyeni 451 Quasirotalia guamensis 556, 557 Rhipidionina williamsoni 406, 451 Quinqueloculina 9, 33, 226, 322, 338, 565 “Rhizaria” 1, 5 Quinqueloculina sp. 9 Rhodesina avonensis 69 Rhodesinella 71 R- strategists 27 ribosomal RNA gene trees 5 Raadshoovenia 321 Riveroina 546 Raadshoovenia guatemalensis 321 Riveroina caribaea 546 Rabanitina 288, 374 Rivieroinidae 402, 449, 546 Rabanitina basraenesis 288 Riyadhella 225, 229, 232, 245, 246, 273, 297 Rabanitina basraensis 374 Riyadhella arabica 225 Radiocycloclypeus 553, 570 Riyadhella elongata 225 Ranikothalia 413, 444, 456, 494, 519, 529, 532 Riyadhella hemeri 225 Ranikothalia nuttalli 519, 532 Riyadhella inflata 225 Ranikothalia nuttalli kohatica 519 Riyadhella intermedia 225 Ranikothalia sahnii 529 Riyadhella nana 225, 273 Ranikothalia sindensis 532 Riyadhella praeregularis 225, 229, 273 Raoia 405, 517 Riyadhella regularis 225, 273 Raoia indica 405, 517 Riyadhella sp. 232, 273 Rauserella 79 Riyadhoides 225, 246, 273 Rauserella erratica 79 Riyadhoides mcclurei 273 Rectochernyshinella 65 Riyhadella nana 225, 273 Rectocibicides 455 Robuloides 92, 168, 178, 198 Rectocornuspira 95, 180, 188 Robuloides lens 92, 168 Rectocornuspira kalhori 180, 188 Robuloides reicheli 198 Rectocornuspira lituiformis 95 Robuloididae 92, 168 Subject Index 689 Robuloidoidea 45, 90, 100, 106, 178 Schubertella lata 81 ranges in Carboniferous and Permian 106 Schubertella transitoria 82 Robustoconus 215, 218 Schubertellidae 26, 75, 77, 80, 97, 98, 102, 103 Robustoconus tisljari 215, 264 arose from Fusulina in Late Carboniferous 102 Robustoschwagerina geyeri 149 evolutionary diversity during Lopingian 103 Robustoschwagerina tumida 149 Schubertina 82, 102 Rosalina 488, 584 Schubertina circuli 82 Rotalia 15, 26, 319, 339, 340, 358, 433, 435, 453, 481, 488, Schwagerina 11, 82, 87, 88, 90, 98, 99, 103, 145, 146, 147, 541, 584 151, 154, 156, 157 Rotalia mexicana 435 Schwagerina adamsi 147, 154 Rotalia sp. 15, 358 Schwagerina craticulifera 90 Rotalia trochidiformis 541 Schwagerina gigantea 87 Rotaliconus 435 Schwagerina sphaerica var. karnica 82 Rotaliconus persicus 435 Schwagerina wanneri var. sutschanica 87 Rotaliidae 318, 431, 464, 478, 571 Schwagerinidae 35, 85, 86, 97, 98, 99, 102, 103, 111, 121 rotaliides of arose from fusulina in Late Carboniferous 102 Cretaceous 339– 344 of Late Carboniferous 102 Neogene 568–5 81 schematic features of 85 Palaeogene 453–4 84 septa in 103 Rotaliinae 432 Schwagerinids 103 rotaliide test 16 evolution of 103 Rotalina inermis 318 Sellialveolina sp. 385 Rotalioidea 340, 422, 453, 469, 478, 568, 571, 572 Semiendothyra 71 Rotalites trochidiformis 319, 433 Semiendothyra surenica 71 Rotorbinella 433, 478, 480 Semiinvoluta 175, 176, 177 Rotorbinella colliculus 433 Semiinvoluta clari 175 Rotorbinella skourensis 441 Seminovella 73 Rumaila 364 Semitextulariidae 49, 56, 57, 100, 104 Rupertia stabilis 555 Senalveolina 322, 324, 338 Rupertina 555, 571, 619 Senalveolina aubouini 324 Rupertina stabilis 619 Septabrunsiina 62, 64 Rupertininae 423, 454, 555 Septabrunsiina (Spinobrunsiina) ramsbottomi 64 Russia 9, 106, 185, 186, 187, 257, 264 Septaforschia 64 Russiella 82, 111, 119 Septaglomospiranella 62, 64 Russiella pulchra 82 Septatournayella 64, 108 Sabaudia 298, 333, 371 Septatournayella henbesti 64 Sabaudia capitata 371 Septatournayella? Conspecta 65 Saccaminopsis 56 Septatrocholina 207, 227, 233, 234, 247, 261 Saccaminopsis fusulinaformis 128 Septatrocholina banneri 207, 234, 261 Saccammina carteri 56 Serpukhovian– Bashkirian boundary 117 St Pietersberg 383 Serpula pusilla 173 Sakesaria 318, 434, 435, 479, 480, 541 Serraia 409 Sakesaria cotteri 435 Serraia cataioniensis 409 Sakesaria dukhani 434, 541 shallow marine foraminifera 176 Sakesaria ornata 541 Shanita 26, 96, 172 Sakmarella 99 Shanita amosi 96 Sakmarian 72, 77– 81, 84, 86–8 9, 91, 94, 95, 97, 102, 103, Sherbornina 439, 479 112, 119, 120, 137, 167, 178 Sherbornina atkinsoni 439 fusulinides distribution in Tethys from Visean Siberian traps 123, 124, 188 to 119–1 20 Sichotenella 78 fusulinides diversity in 120 Sichotenella sutschanica 78 schwagerinidae evolution 103 Siderina douvillei 312 Siderolites 314, 319, 320, 339, 340, 346, 347, 383, 384, 439 salinity crisis 184, 596 Siderolites calcitrapoïdes 319 Sanderella 211, 212, 236, 238, 264, 267 Siderolites calcitrapoides 383, 384 Sanderella laynei 211, 238, 264 Siderolites sp. 383 Satorina 211, 212, 236, 238, 267 Siderolites spengleri 383 Satorina apuliensis 211, 238, 267 Siderolites vidali 314 Saudi Arabia 203, 249, 251, 259, 264, 267, 270, 273, Silvestriella 439, 521 358, 374 Silvestriella tetraedra 521 Saudia 267, 391, 502 Simplalveolina 322, 324, 384, 385 Saudia discoidea 391, 502 Simplorbites 26, 309, 310, 382 Saudia labyrinthica 391 Simplorbites cupulimis 309 Scandonea 327, 402 Simplorbites gensacicus 382 Scandonea samnitica 327 Simplorbitolina 302, 374 Schlumbergerella 425, 556 Simplorbitolina (?) miliani 302 Schlumbergerina 452 Simplorbitolina manasi 302, 374 Schubertella 77, 81, 82, 97, 102, 110, 147 Siphonaperta 452 690 Subject Index Siphovalvulia sinemurensis 233 Spiraloconulus perconigi 232, 273 Siphovalvulina 26, 210, 212, 227, 229, 232, 233, 247, 248, Spirapertolina 325 253, 267, 297 Spirapertolina almelai 325 Siphovalvulina beydouni 267 Spirillina irregularis 55 Siphovalvulina colomi 229, 232, 233, 267 Spirillina subangulata 63 Siphovalvulina colomi biozone 229 Spirillinidae 179 Siphovalvulina gibraltarensis 229, 232, 267 Spiroclypeus 19, 26, 29, 412, 458, 464, 465, 498, 521, 570, Siphovalvulina sp. 232, 267 593, 619, 628 Siphovalvulina variabilis 210 Spiroclypeus leupoldi 521 Sirelella 381, 430, 479 Spiroclypeus orbitoideus 412 Sirelella safranboluensis 381, 430 Spiroclypeus tidoenganensis 628 Sirtina 310, 314, 559 Spiroclypeus umbonata 521 Sirtina orbitoidiformis 314 Spiroclypeus vermicularis 521 Skinnerella 85 Spiroconulus perconigi 218, 295, 361 Skinnerina 88, 103 Spirocyclina 291, 330, 331, 332, 361, 364 Skinnerina typicalis 88 Spirocyclina choffati 364 Socotraina 216, 232, 235, 240, 241, 248, 264, 276 Spirocyclinidae 391 Socotraina serpentina 216, 232, 233, 264, 276 Spirolina 218, 292, 405, 447, 451 Sogdianina 50, 130 Spirolina agglutinans 164, 218, 292 Sogdianina angulata 50, 130 Spirolina cylindracea 405 Solenomeris 425 Spiroloculina 404, 568 Solenomeris ogormani 425 Spiroloculinidae 403 Somalina 407, 444, 447, 517 Spiroplectammina mirabilis 65 Somalina hottingeri 517 Spiroplectammina tchernyshinensis 65 Somalina stefaninii 407 Spiroplectammina venusta 65 Sorites 25, 26, 404, 407, 447, 452, 517, 547, 549, 550, 567, Spiroplectamminoidea 224, 245 612, 619 Spiropsammia 389, 502 Soritidae 325, 405, 452, 547, 566 Spiropsammia primula 502 soritids 36, 37, 490, 566, 568, 585, 593, 595– 596, 598 Spiropsammia uhligi 389 Soritoidea 338, 339, 404, 446, 447, 451, 452, 563, 565, 566, Spiropsammiinae 389 567, 584, 585, 594, 595, 597 Spirorutilus 224 genera of Neogene 584 Sporadotrema 425, 454, 455, 491, 619 Sornayina 291, 330, 331, 332, 364 Sporadotrema cylindricum 619 Sornayina foissacensis 291, 364 Sporadotrema mesentericum 619 South Africa 255, 311, 483, 497, 523, 573, 589, 599, 600, Staffella 76, 77, 79, 97, 103, 108 604, 622, 628 Staffella discoides 79 South America 118, 191, 254, 255, 348, 350, 588, Staffella (Eostaffella) parastruvei 76 589, 597 Staffellidae 75, 77, 79, 97, 98, 101, 103, 116 South France 505 schematic features of 77 South Tibet 282 Steinekella 211, 212, 236, 267 Southeast Asia 565 Steinekella steinekei 211, 267 Southern Europe 236, 249, 258 Stenocyclina 416, 456 southern Iberia 203 stolon system 23, 33, 306, 307, 318, 420, 424, 430, 458, Southern Spain 270, 276 467, 473, 481, 553 Southern Turkey 251 Stomatostoecha 288, 289, 290, 361 southwestern Iberia 191 Stomatostoecha plummerae 289, 361 Spain 118, 164, 248, 270, 273, 280, 282, 283, 295, 346, 358, stratigraphic correlation 191 361, 364, 367, 371, 374, 384, 408, 432, 468, 508, 514, Streblospira 180 526, 529, 601, 631, 634 Streptalveolina 324 Sphaerogypsina 17, 18, 311, 425, 455, 470, 471, 491, Streptalveolina mexicana 324 536, 612 Streptocyclammina 218, 231, 280 Sphaerogypsina globula 470, 612 Streptocyclammina liasica 231 Sphaerogypsina globulus 612 Streptocyclammina parvula 280 Sphaeroschwagerina 82, 102, 103, 119, 149 Subalveolina 20, 324, 338, 385, 449 Sphaeroschwagerina carniolica 149 Subalveolina dordonica 324 Sphaeroschwagerina citriformis 149 Subalveolina sp. 385 Sphaeroschwagerina fusiformis 103 subepidermal networks 21 appearance in Tethyan realm 102 Sulcoperculina 310, 314, 315, 316, 317, 343, 456, 463 Sphaeroschwagerina pulchra 149 Sulcorbitoides 310, 315, 316, 343 Sphaerulina 103 Sulcorbitoides pardoi 316 Spinobrunsiina 64 Sumatra 145, 409, 520, 552, 615, 617, 619, 622, 628, 631 Spinoendothyra 71, 139, 231 Sumatrina 90, 103 Spinoendothyra phrissa 139 Sumatrina annae 90 Spinolaxina 64 Surinam 191 Spinothyra 72 Switzerland 208, 267, 270, 276, 358, 361, 371, 376 Spinotournayella 64 Syria 251, 264, 377, 508, 535, 538, 541 Spiraloconulus 217, 218, 232, 235, 240, 241, 252, 273 Syriana 222, 270 Subject Index 691 Syriana khouri 270 Textulariidae 3, 56, 105, 562 Syriana khouryi 222 textulariides 34, 36, 161, 203, 227, 234 Syzrania 91, 106, 107, 178 Cretaceous 329– 335 Syzrania bella 91 Neogene 581– 582 Syzraniidae 91, 106, 178 Palaeogene 445– 446 Syzriana 166 Textularioidea 224, 227, 228, 234, 245, 246, 329, Taberina 327, 381, 407 445, 581 Taberina bingistani 327, 381 of Jurassic 245–2 46 Taberina cubana 407 Textulariopsidae 224 Taiyuanella 84, 88, 99 Textulariopsis 224, 229, 232, 245, 270 Taiyuanella subsphaerica 88 Textulariopsis areoplecta 224 Takkeina anatoliensis 361 Textulariopsis portsdownensis 224 Tania 558 Textulariopsis sinemurensis 229, 270 Tania inokoshiensis 558 Textulariopsis sp. 232 Tansinhokella 412, 456, 465, 521, 523, 570, 628 Tezaquina 91, 166, 178 Tansinhokella sp 521 Tezaquina clivuli 91 Tansinhokella tatauensis 412, 521, 523 Thaumatoporella 229, 248, 267, 283 Tansinhokella yabei 521, 628 Thaumatoporella parvovesiculifera 248, 267, 283 Tanzania 249, 609, 622, 631 Thomasella 391, 502 Tarburina 327 Thomasella labyrinthica 502 Tarburina zagrosiana 327, 361 Tikhinella cannula 55 Tawitawia 562 Timanella 72 Tawitawiinae 562 Timidonella 216, 232, 233, 235, 240, 241, 276 Tayamaia 421 Timidonella sarda 216, 232, 276 tectonic reconstruction Titicites 98 Early Permian 123 Torinosuella 219, 234, 245, 280, 289, 331, 332 Late Devonian 115 Torinosuella peneropliformis 280 Late Pennsylvanian 120 Torinosuella peneroplliformis 234 Visean 118 Toriyamaia 79 Toriyamaia latiseptata 79 Tenisonina 556 Torreina 310, 311, 343 Tenisonina tasmaniae 556 Torreina torrei 311 Tethyan realm Tournaisian, recovery of shallow reefal environment fusulinide during Late Carboniferous at 119 during 114 Sphaeroschwagerina fusiformis appearance 102 Tournaisian– Visean boundary 114, 115 Tethys 25, 34, 36, 102, 112, 114, 115, 116, 117, 119, 120, Tournayella 61, 62, 64 177, 179, 180, 181, 182, 183, 184, 187, 188, 189, 203, Tournayella, chamber arrangement of 18 225, 228, 247, 248, 249, 251, 253, 257, 285, 329, 330, Tournayella costata 62 345, 347, 348, 350, 352, 353, 354, 355, 387, 392, 428, Tournayella discoidea 64 440, 444, 445, 446, 452, 453, 459, 461, 471, 493, 495, Tournayella (Eotournayella) jubra 63 497, 498, 499, 500, 566, 570, 571, 578, 580, 588, Tournayella moelleri 63 589, 591, 593, 595, 596, 599, 600, 601, 602, 603, Tournayella questita 64 605, 606 Tournayella segmentata 64 fusulinides distribution in 118–1 19 Tournayellida 7, 45, 47, 61, 108, 114, 179 peri- Gondwanan parts of 119– 120 Tournayellidae 49, 61, 62, 65, 100, 104 Tetrataxidae 49, 59, 100, 209 lineages 61– 62 Tetrataxis 59, 60, 107, 108, 128, 132, 136 schematic evolution of 62 Tetrataxis Bradyi 136 Tournayellina 64 Tetrataxis conica 60, 128, 132, 136 Tournayellina (Rectotournayellina) 64 Tetrataxis eominima 60 Tournayellina vulgaris 64 Tetrataxis (Globotetrataxis) elegantula 59 Tournayelloidea 100 Tetrataxis pusillus 128 Triadodiscidae 174 Tetrataxis subcylindricus 136 Triadodiscus 26, 174, 176, 177, 180, 181, 185, 201 Tetrataxoidea 59, 100 Triadodiscus eomesozoicus 201 Textularia 57, 224, 227, 232, 247, 390, 560, 562 Triadodiscus mesotriasica 185 Textularia antiqua 57 Triasina 175, 176, 177, 181, 196 Textularia barrettii 560 Triasina hantkeni 175, 181, 196 Textularia immensa 562 Triasina oberhauseri 175, 181 Textularia lirata 390 Triasinidae 175 Textularia sp. 232 Triassic 27, 34, 45, 54, 65, 66, 91, 92, 94, 95, 107, 121, 124, Textulariella 20, 298, 560 125, 161, 162, 163, 164, 165, 166, 167, 168, 169, 170, Textulariella cretosa 298 171, 172, 173, 174, 175, 176, 177, 178, 179, 180, 181, Textulariella minuta 298 182, 183, 184, 185, 186, 187, 188, 189, 190, 191, 192, Textulariellidae 560, 581 193, 194, 195, 196, 198, 201, 203, 205, 206, 207, 209, Textulariida 7, 14, 36, 45, 161, 176, 203, 204, 209, 285, 214, 218, 221, 226, 228, 247, 250, 251, 253, 256, 286, 286, 388, 543 287, 292, 293, 294, 296, 298, 307, 328, 389, 390, 407, evolved from Allogromiida during Cambrian 45 546, 551, 559, 587 692 Subject Index Triassic larger benthic foraminifera 161– 201 Turkey 107, 193, 204, 264, 267, 270, 273, 296, 361, 367, Allogromiida 161–1 62 381, 452, 502, 511, 514, 526, 534, 609 biostratigraphy and phylogenetic evolution 175– 181 Turriglomina 171, 177, 180, 188, 196 biostratigraphy 175– 177 Turriglomina conica 188 fusulinides of Triassic 179 Turriglomina mesotriasica 188, 196 involutides of Triassic 180–1 81 Turriglomina scandonei 188 lagenides of Palaeozoic- Triassic 177– 179 Turritellella mesotriasica 171 miliolides of Triassic 179– 180 Twaraina 407, 447, 517 Palaeozoic– Triassic lagenides 177–1 79 Twaraina seigliei 407, 517 development and evolution of superfamilies 162 Uhligina boninensis 423 disaster forms 176, 178, 180 Ukraine 189, 257, 356 Fusulinida 161–1 62 Umbellinidae 49 Involutinida 161–1 62 United Arab Emirates 273, 364, 377 Lagenida 161– 162 Uralinella 51 Miliolida 161– 162 Uralinella bicamerata 51 morphology and taxonomy of 161– 175 Uralinellidae 50, 100 palaeoecology of 182– 185 Urbanella 72 palaeogeographic distribution of 185– 192 Urgonina 214, 239, 300, 302, 334 Textulariida 161–1 62 Urgonina (Parurgonina) caelinensis 214, 239 Triassic– Jurassic extinction 190 Urgonina protuberans 302 Tribliolepidina 579, 580 Usloniidae 51, 100 Trillina howchini 402 Uviella 65 trimorphic life cycle, of larger benthic foraminifera 12– 14 Uviella aborigena 65 Trinidad 468, 502, 523, 532, 541, 599, 625 Vacuolispira 407, 409, 456, 467, 520 Triplalepidina veracruziana 523 Vacuolispira inflata 520 Tristix (Pseudotrislix) tcherdvnzevi 92 Vacuovalvulina 391, 392, 445 Triticites 16, 84, 85, 86, 88, 97, 99, 102, 103, 110, 113, 119, Valdanchella 300, 302 137, 141, 145, 151, 157 Valserina 302 appears at the end of Moscovian 119 Valserina broennimanni 302 from Wolfcamp formation 86 Valvulina 59, 61, 209, 214, 237, 273 keriothecal wall structure 102 Valvulina bulloides 59 Late Carboniferous 102 Valvulina sp. 273 palaeoecology of 110 Valvulina triangularis 214 Triticites patulus 141, 157 Valvulina youngi 61 Triticites ventricosus 137, 145 Valvulinella 61, 106, 128, 138 Trochamijiella 216, 217, 276 Valvulinella lata 128 Trochamijiella gollesstanehi 216, 276 Valvulinella tchotchiai 128 Trochammina 201 Valvulinella youngi 106, 138 Trochammina squamata 163 Valvulinellidae 60, 61, 100 Trocholina 26, 174, 175, 176, 177, 193, 206, 207, 208, 227, Valvulinidae 213, 237 232, 247, 253, 259, 264, 377 Valvulininae 213 Trocholina altispira 206, 377 Vanderbeekia 310, 314, 559, 575 Trocholina arabica 206 Vanderbeekia trochoidea 314 Trocholina cf. granosa 259 Vaughanina 310, 315, 317, 343, 347, 383 Trocholina conica 206, 208, 232, 259 Vaughanina cubensis 317 Trocholina granosa 232 Vaughanina sp. 343, 383 Trocholina multispira 193, 206 Vaughaninae 317 Trocholina palastiniensis 259, 264 Venezuela 303, 364, 420, 535, 602 Trocholina (Paratrocholina) eomesozoicus 174 Ventrolaminidae 209 Trocholina (Paratrocholina) oscillens 174 Verbeekina 20, 82, 85, 89, 97, 99, 103, 111, 121, 145 Trocholina permodiscoides 174 Verbeekina verbeeki 145 Trocholina transversaria 232 Verbeekinidae 75, 85, 89, 97, 98, 99, 103 Trocholina (Trocholina) biconvex 174 Guadalupian marks first occurrence of 103 Truncatulina grosserugosa 424 schematic features of 85 Trybliolepidina 475 Tubeporina 50, 52, 128 Verbeekininae 26 Tubeporina gloriosa 52 Vercorsella 298, 299, 333, 334, 371 Tubeporina magnifica 128 Vercorsella arenata 298, 299, 371 Tuberendothyra 72 Vercorsella camposaurii 371 Tuberitina 51, 52, 128, 131 Verneuilinella 297 Tuberitina bulbacea 52 Verneuilinidae 209 Tuberitina collosa 51 Verneuilinoidea 209 Tuberitinidae 49, 51, 100, 114 Verneuilinoides 297 Tuborecta 92, 106 Verneulinidae 347 Tuborecta vagranica 92 Verneulinoides 330 Tunisia 154, 179, 452 Verseyella 392, 396, 444, 446, 502 Turborotalia 296, 369 Verseyella jamaicensis 502 Subject Index 693 Victoriella 424, 425, 454, 455, 491, 536, 617, 619, 622 Wilfordia 425, 471, 491, 514, 517, 558, 575 Victoriellidae 423, 555 Wilfordia sarawakensis 425, 514, 517 Victoriellinae 423, 454 Wolfcamp Vidalina 207 beds 86, 137, 141, 145, 150, 151 Vietnam 580, 625, 634 formation 137, 145, 151 Visean, palaeogeographic and tectonic hills 86, 137, 150 reconstruction of 118 Wrangellia Terrane 188 Viseina 65 Wrangellia volcanic sequences 188 Vissariotaxis 60, 106, 140 Yabeina 85, 90, 97, 98, 103, 111, 156 Vissariotaxis cummings 106 Yabeina globosa 156 Vissariotaxis cummingsi 140 Yaberinella 407, 447, 452, 508, 514, 517 Voloshinoides 296, 297, 298, 333, 371 Yaberinella hottingeri 514 Voloshinovella 298 Yaberinella jamaicensis 407, 508, 517 Wadella 424, 454, 491 Yangchienia 82 wall structures 14, 74, 90, 92, 99, 178, 345 Yangchienia iniqua 82 in bradyinidae 96 Yemen 276, 498, 499, 601 characterise fusulinoidea 96 Yugoslavia 121, 179, 197, 267, 282, 358, 514 development in fusulinoidea during Late Carboniferous Zagrosella 289 and Permian 101 Zagrosella rigaudii 289, 361 Zarodella 78, 102 Wanganella ussuriensis 94 Zarodella zhamoidai 78 Wedekindellina 85, 145 Zekritia 327, 385 Wernlina 164 Zekritia langhami 327, 385 Wernlina reidae 164 Zellerinella 72 West Iraq 264, 273, 280 Zellia 89, 103, 119, 157 western Africa 191 Zellia heritschi mira 157 Western Europe 254, 606 Zotheculifida 390, 502 Western Mediterranean 203, 229, 234, 595, 599 Zotheculifida lirata 502 Second Edition Evolution and Geological Significance of Larger Benthic Foraminifera  is a unique, comprehensive reference work on the larger benthic foraminifera. E vo l u t i o n this second edition is substantially revised, including extensive re-analysis of  the most recent work on cenozoic forms. it provides documentation of the biostratigraphic ranges and paleoecological significance of the larger foraminifera, which is essential for understanding many major oil-bearing sedimentary basins. in addition, it offers a paleogeographic interpretation of a n D G E o l o G i c a l the shallow marine late Paleozoic to cenozoic world. Marcelle K. BouDagher-Fadel collects and significantly adds to the information already published on the larger benthic foraminifera. new research in the Far East, the Middle East, South africa, tibet and the americas has provided fresh insights into the evolution and paleographic S i G n i F i c a n c E significance of these vital reef-forming forms. With the aid of new and precise biostratigraphic dating, she presents revised phylogenies and ranges of the larger foraminifera. the book is illustrated throughout, with examples of different families and groups at the generic levels. Key species are discussed and their biostratigraphic ranges are depicted in comparative charts (available o F l a r G E r B E n t h i c separately online). PROF MaRcelle K. BOuDagheR-FaDel is a Professorial research Fellow in the office of the vice-Provost (research) at ucl. She graduated with a BSc from the lebanese university and has an MSc and PhD from ucl. She has an extensive publication record, having written three major books and F o r a M i n i F E r a over 130 papers. She is an established consultant with several oil companies, lectures widely, and supervises PhD students from around the world. Marcelle K. BouDagher-Fadel FRONT AND BACK COVER IMAGES: Marcelle K. BouDagher-Fadel Cover design: Rawshock design www.ucl.ac.uk/ucl-press Evolution anD GEoloGical SiGniFicancE oF larGEr MarcEllE K. BouDaGhEr-FaDEl BEnthic ForaMiniFEra