u> -Z in •** z _ JTION NOIinillSNI_ NVINOSHIIWS S3iyvy8H LIBRARIES SMITHSONIAN INSTITUTION NOlin. \ Z in o -^vjr ~ o 5 2 J 2 j z _i M8n LIBRARIES SMITHSONIAN INSTITUTION NOlinillSNI NVINOSHIIWS S3IHVHan LIBRA z r~ 2 r- > 30 (- ™ ' 2 m co m co m co - co ± — w \ z tn jtion NouniiiSNi NViNOSHims saiyvyan libraries Smithsonian institution Noun Z _ CO Z , y to 2 co ^ E . . <■* ^ E ^ \ b. y a n_ u b r ar i es Smithsonian institution NouniiiSNi nvinoshiiws sgiyvaaii libra -?> tn c/> — . c/> — v}S\ / « /^SOf^/Sv Ui / >s XlMSOV^S. UJ iJi\ • * l y^TifwyTS. ^ O xgN.pc^ ~ W" O 5 ^ 5 N^%j JTION ~ NOlinillSNI^NVINOSHimS^Sa lyvyail^LIBRAR IESZSMITHSONIAN~'lNSTITUTIONZNOIin z — ” > z r~ 2 f~ 2 m ' \x> -7- m v/\os*sz ^ m — CO — to Z o _ y 8 IT LI B RAR I ES^SMITHSONIAN ^INSTITUTION NOlinillSNI _NVIN0SH1IWS _ S3 I y VH 8 |"| _LI B RA < v''W* - 'v!^fNW^9/ t Z , S s > s X^vos^ > in c/> *'* 2! to — 2! CO TION NOlinillSNI _ NVINQSH1IWS S3iyvy8IT LIBRARIES SMITHSONIAN INSTITUTION NOlin. z \ y? r-. 5 CO ^ - co < i " O 5 2- ' O 2 Han LIBRARIES SMITHSONIAN INSTITUT!0N~N0liniJlSNrNVIN0SHlllNS~S3iyvy8n~LIBRA f z y r- z r- ^ m 2 , . ,cd 2 m ^>v 2 ro m x^vpc^ — m -^2i^ ~ xvo7iB2X m X v> n^u^/ m co — co £ — to \ £ to JTION NOlinillSNI NVIN0SH1IINS S3IHVyail LIBRARIES SMITHSONIAN INSTITUTION NOlin. CO 2 to Z CO 2 CO S •- t ^ 4KW E i s z -I .y ,.:/ V?* Z «\> =? /&f?T>&v z £ I (Ir SI) ° il6k x w ^ § W £ i IW?# g *% § % * Z CO Z CO * Z ^ V 2 y 8 n_ LI B RAR I ES SMITHSONIAN INSTITUTION NOlinillSNI NVIN0SH11WS S3iyvyon LIBRA CO ^ CO — CO z K > ' ^ X£V OS^iX > -C- ,W £" *2 ^ — z to N z to noshxiws samvaan libraries smithsonian institution NoixnxixsNi_NViNOSHxms sh to _ = to zz _ - X'lZtorAkwj/ — "*•' "s&x •- \vxm35','xy — vc/w^xy wvsk m -vjv. to xSqxn^z m E2 x^msgx rn xfo dcz z f11 CO « to . ^ co — to INOSHXIWS SBIBVyen LIBRARIES SMITHSONIAN INSTITUTION^ NOIXnXIlSNI NVINOSHXIWS S3 il 8 JSs x ^Hk' I fJ^jl I 1 ^ 5 s ^ 5 > 'WA 2 ^ (/) * Z (/) 2 CO ^ ITHSONIAN INSTITUTION NOlinillSNI NVIN0SHX11MS S3 I a VH 8 ll_ LI B R AR I ES SMITHSONIAN _ IN — CO ~ _ co -p ^ ^ / Aj , , , Jpw c 5 5 ^ 5 U| 'W o INOSHIIWS^S 3 I aVBSIl^UBRARI ES^ SMITHSONIAN^INSTITUTION ^NOIXnXIXSNI^NVINOSHXIWS^S) O /TXtTtjTn. Z^o~aSv. O m zf^o^oX 5 m m £ * m ^ X_;v^z rn *— (f) ^ iithsonian'institution NouniiiSNi~NviNOSHims S3 1 a va 8 n li b r ar i e$ smithsonian ir -r ■ rn Z CO Z •«. ~f ^ fii ' x Wv S ' N^v > " S X^osv^y > ^ g 2 nNOSHlUNS^0 S3 I ava 8 nZLI B RAR I ES^SMITHSONIAN INSTITUTION NOIinillSNI_NVINOSHillMS S to - to - ^ to — z \ “2 uj xS5Rn^s. ^ ui ,., ^ t\ _i A^ma^SK — mwAA O z 73 rn Dt^z ^ m S2 XvgjiTs^y m ^^dcJX ^ C/) ~ CO ^ CO — tflNOSHXIINS S3 lava 8 11 LIBRARIES SMITHSONIAN INSTITUTION NOIXnXIXSNl NVINOSHXIIMS S to z .,. w z >v. #0# o |Sg% 3 /i^l i 3 5 = CO CO 4 £ 5 z _ ' ^ ^ > ^ co ^ z co v z co VIITHSONIAN INSTITUTION NOIXOXIXSNI NVINOSHXIWS SBiavaail LIBRARIES SMITHSONIAN _ I — to — i3[/\\an>I L9 31Vld H -/OA ‘AS0/O}U0dV/DJ RUDWICK: TRIASSIC BRACHIOPODS THECOSPIRA AND BACTRYNIUM 339 indentation are often coalesced into a single ridge or ‘lateral septum’; but here too the best specimens show that the ridge is double in origin (PI. 67, fig. 4; PI. 68, fig. 1). Around the distal end of each lobe the ridge may be clearly marked or may merge insensibly into the peripheral part of the valve surface. The groove likewise is variably expressed; it may be clearly differentiated from the central area of the lobes or may merge insensibly into it. Where the inner edge of the groove is distinct, it is shown by a dashed line in text-fig. 5. The central area of the lobes is differentiated, at least in the best specimens, by a distinctly pustular surface (PI. 67, figs. 2-5). The posterior extremity of the groove/ridge structure is not entirely clear. The most posterior pair of secondary lobes are generally short, and lie antero-lateral to the cardinal process. At first sight it might appear that the groove runs into the base of the cardinal process (PL 66, figs. 3, 8, 11; cf. Williams 1965, p. H521). But this is an illusion. A specimen of Zugmayer’s, ground down to expose the ridges, shows clearly that the ridges of left and right sides fused with one another in the median plane to form a forwardly projecting V-shaped structure (PI. 67, fig. 6). Zugmayer himself referred to this as a bridge, and this interpretation appears to me to be correct. In other words, at points antero-lateral to the cardinal process the ridge rises away from the floor of the valve and crosses the postero-median cavity as a bridge-like lamella (text-fig. 4a). In the best-preserved specimens (PI. 66, fig. 3; PI. 68, fig. 1) the ridges appear to end abruptly, but these are evidently breakage surfaces at the points where the ridges became a free lamella. It is clearly distinct from the ‘socket plate’ running into the base of the cardinal process (text-fig. 2a). Brachial axis o/Thecospira. The entire brachidium of Thecospira, like that of any other articulate, is clearly an outgrowth of the secondary-layer material of the dorsal valve. As such, it must have been sheathed by, and secreted by, an extension of the outer mantle epithelium (cf. Williams 1956). As with the spiral brachidia of spiriferides, geo- metrical corsiderations alone are sufficient to prove that this epithelium must have had the power of resorption as well as secretion, or else the brachidium could not have grown in size while retaining its form and relative position. In their reconstruction of the lophophore of spiriferides, Williams and Wright (1961) make the assumption that the growing tips of the brachia necessarily remained together in the mid-line, on the jugum, throughout ontogeny. But this assumption is not fully supported even by their cited analogy of plectolophes in living brachiopods; in Gryphus vitreus, for example, the growing tips are far apart from one another at the distal end of the median coil. Even supposing that a complete jugum was present in Thecospira, therefore, I see no justification for reconstructing a ‘deuterolophe’ (i.e. doubled brachial axes) on the spiral lamellae. The simpler reconstruction of an ordinary spirolophe seems to me more plausible, in view of the close resemblance between the form of the spiral brachidium and that of the spiral lophophores of living rhynchonellides. On this reconstruction each whorl of the lamellae would have borne a corresponding single whorl of the brachial axis, and the growing tips of the brachia were at the tips of the lamellae. As suggested previously (Rudwick 1960, text-fig. 6d), the brachial axis is shown, conjecturally, nestling within the ‘gutter’; if this is correct the filament-row, at least near its base, would have projected outwards around each whorl. If the lophophore 340 PALAEONTOLOGY, VOLUME 11 was supported throughout its length by the brachidium, the great brachial canal, which in most living articulates acts as a hydrostatic skeleton, may have been vestigial or absent; and it is so shown, conjecturally, in the reconstruction given here (text-fig. 3b). Proximally, by homology with living brachiopods, the axis would have continued towards the mid-line, lying adjacent to the crural processes or jugum. The mouth would have been situated in the mid-line. At this point, if the basic orientation of the lopho- phore was like that of all living brachiopods, the filament-row must have been on the posterior or ventral side of the mouth and food-groove, and the lip on the anterior or dorsal side. Then, by tracing the axis laterally on to the spirals, it follows by simple topology that the filament-row must have been on the dorsal side of the food-groove, and the lip on the ventral, on each of the outwardly facing whorls of the spirolophe. The brachial axis is similarly twisted in the proximal parts of the spirolophes and plecto- lophes of living brachiopods (Rudwick 1962, text-figs. 7c, 9c, 12b). Brachial axis of Bactrynium. The mode of growth of the lobate apparatus can be in- ferred quite simply (text-fig. 6). A section of Zugmayer’s, cut longitudinally through the dorsal valve parallel to the median plane, shows clearly that the lobate apparatus grew by simple accretion. The ‘lateral septa’ (i.e. the coalesced doubled ridges in the lateral indentations) remained in the same absolute position during the growth of the valve, for their growth tracks can be seen projecting through the thickness of the valve perpendicu- lar to its outer surface (text-fig. 2b; PI. 68, fig. 9). Since the valve grew not only in thick- ness but also in area, any given lobe must therefore have changed in relative position during ontogeny, becoming progressively more posterior. The distal ends of the lobes are near the valve edge in specimens of all sizes: therefore the lobes must have grown in EXPLANATION OF PLATE 68 Bactrynium, Thecospira and other brachiopods for comparison. All figures X 6. All except fig. 9 whitened with ammonium chloride. Figs. 1, 2. Bactrynium bicarinatum Emmrich; ventral and posterior views of fragment of dorsal valve; see also text-fig. 2a. Kossener Schichten of Diirnbach (PIUW; figured Zugmayer 1880, pi. 2, fig. 18). Fig. 3. Eudesella mayalis (Deslongchamps), from the ‘Couche a Leptaena’ (Upper Lias), of May, Calvados, France; interior of dorsal valve (SM.F.20273). Figs. 4, 5. Davidsonella sinuata (Deslongchamps), from the same locality and horizon; postero-ventral and ventral views of dorsal valve (SM.F.20199). Fig. 6. Moorellina leptaenoides (Deslongchamps), from the same locality and horizon; interior of dorsal valve (SM.F.20258). Figs. 7, 8. Elliottina deslongchampsi (Davidson), from the Middle Lias of May, Calvados, France; ventral and postero-ventral views of dorsal valve (SM.F.20288). Fig. 9. Bactrynium bicarinatum Emmrich, longitudinal section of dorsal valve (and part of ventral), showing growth tracks of lateral septa; see also text-fig. 2b. (PIUW; figured Zugmayer 1880, pi. 2, fig. 32.) Locality unknown. Fig. 10. Lyttonia sp., internal surface of part of dorsal valve, for comparison with Bactrynium. Sandy limestone beds above Middle Productus Limestone (Permian), Banschang, East of Chideru, Salt Range, Pakistan (Yale University, Peabody Museum, 22954). Figs. 11, 12. Thecospira sp. Dorsal views of ventral valve alone and of both valves as found in loose association, to show pseudodeltidium, articulation and radial costellae. Raibler Schichten, Rumerlo, near Cortina d’Ampezzo, Italy (R. Zardini coll.). Fig. 13. ‘ Thecidea sp.’ (Thecospira?), dorsal view of shell, to show convex pseudodeltidium, from Helenenthal, Baden (PIUW; figured Zugmayer 1880, pi. 2, fig. 41). Palaeontology, Vo I. 11 PLATE 68 12 13 RUDWICK, Bactrynium, Theeospira and comparative species RUDWICK: TRIASSIC BRACHIOPODS THECOSPIRA AND BACTRYNIUM 341 length during ontogeny, extending laterally as the valve increased in breadth. At the same time the valve was also growing in length, and therefore the most anterior secondary lobe must have become progressively further from the anterior edge of the valve. But the most anterior secondary lobes, and the primary lobes themselves, are near the anterior edge in specimens of all sizes: therefore new secondary lobes must have been text-fig. 6. Diagrams to show inferred mode of growth of ‘lobate apparatus’ of Bactrynium (a, b), compared with Thecidiopsis (c, d). Figs. A, c show course of lobed groove and ridge; figs, b, d show corresponding growth tracks of axes of lobes ; lobes are numbered consecutively in order of derivation from the persistent primary lobes P. As dorsal valve grows in size from AA to BB, lobes grow in length but do not change in absolute position; new lobes (no. 8 in fig. a, no. 5 in fig. c) are budded off from persistent primary lobes. budded off at intervals during ontogeny from the pair of persistent primary lobes. Thus in text-fig. 6 a, during growth of the valve margin from A A to BB, secondary lobes 1 to 7 extended laterally while remaining in the same absolute position on the valve, while the primary lobe extended forwards and budded off the new secondary lobe 8. The whole process of growth of the lobate apparatus could have been achieved simply by differen- tial rates of secretion of shell material over the surface of the valve as it increased in thickness and over-all size: there is no need to postulate any shell resorption (except possibly on the posterior edge of the ‘bridge’). 342 PALAEONTOLOGY, VOLUME 11 This analysis of growth justifies the terminology so far used in a purely descriptive sense. The primary lobe may be so called because it must have persisted throughout onto- geny in a submedian position, growing progressively forwards. The secondary lobes must all have been derived from the primary lobe by lateral budding, in order from the most posterior lobe forwards. The numbering of the secondary lobes thus corresponds to their order of appearance (text-fig. 6b). The over-all effect of the growth of the lobate apparatus was to keep the internal surface of the valve fully covered (except in the peripheral zone) with a series of lobes and indentations of a constant standard width, whatever the absolute size and area of the valve. There is a close analogy between the lobate apparatus of Bactrynium and the brachia- bearing grooves and ridges in living Thecideacea. On the inner surface of the dorsal valve there is a lobed groove, two-lobed in Thecidellina, four-lobed in Lacazel/a (Elliott 1965, fig. 742, 3a; 744, 2a). On the outer side the groove is separated by a conspicuous slope from an outer ridge, though the latter may not be clearly distinguishable from the peripheral zone of the valve. Postero-laterally the outer ridge extends freely above the postero-median cavity, projecting towards the mid-line, or it may fuse with that pro- jecting from the opposite side to form a complete bridge over the cavity. This bridge is clearly separate from the ‘socket plates’ and cardinal process. In the median indentation the ridges of left and right sides gradually coalesce and project backwards as a spike overhanging the postero-median cavity. Indeed in Lacazella all the ‘septa’ in the indentations may overhang in this way at their inner ends. Internal to the lobed groove is a second, inner ridge, the so-called ‘brachial ridge’; within each lobe this is clearly double in origin, but may be coalesced into a single ridge running along the axis of the lobe. This inner ridge, like the outer, may project freely so as to overhang the postero- median cavity. Its crest is often distinctly pustular. Although the anatomy of living thecideaceans remains imperfectly known in many respects, it is at least clear that the brachial axis of the lophophore is attached throughout its length to the dorsal mantle, and that it lies in the lobed groove. This can be seen in both Thecidellina (cf. Elliott 1965, fig. 742, 3a) and in Lacazella (Lacaze-Duthiers 1861, pi. 3, fig. 1). The frontal side of the axis, bearing the lip, faces inwards, towards the inner ‘brachial’ ridge; while the outer side, bearing the filament-row, faces outwards, towards the outer ridge. The distal growing tips of the brachial axis are located close together at the tip of the median indentation (Lacaze-Duthiers 1861, pi. 2, fig. 7). The mouth is located medially, on the anterior side of the bridge over the postero-median cavity. The arrangement of the lobes is such that the lophophore of Thecidellina is a schizolophe, while that of Lacazella is a simple (four-lobed) ptycholophe. (Williams’s and Rowell’s (1965, p. H38) novel definition of the trocholophous stage, to include all lophophores in which the filaments are arranged in a single series, seems to me to be unwarranted: lophophore growth stages were originally defined by the arrangement of the brachial axis, and should remain so.) The brachial axis of Bactrynium may be reconstructed in a comparable position, attached to the dorsal mantle and lying in the lobed groove, with the filament-row facing the outer bounding ridge and the lip facing the inner pustular area (text-fig. 7d). Then proximally, at the mouth, the axis would have had the normal orientation, with the filament-row on the posterior or ventral side of the mouth and the lip on the anterior RUDWICK: TRIASSIC BRACHIOPODS THECOSP1RA AND BACTRYNIUM 343 or dorsal side. Thus the lophophore would have been a complex ptycholophe, with up to twenty or more lobes (text-fig. 4b). It may be assumed that the lophophore would have grown, like those of all living brachiopods, only by the lengthening of the brachial axis and the formation of new filaments at the extreme tip of the brachia, in the median indentation. Any given filament would therefore have shifted continually in its relative position during the growth of the lophophore. The growth zones of the lophophore and those of the supporting lobate apparatus were thus entirely distinct from one another. text-fig. 7. Lophophore and inferred feeding mechanism of Lacazella (a-c) and Bactrynium (d). a, dorsal valve of Lacazella with lophophoral filaments in contracted state, as illustrated by Lacaze- Duthiers (1861, pi. 3, fig. 1); b, c, lophophore of Lacazella in inferred feeding orientation, based on Lacaze-Duthier’s illustrations of opened shell (1861, pi. 1, figs. 2, 7) and on analogy with Megathiris (Atkins 1960) ; d, reconstruction of brachial axis and filaments of Bactrynium, shown as section through two lateral lobes; note brachial axis (dark stippled) in groove in surface of dorsal valve (light stippled), with filaments arching over ‘lateral septa’ to form ‘tunnels’; arrows show water currents and food currents. FEEDING MECHANISMS Some elements of the foregoing reconstructions of the lophophore are necessarily conjectural. But of the assumptions made, only three are essential for the validity of the functional interpretation given here : (n) that the basic anatomy of the lophophores of Thecospira and Bactrynium resembled that of all living brachiopods; (b) that their basic orientation relative to the body likewise resembled that of all living brachiopods; and ( c ) that in each genus the grooves in the skeletal apparatus bore single rows of the brachial axis, so that the lophophore was spirolophous in Thecospira and ptycholophous in Bactrynium. For the reconstruction of the feeding mechanisms, i.e. of the lophophores 344 PALAEONTOLOGY, VOLUME 11 in operation, one further assumption is necessary: ( d ) that the basic functional mechan- isms of the lophophores were normal, i.e. that they resembled those of all living brachio- pods (Rudwick 1962, and references therein). These assumptions of anatomical and physiological uniformity are not of course capable of direct proof. But it is methodo- logically a sound procedure to begin by testing them indirectly, by determining whether they give a consistent and intelligible explanation of the morphology of the fossils. Only if they fail to do so should departures from uniformity be postulated (cf. Rudwick 1964). For effective ciliary filter-feeding in a brachiopod, it is essential that the tips of all the filaments (except those projecting at the gape and delimiting the apertures) should touch the mantle surface or the tips of other filaments. Only by so doing can the lopho- phore divide the mantle cavity into separate inhalant and exhalant chambers and thereby prevent wasteful re-circulation and re-filtering of the water (Rudwick 1960). The possible ways in which a pair of linear brachia can divide the mantle cavity is limited by inherent topological considerations, and only three essentially different arrangements are known in living brachiopods (Rudwick 1962). The spirolophe and ptycholophe, here posulated for Thecospira and Bactrynium respectively, are two of these three known arrangements. Current-system o/Thecospira. A topological argument developed for the reconstruction of any spirolophe (Rudwick 1960) can be used to infer the attitude of the filaments in Thecospira. Inhalant and exhalant chambers could not be isolated unless the filaments on the spiral brachia touched the brachial axis of either the next proximal or the next distal whorl. But on the latter assumption the filaments on the first or most proximal whorl would have touched the axis of the second whorl, and no filaments would have been available to bridge the gap between the axis of the first whorl and the dorsal mantle. Therefore only the former alternative is feasible. On the distal whorls, the tips of the filaments would thus have touched the brachial axis of the next proximal whorl; on the medial part of the proximal whorl they would have touched either the tips of the corresponding filaments on the other brachium, or else the ventral mantle; anteriorly on this whorl, they would have passed across the gape, projecting freely as apertural filaments; and laterally they would have touched the dorsal mantle. Posteriorly, in the most proximal part of the lophophore, they would have touched the lateral body wall and finally the anterior body wall or ventral mantle (text-figs. 8, 9). Only by this precise orientation of all the filaments could the spirolophe of Thecospira have divided the mantle cavity into separate inhalant and exhalant chambers; by no other arrangement is it topologically possible for this essential functional condition to have been met. Of the only two possible varieties of the spirolophous system, the orientation of the brachidial lamellae thus shows that the interior of the spirals must have been an exhalant space, so that Thecospira must have operated the ‘ exhalant ’ or ‘ Spirifer' type, which has also been inferred for most spiriferides (Rudwick 1960) and is known in the living Discinisca (Paine 1962). The inhalant chamber in Thecospira would thus have comprised the ventral part of the mantle cavity, external to the spirals; while the exhalant chamber would have occu- pied the dorsal part, together with the interior of the spirals, and would also have included a narrow exhalant space between the proximal filaments and the body wall. RUDWICK: TRIASSIC BRACHIOPODS THECOSPIRA AND BACTRYNIUM 345 text-fig. 8. A. Reconstruction of T/iecospira, lateral view, cut along median plane. Inhalant water (tailed arrows) is seen entering laterally and circulating around exterior of spiral brachium; exhalant water (tailless arrows) is seen emerging from interior of spiral and leaving mantle cavity by a median aperture. To clarify the current system, brachial axes are shown more slender, and filaments more widely spaced, than in reality. Digestive glands omitted. b. Similar reconstruction, with shell cut in plane parallel to median plane on right side, but with right brachium preserved intact (water currents are those that would operate with whole shell intact); showing crus embedded in lateral body wall, through which muscles and gut are faintly visible. For further explanation see text. Given this orientation of the filaments, the current system of Thecospirci follows necessarily (text-figs. 8, 9). Water must have entered the mantle cavity laterally, being drawn down into the inhalant chamber by the pumping action of the filaments removing water into the exhalant chamber. After being filtered between the filaments on the spirals, the water would have flowed as an exhalant current up the axis of each spiral, and then forwards to leave the mantle cavity medially. Water filtered by the most proximal filaments would have flowed along the narrow space between those filaments and the body wall, emerging to join the main exhalant current on either side of the dorsal part of the body wall (text-fig. 9, current ‘X’): a similar current can be observed in living articulates (cf. Rudwick 1962, text-figs. 7c, 9a, c, 12b). Thus the morphology of the brachidium of Thecospirci is consistent with a ciliary feeding mechanism essentially similar to that of living brachiopods. The rate at which a lophophore can pump water through a mantle cavity is roughly proportional to the total area of the filament-row. On a spirolophe, the rate is therefore approximately proportional to the area of the roughly conical surfaces which are formed by the spiral brachia when all the filaments are in their natural orientation. Thus in a a C 5586 346 PALAEONTOLOGY, VOLUME 11 a spire-bearing brachiopod the larger the conical ‘framework’ formed by each spiralium, the greater the filtering capacity of the brachium must have been. For a shell of given size, the most effective arrangement of the spiralia would be that which most fully occupies the available space of the mantle cavity, while leaving sufficient space for the circulation of the water currents outside the spirals. text-fig. 9. Reconstruction of Thecospira , ventral view of dorsal valve, and dorsal part of ‘body’, with lophophore (right brachium cut off near base to clarify course of exhalant currents). Other conventions as in text-fig. 8. The observed arrangement of the brachidium in Thecospira approximates to this paradigm. The spiralia are so closely moulded to the available space that when the shell was closed the filaments must have been almost or quite in contact with the ventral mantle. When the shell opened, however, the brachidium and lophophore, being attached rigidly to the dorsal valve, would have been raised clear of the ventral mantle, thus allowing the circulation of water outside the spiral brachia. At the degree of opening selected for illustration here (text-figs. 8, 9), the circulation RUDWICK: TRIASSIC BRACHIOPODS THECOSPIRA AND BACTRYNIUM 347 space outside the spiral brachia approximates to that observed outside the spirolophes of the living rhynchonellide Notosaria and the living Crania (Rudwick 1962, text-fig. 7; Atkins and Rudwick 1962, text-fig. 1). If the dorsal valve opened through a wider angle than that shown in this reconstruction, the same current system would have operated, except that the lateral inhalant apertures would have coalesced anteriorly, on the ventral side of the exhalant aperture, to form a single inhalant aperture. The functional advantage of developing a brachidial support for the spirolophe is not entirely clear. A possible explanation is that a calcareous support for the brachial axis takes up less space than the rather bulky hydrostatic and muscular skeletons seen in the spirolophes of living brachiopods (e.g. Atkins 1963, fig. 8a). Consequently the successive whorls of the spiral brachia can be set more closely together without substantially impeding the flow of water; and thus a greater area of filaments can be accommodated within a mantle cavity of given form and volume. But, on the other hand, this increased efficiency of the lophophore is obtained at the expense of having a supporting structure that is less flexible and more vulnerable to damage than is a hydrostatic skeleton. Thus, a spiral brachidium may on balance be advantageous in some circumstances; but, of course, it is only an available option in those brachiopods that have developed the power of resorption in the epithelia of the mantle. Current-system of Bactrynium. In a schizolophe or ptycholophe the only effective arrangement of the filaments is for those in the indentations to be flexed abfrontally so that their tips are contiguous. Thus each indentation forms a ‘tunnel’ of exhalant water, blind-ended proximally but opening peripherally. Such an orientation is in fact known in all living species in which schizolophes or ptycholophes have been observed in operation. In these lophophores the filaments are flexed abfrontally all along the brachial axis. In a schizolophe the tips of the filaments touch each other across the median indentation, forming a median ‘tunnel’ which is blind-ended posteriorly but open anteriorly; this has been observed in Argyrotheca and Pumilus and in the schizolophous growth-stages of Notosaria and Waltonia (Atkins 1960, fig. 2; Rudwick 1962, fig. 6, 7a, 9a). The schizolophe of Thecidellina has not yet been described in operation. In a ptycholophe there are similar ‘tunnels’ in the lateral indentations as well; this has been seen in Megathiris (Atkins I960, fig. 6). Lacaze- Duthiers (1861) only observed Lacazel/a with the filaments contracted frontally (text- fig. 7a), but their probable natural orientation can be inferred (text-figs. 7b, c) by analogy with Megathiris. In all these schizolophes and ptycholophes the space near the dorsal valve surface is divided into two parts; the space corresponding to the lobes is filled with inhalant water, whereas the space corresponding to the indentations is filled with exhalant water. Water is filtered through the filament-row around the periphery of the lophophore, but also passes into the blind-ended tunnels in the indentations, from which it emerges peri- pherally (cf. Atkins 1960, fig. 6; Rudwick 1962, fig. 6). The wide gape between the valve edges of Megathiris (and probably LacazeUa also) is divided in effect into two apertures. The exhalant aperture extends all round the gape as a fairly narrow zone near the dorsal valve edge; the inhalant aperture occupies the remainder of the gape, nearer the ventral valve but not extending as far posteriorly. The apertures are separated only by the tips of the filaments around the periphery of the 348 PALAEONTOLOGY, VOLUME 11 lophophore. In principle there might be re-circulation around these apertural filaments; but in practice the wide opening of the shell exposes the periphery of the lophophore to sufficient external currents and turbulence to render any re-circulation insignificant. It is reasonable to reconstruct the filaments of Bactrynium in a similar orientation, forming ‘tunnels’ across each of the indentations (text-fig. 7d). Then its current system would also have been similar (text-fig. 10). Water would have been drawn towards the dorsal valve surface, and would then have been filtered either through the filament-row around the periphery or else into the tunnels; within each tunnel an exhalant current would have flowed laterally or anteriorly to emerge near the edge of the valve. Thus in effect the part of the gape nearest the dorsal valve edge would have been an exhalant A B text-fig. 10. Reconstruction of Bactrynium in feeding position, showing inferred orientation of filaments, and consequent paths of inhalant and exhalant currents, a, anterior view; b, lateral view. aperture, and the rest of the gape a wide inhalant aperture. Unless Bactrynium lived in exceptionally still water there would have been little risk of any substantial re-circulation of the water. As in Thecospira, therefore, the observed morphology of Bactrynium is consistent with a ciliary feeding mechanism essentially similar to that of living brachiopods. On this interpretation Bactrynium developed the ptycholophous alternative to a greater degree of elaboration than in any living brachiopod. This may be related func- tionally to its greater size. Observations on living brachiopods of various species and at various growth stages suggest that the pumping capacity of filaments is roughly RUDWICK: TRIASSIC BRACHIOPODS THECOSP1RA AND BACTRYNIUM 349 constant, as might be expected of a property dependent on a mechanism (i.e. the ciliary action) operating on the cellular level (Rudwick 1962, p. 611). But the pressure dif- ference (between inhalant and exhalant chambers or spaces) at which a lophophore can operate is limited by the ability of the slender and unfused filaments to withstand the pressure and remain in position. Therefore if a blind-ended tunnel formed from two apposed stretches of filament-row were below some critical width, the rate at which water was being pumped into the tunnel might exceed the rate at which it could escape down the tunnel, and the water might then force the filaments out of position and flow back into the inhalant space. On the other hand it is clearly advantageous to the effi- ciency of the whole ciliary feeding mechanism that each tunnel should be as narrow as possible without incurring the problem mentioned above: for only so can the maximum total length of filament-row, and hence the maximum filtering capacity, be accommo- dated within a shell of given size. There must, therefore, be some optimal width for the tunnels, which would be the same for any normal ptycholophe. Similarly, it is clearly advantageous that the inhalant ‘troughs’ between the exhalant tunnels should also be parallel-sided and of some uniform width, so that as many as possible can be provided in a given space. Thus the entire form of a ptycholophe can be interpreted as a means of maintaining the maximum effective filtering capacity within a shell of limited size. Any ptycholophe is inherently limited by being attached throughout to the mantle, so that its elaboration to provide greater filtering capacity can only be in two dimensions, making the best use of the area of the dorsal valve surface (cf. Rudwick 1962, p. 611). The only means of lengthening an attached filament-row within such a limited area is to buckle it into lobes and indentations. But if there are also definite optimal widths for these lobes and indentations, the lophophore must be elaborated during growth by the sequential addition of new lobes and indentations to an initially simple schizolophe. Hence it is to be expected that the degree of elaboration of the ptycholophe should be directly related to the absolute size of the dorsal valve, both during the ontogeny of a single species and also among species of different adult sizes. It should thus show ‘size- required allometry’ (Gould 1966). This seems to be true of Bactrynium as far as the limited number of specimens allows such a conclusion (text-fig. 5); and a comparison with some thecideaceans (text-fig. 11) shows that they too have lobes and indentations of about the same size. Unfortunately, quantification of this allometric relation is hardly possible, in view of the difficulties of estimating the size of the true ‘body’ of a brachiopod, as opposed to the size of its mantle and shell. AFFINITIES AND EVOLUTION Affinities o/ Thecospira. Williams (1953n) pointed out that Thecospira had such strik- ingly strophomenide characters that it could not be excluded from that group solely because of its possession of a spiral brachidium, the only character which suggested affinity with the spiriferides. Thecospira was therefore included within the Orthotetacea (now Davidsoniacea). Like the davidsoniaceans of the Palaeozoic, Thecospira was cemented by the ventral valve during the earlier part of its life history; it has a strophic hinge-line, well-developed 350 PALAEONTOLOGY, VOLUME 11 teeth and sockets, and a large rectangular cardinal process. Most Palaeozoic davidsonia- ceans (except the earliest) are pseudopunctate, but Schuchertella is said to be impunctate (Williams 1965, p. 408), and an apparently punctate structure has been reported in Streptorhynchus (Thomas 1958). Against this background of diverse shell structure, the existence of punctate, impunctate, and obscurely pseudopunctate species of Thecospira is consistent with an argument for affinity. text-fig. 11. Drawings of the grooves and ridges on the dorsal valve of various species of Thecidiopsis, to show the uniformity of dimensions of the lobes. All x5; compare with Bactrynium, text-fig. 5. (Sources: a, Nekvasilova 1964, pi. XI, fig. 4, 6, 8; pi. XII, fig. 5; b-f, Backhaus 1959, b, pi. 5, fig. 4; c, pi. 4, fig. 2; d, pi. 7, fig. 1,2; e, pi. 4, fig. 8; f, pi. 5, fig. 7.) Lobes of e and F marked with inferred order of development ; compare text-fig. 6c. Scale represents 5 mm. Thecospira is exceptional among davidsoniaceans in having an ‘entire’ (i.e. plane) pseudodeltidium (Williams 1965, p. H366), but this is a development that certainly occurred several times among the strophomenides, and apparently related shells have a convex pseudodeltidium. Likewise it has lost all but a faint trace of the costellate shell surface characteristic of davidsoniaceans, but this too occurred in other strophomenide groups (Williams 1965, p. H364). When the costellae are most clearly visible they show a pattern of increase by intercalation, as in other davidsoniaceans, and not by lateral branching. The pustular or spinule-bearing shell surface which apparently replaced the RUDWICK: TRIASSIC BRACHIOPODS THECOSPIRA AND BACTRYNIUM 351 costellae is not found, so far as I am aware, in any other davidsoniacean, and is more reminiscent of productaceans such as Waagenoconcha. But Thecospira lacks the larger tubular spines that seem to have been universal among the productaceans and stropha- losiaceans. The most important character distinguishing Thecospira from other davidsoniaceans is, of course, the spiral brachidium. But the unique structure of the lamellae suggests that the brachidium developed independently from those of the Spiriferida. Spiral brachidia have recently been reported in Cadomella (Cowen and Rudwick 1966), which is generally regarded as related to the chonetaceans, so that there is now evidence that spiral brachidia evolved at least twice outside the Spiriferida. Even among the Spiriferida themselves spiral brachidia may have evolved more than once, the atrypaceans being in many respects closer to the rhynchonellides than to most other spiriferides (Copper 1965). Palaeozoic davidsoniaceans probably possessed spirolophes supported by a hydro- static or other ‘soft’ skeleton, for spiral impressions are known in Davidsonia itself. Davidsoniaceans existed in considerable abundance and variety even in Upper Permian time. The teeth of Thecospira , being unsupported by dental lamellae, suggest the Schuchertellidae or Orthotetidae as possible ancestors. The cardinal process and ‘socket plates’ have perhaps the greatest resemblance to those of Streptorhynchus ; Orthotetes is another possible ancestor. From some such Permian ancestor, the evolution of Thecospira would have involved (a) acquisition of brachidium ( b ) loss of convexity in pseudodeltidium and chilidium (c) obsolescence of costellae and (in some species) replacement by pustules, and possibly ( d ) loss of pseudopunctate structure and (in some species) acquisition of punctae. There is, however, no reason to suppose that these changes took place simultaneously. As already mentioned, the diversity of shell structure in Thecospira was foreshadowed in the Permian davidsoniaceans. Even the spiral brachidium may have evolved earlier than is yet apparent, and it might be worth examining Permian davidsoniaceans for any trace of the bases of crura. In any case, the evolution of Thecospira would not have involved any major changes in the relation of the shell to the substratum, in the hinge and musculature, or in the ciliary feeding mechanism. Only the supporting structures of the brachial axis would have changed, from a probably hydrostatic to a purely calcareous skeleton. As already suggested, there may have been an adaptive advantage in this change, in allowing a greater filtering capacity within a shell of given size. The Koninckinidae, which have generally been regarded as abnormal Spiriferida, should probably now be placed with Cadomella as post-Palaeozoic descendants of the chonetaceans (Cowen and Rudwick 1966). Like Thecospira , the koninckinids are first known from Middle Triassic strata, but they are significantly different in the structure of the brachidium and many other characters. If this interpretation of koninckinids is correct, it implies that two relict strophomenide groups evolved spiral brachidia inde- pendently at about the same time, after most of the related ‘normal’ genera had become extinct. Affinities of Bactrynium. As already mentioned, the affinities of Bactrynium have re- mained more controversial than those of Thecospira. Affinity to Lyttoniacea and to Thecideacea are interpretations that have been almost equally supported for many 352 PALAEONTOLOGY, VOLUME 11 text-fig. 12. Diagram to show inferred phyletic and functional relationships of Bactrynium and T/ieco- spira to other brachiopods. Outlines of dorsal valves of respresentative species are shown at uniform PERMIAN TRIASSIC JURASSIC CRETACEOUS TERTIARY RUDWICK: TRIASSIC BRACHIOPODS THECOSPIRA AND BACTRYNIUM 353 decades. The form of the lobate apparatus of Bactrynium has an immediate resemblance to that of the lyttoniaceans. The strongly concave form of the dorsal valve gives it an especially close resemblance to Oldhamina. It is necessary, however, to examine these similarities critically, in order to assess their value as clues to affinity. In a lyttoniacean the highly modified dorsal valve (‘internal plate’ of Williams 19536, 1965) has a pair of sub-parallel submedian primary lobes, from which secondary lobes branch laterally. Between the lobes are lateral indentations and a single median indenta- tion. The degree of symmetry is variable, but in Oldhamina and Leptodus, for example, it is as great as in Bactrynium. In such genera as these, there is also an obvious approach to uniformity in the width of the lobes and of the indentations. The mode of growth of the lobes can be inferred, by analogous reasoning, to have been similar to that of Bactrynium, with the secondary lobes extending laterally, while each primary lobe extended forwards and periodically ‘budded off’ a new secondary lobe. This is con- firmed by the evidence of the growth-lines on the lobes. The similarities even extend to the detailed morphology, for the inner side of the dorsal valve of lyttoniaceans has a contiguous ridge and groove (the ridge being external to the groove) running around the edge of the lobes and indentations, while the area enclosed by the groove is often pustu- lar in appearance (PI. 68, fig. 10; see also Stehli 1956, pi. 41, fig. 3; Williams 1965, fig. 393, 3d). But the differences between the lobate apparatus of Bactrynium and the dorsal valve of a lyttoniacean are as striking as the resemblances. The strophic hinge and articulation of Bactrynium are apparently quite normal, and are certainly unlike the highly aberrant hinge structures of the lyttoniaceans. The dorsal valve of Bactrynium is a thick massive plate, whereas that of a lyttoniacean is excep- tionally thin and delicate. More significantly, the dorsal valve of Bactrynium has a normal ‘entire’ outline, whereas that of a lyttoniacean is highly indented and corre- sponds to the course of the lobed ridge. It is true that its indented outline is often modified by ‘bridges’ of shell material across the indentations, either occasional as in Oldhamina (cf. Williams 1965, fig. 293, 2b), confined to the median indentation as in Gubleria (Termier and Termier 1960), or regularly across all the indentations as in Coscinophora (cf. Williams 1965, fig. 393, la, lc). The inner ends of the posterior magnification (x2), with course of brachial axes (reconstructed from brachial grooves or brachidial lamellae) marked in thicker black lines (hatched where reconstruction is tentative). Note fairly uni- form dimensions of lobes of Thecideacea (1-17) and spiralia of Thecospira (18) in contrast to much larger lobes of Lyttoniacea (20). Variations in adult body size are indicated approximately by positions on horizontal scale, smallest species being roughly central. (If space allowed, the lyttoniacean (20) would be much further to right.) Stratigraphical horizons of species indicated approximately against Geological Society time-scale (.Quart. J. geol. Soc. Loud. 120S, 260-2). Broad outline of suggested phylogeny shown by functional ‘zones’ based on type of lophophore in adult stage: specific phyletic pathways are not given. For further explanation see text. Key to genera: thecideacea: 1. Lacazella ; 2. Glazewskia; 3-5. Thecidellina\ 6. Parathecidea; 1-9. Thecidiopsis', 10. Bifolium ; 11. Glazewskia', 12. Eudesel/a; 13. Elliottina; 14. Davidsonella : 15. Moorellina; 16. Cooperina ; 17. Bactrynium ; david- soniacea: 18. Thecospira ; 19. A Permian davidsoniacean (based on Diplanus), with (a) juvenile stage with inferred schizolophe, and (b) adult with inferred spirolophe. lyttoniacea: 20. Leptodus (Frag- ment only). (Sources: 2, Pajaud 1965; 4, Thomson 1915; 5, Toulmin 1940; 6-9, Backhaus 1959; 10, Nekvasilova 1964; 11, Glazewski and Pajaud 1965; 16, Termier, Termier and Pajaud 1966; 1, 3, 12-15, 17-20 original.) 354 PALAEONTOLOGY, VOLUME 11 indentations may also be filled up progressively, as in Leptodus (cf. Williams 1965, fig. 393, 3d). But in no lyttoniacean were the indentations completely filled in: even in Coscinophora gaps of uniform length were maintained between the regularly spaced ‘bridges’ across the indentations. As Sarycheva (1964) has pointed out, Stehli’s (1956) suggested derivation of Bactrynium from Coscinophora is therefore improbable. The indented edge of the dorsal valve of lyttoniaceans rests on a series of septa rising from the floor of the ventral valve. The interior of the ventral valve of Bactrynium is not well known, but it certainly has no such septa rising to make contact with the lobate apparatus on the dorsal valve (text-fig. 2b; PI. 68, fig. 9). Another difference of great importance is that of absolute size. The lobes and in- dentations on a lyttoniacean are nearly three times as large (in linear dimensions) as those of Bactrynium (text-fig. 12; PI. 68, compare fig. 10 with fig. 1). This difference is independent of the absolute size of the whole shell, since in both groups the width of the lobes and the width of the indentations were clearly kept more or less constant during growth. As Gould (1966, p. 604) suggests, such a difference should draw attention to a probable adaptive discontinuity between the groups concerned. The morphology of lyttoniaceans is so abnormal that any reconstruction of their feeding mechanism on grounds of analogy alone would be hazardous. Only a full functional analysis of their morphology will give rational grounds for any such recon- struction. In the meantime, it should not be assumed that their feeding mechanism was necessarily the same as that of Bactrynium. In particular the very thin and delicate dorsal valve, recessed within a more robust ventral valve, and the persistently main- tained slots or perforations between the lobes of the dorsal valve, suggest a significantly different functional organization; and the much larger dimensions of the lobes and indentations point to the same conclusion. On the other hand it is probable that the groove running round the lobes and indenta- tions bore a brachial axis as in Bactrynium , and that the lophophore therefore was a ptycholophe. If this is correct, the same topological considerations could have been responsible for the very similar arrangement and mode of growth of the lobes. The over-all similarity could, therefore, be due to simple functional parallelism. There is now no good evidence for the survival of true lyttoniaceans after the end of the Permian period. The large stratigraphical gap between the last undoubted lyttonia- ceans in the Dzjulfian (late Permian), and Bactrynium in the Rhaetian (late Triassic), is therefore no longer a problem. A much closer comparison can be made between Bactrynium and the thecideaceans. In both, the dorsal valve is a massive plate with a thickened sub-marginal rim and ‘entire’ outline. The ventral valve is cemented to the substratum, at least in earlier growth-stages. The rectangular cardinal process of Bactrynium , with flanking sockets and socket plates, is closely similar to the cardinalia of thecideaceans. The shell surface of thecideaceans has no radial ornament. The hinge is strophic, with a low dorsal inter- area and higher ventral interarea. Many other characters of Bactrynium can be matched among the earlier thecideaceans of the Lower Jurassic. Thus Davidsonel/a has already been reported as pseudopunctate (Elliott 1965), and there is a similar obscurely pseudopunctate structure in Moorel/ina leptaenoides (Deslongchamps). The cardinal process and sockets are similar to those of Bactrynium, and lateral to RUDWICK: TRIASSIC BRACHIOPODS THECOSPIRA AND BACTRYNIUM 355 the cardinal process is a pair of elliptical areas similar to the postulated lateral adductor scars of Bactrynium. The ventral valve is generally strongly convex, and the dorsal valve may be concave as in Bactrynium (e.g. Davidsonella sinuata (Deslongchamps)). There is a similar lobed ridge and groove in Eudesella (PL 68, fig. 3). The area within the groove is generally pustular (PI. 68, figs. 3, 7, 8). In Davidsonella pustules are highly developed (PI. 68, figs. 4, 5), but do not appear to be separate spicules as implied in the Treatise description (Elliott 1965, p. H859). Posteriorly the ridge projects across a postero-median cavity, often as a complete ‘bridge’, which is quite distinct from the ‘socket plates’ (PI. 68, fig. 4). The median indentation of the ridge may project pos- teriorly as a distinctly doubled spur or spike closely resembling that of Bactrynium (PI. 68, figs. 4-6). In thecideaceans there is, as already mentioned, a clear approximation to a uniform width for the lobes and for the indentations, both in the different growth stages of a species and between different species; and these uniform dimensions are closely com- parable to those of Bactrynium (compare text-fig. 1 1 with text-fig. 5; also PI. 68, compare figs. 3-5, 7, 8 with fig. 1). This suggests a common feeding mechanism for Bactrynium and thecideaceans, with ‘tunnels’ of exhalent water stabilized at the optimum dimen- sions inherent in any normal (i.e. non-lyttoniacean) ptycholophous system. The chief difference between the lobes of Bactrynium and those of the thecideaceans lies in their arrangement. In genera with several lobes the lobes project anteriorly or antero-medially, and never laterally (text-fig. 11). It can, however, be inferred that they grew in the same manner as that postulated for Bactrynium. The difference is that the primary lobe of thecideaceans is that furthest from the mid-line, not that nearest the mid-line. The primary lobe extends parallel to the postero-lateral sector of the valve edge ; the secondary lobes project anteriorly or antero-medially from it, and must have been ‘budded’ from it serially, the postero-median lobes being the oldest (text-figs. 6c-d). Expressed another way, the secondary lobes were budded from the medial side of the primary lobe in thecideaceans, but from the lateral side in Bactrynium. Beginning from a schizolophous stage, these are, in fact, the only two distinct ways in which the number of lobes can be increased (without involving an unlimited amount of shift in the absolute positions of all the lobes). As already shown, it appears that Bactrynium actualized one of these alternative possibilities, with secondary lobes being formed serially on the lateral side of each primary lobe; whereas the thecideaceans actualized the other alternative, with secondary lobes being formed serially on the medial side of each primary lobe (in some thecideaceans the budding pattern may have been irregular — see text-figs. 11a-c). These alternatives, though topologically equivalent, are probably not equal in func- tional efficiency. On the first alternative ( Bactrynium ) all the secondary lobes project laterally, and hence all the ‘tunnels’ (except the median one) opened laterally. On the second alternative (Thecideacea) all the secondary lobes project antero-medially, anteriorly, or at most antero-laterally, so that all the ‘tunnels’ open around the anterior part of the gape. In terms of the current systems, the greatest outflow of exhalent water would be, on the first alternative, in the lateral part of the gape, and on the second, in the anterior part. But the area available to act as exhalant aperture is limited laterally by the degree of separation of the valve edges, whereas anteriorly the much wider total gape can be divided between inhalant and exhalant apertures simply by an appropriate 356 PALAEONTOLOGY, VOLUME 11 orientation of the apertural filaments. In other words it would seem simpler to ensure an unimpeded outflow of exhalant water on the second alternative than on the first. This suggests a possible adaptive advantage in favour of the thecideaceans, which may be reflected in their longer range and greater diversity and abundance. It is interesting that within the lyttoniaceans there are genera (e.g. Paralyttonia, Rigbyella ) with forwardly projecting lobes like those of the thecideaceans, as well as the better-known genera (e.g. Oldhamina, Leptodas ) with laterally projecting lobes like those of Bactrynium. This fact was used by Wanner (1935) as an argument for affinity between lyttoniaceans and thecideaceans (including Bactrynium). He maintained that any other interpretation would involve postulating a double convergence. But the present analysis of the growth of ptycholophes implies on the contrary that the parallel development of both alternative arrangements of the lobes is by no means improbable. These similarities together make a strong case for affinity between Bactrynium and Thecideacea. Indeed, in my opinion there is now no difference warranting supra- familial recognition. I therefore suggest that Bactrynium, while retaining its family Bactryniidae in recognition of the distinctive form of its lobate apparatus, should be assigned to the Thecideacea. In phylogenetic terms Bactrynium can best be regarded as a derivative of small and simple schizolophous thecideacean ancestors. Such species are known in the Permian ( Cooperina ) and Rhaetian (e.g. Moorel/ina), and are probably represented in inter- mediate strata among the poorly known Triassic ''Thecidea' spp. From some such ancestor, Bactrynium could have evolved allometrically by elaboration of the schizo- lophe into a ptycholophe and by concurrent increase in absolute size. Later ptycholo- phous forms, such as Eudesella in the Lower Jurassic, utilized the other alternative arrangement of lobes, and therefore probably evolved independently from schizolophous forms. Bactrynium might, however, have left some schizolophous descendants by neo- teny; the pseudopunctate Davidsonella, for example, might have had such an origin. Without attempting to reconstruct the detailed course of phylogeny in the Thecideacea, the functional analysis given here leads to an interpretation in terms of ‘functional zones’ (text-fig. 12). Each ‘zone’ represents the utilization of one possible mode of organiza- tion of the lophophore, without any necessary effect on the mode of life of the whole organism; functional zones are thus not synonymous with adaptive zones. Using this concept, it would seem that from the earliest thecideacean ( Cooperina ) onwards, there has always been a ‘zone of schizolophes’, occupied by species small enough for a schizolophe to be an adequate form for the lophophore. At certain times there have also been species which, by utilizing one or other ptycholophous arrangement, were able to increase in absolute size and evolve into ‘zones of ptycholophes’. Gould (1966) has already cited some thecideaceans as an example of such ‘size-required allometry’. Such species would, of course, have traversed the ‘zone of schizolophes’ during ontogeny, as, for example, Lacazella is known to do at the present day. I have made a somewhat arbitrary distinction between ‘zones of simple ptycholophes’, containing species with no more than three pairs of lobes, and ‘zones of complex ptycholophes’, containing species with larger numbers of lobes. The ‘envelopes’ outlining the zones on text-fig. 12 are merely to show the limits of known species in each zone: they do not imply that any zone was in any sense unavailable at other times, nor that all the specimens in a given zone are closely related to each other (e.g. Eudesella in the Lower Jurassic may have RUDWICK: TRIASSIC BRACHIOPODS THECOSPIRA AND BACTRYNIUM 357 evolved from schizolophous ancestors independently from the much later, Cretaceous, ptycholophous species). Affinities ofThecideacea. Further discussion of the affinities of Bactrynium thus involves the question of thecideacean affinities. Lacazeda and its fossil relatives formed an im- portant element in the earlier concept of the ‘Protremata’, and underlying this was the belief that the Thecideacea were related to the brachiopods now grouped together as Strophomenida. More recently, however, this early view has been regarded as doubtful, and an affinity with the Terebratulida or Spiriferida has been favoured (Williams and Rowell 1965). In this interpretation undue weight has perhaps been given to the fact that most Thecideacea are punctate. But it is now recognized that the punctate structure must have evolved independently several times during the history of the Brachiopoda (Williams and Rowell 1965, p. H68; Wright 1966). Therefore the punctation ofThecideacea is not by itself a reliable criterion of affinity to the Terebratulida or punctate Spiriferida. The existence of diverse shell structures among the earlier Thecideacea increases the likelihood that punctation has evolved independently in this group. Two other important characters of Thecideacea, their cementation and their lobed brachial grooves, find no parallel among Spiriferida or Terebratulida. There is now no authenticated case of cementation attachment in any articulate brachiopod outside the Strophomenida; for Dagis (1965) has justly thrown doubt on Thecocyrtella, and Bittnerula is now known to have a pedicle foramen (Cowen and Rudwick 1967). The cemented attachment of Thecideacea is, therefore, strong evidence against any but a strophomenide affinity. Linked with the possession of cementation is the fact that Thecideacea have a pseudodeltidium and no pedicle foramen; and their cardinalia cannot be matched closely among terebratulides or spiriferides. The brachial supporting structures of Thecideacea, with the brachial axes in grooves excavated in the dorsal valve surface, are entirely different from any of the varied spiral and looped brachidia of Spiriferida and Terebratulida, all of which are composed of calcareous lamellae growing independently from the dorsal valve floor, though attached to the cardinalia and also (in some forms) to septa. A far stronger case can be made for affinity between the Thecideacea and the Stropho- menida. Of strophomenide brachiopods, the Davidsoniacea show the greatest resem- blances to the Thecideacea. Both groups have cementation attachment and therefore no pedicle foramen, and both lack the tubular spines of chonetoids and productoids. Both have strophic hinges, with strong articulation flanking a fairly massive cardinal process which is generally covered by a convex pseudodeltidium. This conclusion would be given taxonomic recognition by the assignment of the Theci- deacea to the Strophomenida. Their possession of brachial grooves and related struc- tures, implying schizolophous or ptycholophous but never spirolophous lophophores, their lack of costellate ornament, and their generally (but not universally) punctate shell structure are collectively sufficient grounds for retaining them as a superfamily distinct from the Davidsoniacea. Until recently the Thecideacea were unknown before the Rhaetian (Elliott 1965); but with the discovery of the simple but true thecideacean Cooperina in the Word Formation (Upper Permian) of Texas (Termier, Termier, and Pajaud 1966), it is now 358 PALAEONTOLOGY, VOLUME 11 clear that the ancestors of the group must be sought among Lower Permian or even earlier davidsoniaceans. These, like the Devonian Davidsonia itself, probably had spirolophous lophophores supported only by a hydrostatic skeleton or by muscular and connective tissue. By analogy with living brachiopods, most Permian davidsonia- ceans were almost certainly too large for a schizolophe to have been adequate for their needs; but like all living spirolophous species they would have passed through a schizo- lophous stage early in ontogeny while they were still small in size. In the light of these functional considerations a neotenous origin for the Thecideacea seems most probable. Only the development of a sessile schizolophe, accommodated in a groove in the dorsal valve, would have been required to convert a small and young davidsoniacean into a true thecideacean (text-fig. 12). CONCLUSIONS If the foregoing arguments are correct, the great Palaeozoic order of Strophomenida, which suffered a drastic degree of extinction at about the end of the Permian period, survived into the Mesozoic only as three groups of small, rare, and inconspicuous brachiopods. One group was the Koninckinidae, modified descendants of the Chone- tacea (Cowen and Rudwick 1966). The second group consisted of small and simple davidsoniaceans together with Tliecospiro. The spiral brachidium of the latter genus may have evolved at about the same time (early or middle Triassic) as the parallel develop- ment in the Koninckinidae, both being independent of the spiral brachidium of the Spiriferida. The brachidium of Thecospira may have increased the efficiency of the pumping action of its lophophore, but did not entail any essential change in its ciliary feeding mechanism, or in the exhalant spirolophous current system by which that mechanism operated. Apart from its spiral brachidium Thecospira was functionally similar to more ‘normal’ davidsoniaceans. The third group, the Thecideacea, was already in existence even before the end of the Permian, and was probably derived by neoteny from the davidsoniaceans. In late Triassic time Bactrynium evolved from some small and simple thecideacean by elaboration of a schizolophe into a ptycholophe and by corresponding increase in size. This ‘size-required allometry’ involved no essential change in its ciliary feeding mechanism, but its concavo-convex shell form adapted it to a free-lying mode of life. Bactrynium itself became extinct at the end of the Triassic, though some neotenous descendants may be represented among the varied Thecideacea of the Lower Jurassic. These also included a new and independent development of ptycholophous forms ( Eudesella ). The Thecideacea survived thereafter, though incon- spicuously, up to the present day. On this interpretation the Strophomenida are not yet extinct. REFERENCES atkins, d. 1960. The ciliary feeding mechanism of the Megathyridae (Brachiopoda), and the growth stages of the lophophore. /. mar. biol. Ass. U.K. 39, 459-79. 1963. Notes on the lophophore and gut of the brachiopod Notosaria (formerly Tegulorhynchid) nigricans (G. B. Sowerby). Proc. zool. Soc. Lond. 140, 15-24. and rudwick, m. j. s. 1962. The lophophore and ciliary feeding mechanism of the brachiopod Crania anomala (Muller). J. mar. biol. Ass. U.K. 42, 469-80. backhaus, e. 1959. Monographic der cretacischen Thecideidae (Brach.), Mitt. geol. St Inst. Hamb. 28, 5-90. RUDWICK: TRIASSIC BRACHIOPODS THECOSP1RA AND BACTRYNIUM 359 bittner, A. 1890. Brachiopoden der alpinen Trias. Abh. geol. Bundesanst., Wien, 14. broili, f. 1916. Die permischen Brachiopoden von Timor. In wanner, j. (ed.), Palciontologie von Timor, 7, 12. copper, p. 1965. Unusual structures in Devonian Atrypidae from England. Palaeontology, 8, 358-73. cowen, R., and rudwick, m. j. s. 1966. A spiral brachidium in the Jurassic chonetoid brachiopod Cadomella. Geol. Mag. 103, 403-6. 1967. Bittnerula Hall and Clarke, and the evolution of cementation in the Brachiopoda. Geol. Mag. 104, 155-9. dagis, a. s. 1965. Triassic brachiopods of Siberia. Nauka, Moscow, 1-186. (In Russian.) elliott, g. f. 1965. Suborder Thecideidina. In moore, r. c. (ed.). Treatise on Invertebrate Paleontology, part (H) Brachiopoda, H857-62. emmrich, h. 1855. Notiz liber den Alpenkalk der Lienzer Gegend. Jb. geol. Bundesanst., Wien, 6, 444-50. glazewskj, k., and pajaud, d. 1965. Sur une nouvelle espece de Thecideidae, Glazewskia sp. (Brachio- pode) du Jurassique de Podolie. Bull. Soc. geol. Fr. (7) 6, 262-8. gould, s. j. 1966. Allometry and size in ontogeny and phylogeny. Biol. Rev. 41, 587-640. grant, r. e. 1966. Spine arrangement and life habits of the productoid brachiopod Waagenoconcha. J. Paleont. 40, 1063-9. kozlowski, R. 1929. Les brachiopodes gothlandiens de la Podolie polonaise. Palaeont. pol. 1, 1-254. LACAZE-DUTfflERS, h. de. 1861. Histoire naturelle des Brachiopodes vivants de la Mediterranee. Annls Sci. nat. (4), 15, 259-330. makridin, v. p. 1960. Superfamily Thecideacea. In sarycheva, t. g. (ed.). Os navy Paleontologii, Mshanki, Brakhiopody, 1-305. (In Russian.) nekvasilova, o. 1964. Thecideidae (Brachiopoda) der Bohmischen Kreide. Sb. geol. Ved. P3, 1 19-62. noetling, f. 1905. Untersuchungen liber die Familie der Lyttoniidae Waagen emend. Noetling. Palaeontographica, 51, 129-53. paine, R. t. 1962. Filter-feeding pattern and local distribution of the brachiopod Discinisca strigata. Biol. Bull. mar. biol. Lab., Woods Hole, 123, 597-604. pajaud, d. 1965. Remarques sur les Thecideidae (Brachiopodes) tertiaires. Bull. Soc. geol. Fr. (7) 6, 258-61. rudwick, m. j. s. 1959. The growth and form of brachiopod shells. Geol. Mag. 96, 1-24. 1960. The feeding mechanism of spire-bearing fossil brachiopods. Geol. Mag. 97, 369-83. 1962. Filter-feeding mechanisms in some brachiopods from New Zealand. J. Linn. Soc. Lond. ( Zool .) 44, 592-615. ■ 1964. The inference of function from structure in fossils. Brit. J. Phil. Sci. 15, 27-40. 1965. Ecology and paleoecology. In moore, r. c. (ed.). Treatise on Invertebrate Paleontology, part (H) Brachiopoda, H199-H214. sarycheva, t. g. 1964. Oldhaminoid brachiopods from the Permian of Transcaucasia. Paleont. Zh. 1964, 3, 58-72. (In Russian.) stehli, f. g. 1956. Notes on oldhaminid brachiopods. J. Paleont. 30, 305-13. suess, e. 1854. Uber die Brachiopoden der Kossener Schichten. Denkschr. Akad. Wiss., Wien, 7, 29-65. termier, h., and termier, g. 1960. Fes Oldhaminides du Cambodge. Bull. Soc. geol. Fr. (7) 1, 233-44. and pajaud, d. 1966. Decouverte d’une Thecidee dans le Permien du Texas. C.R. Acad. Sci. Paris, 263, D332-5. Thomas, g. a. 1958. The Permian Orthotetacea of Western Australia. Bull. Bur. Miner. Resour. Geol. Geophys. Aust. 39. Thomson, j. a. 1915. On a new genus and species of the Thecidiinae (Brachiopoda). Geol. Mag. (6), 461-4. toulmin, l. d. 1940. Eocene brachiopods from the Salt Mountain limestone of Alabama. J. Paleont. 14, 227-33. waagen, w. 1879-87. Productus Fimestone Fossils. Palaeont. indica (13), 1. wanner, j. 1935. Fyttoniidae and ihre biologische und stammesgeschichtliche Bedeutung. Neues Jb. Miner. Geol. Paldont. BeilBd. 74B, 201-81. williams, a. 1953u. The classification of the strophomenoid brachiopods. J. Wash. Acad. Sci. 43, 1-13. 19536. The morphology and classification of the oldhaminid brachiopods. Ibid. 43, 279-87. 360 PALAEONTOLOGY, VOLUME 11 williams, a. 1956. The calcareous shell of the Brachiopoda, and its importance to their classification. Biol. Rev. 31, 243-87. 1965. Strophomenidina, Oldhaminidina. In moore, r. c. (ed.), Treatise on Invertebrate Paleonto- logy, part (H) Brachiopoda, H362-H412, H510-21. and wright, a. d. 1961. The origin of the loop in articulate brachiopods. Palaeontology , 4, 149— 176. and rowell, a. j. 1965. Anatomy, morphology, evolution and phylogeny. In moore, r. c. (ed.), Treatise on Invertebrate Paleontology, part (H) Brachiopoda, H6-H57, H57-H138, HI 64-99. wright, a. d. 1966. The shell punctation of Dicoelosia biloba (Linnaeus). Geol. For. Stockh. Fork. 87, 548-56. zugmayer, h. 1880. Untersuchungen fiber rhatische Brachiopoden. Beitr. Palaont. Geol. Ost.-Ung. 1, 1-42. M. J. S. RUDWICK Department of Geology Sedgwick Museum Typescript received from author 28 April 1967 Cambridge UPPER CRETACEOUS COCCOLITHOPHORIDS FROM ZULULAND, SOUTH AFRICA by RICHARD N. PIENAAR Abstract. A detailed study of some Upper Cretaceous calcareous nannoplankton of Zululand was undertaken, utilizing the optical as well as the electron microscope. Nine new species belonging to five genera are described as viewed in the electron microscope. These are Coccolithus cribosphaerella, Coccolithus zuluensium, Cyclolithus zulua, Discolilhus cristallinus, Discolilltus rhabdosphaericus, Discolithus spiralis, Maslovella africana, Maslovella blackii, and Maslovella pulchra. Using the sequential occurrence of the coccoliths through the top 800 ft. of 'Zululand Oil Exploration’ Borehole 'A’ the age was determined as being Cretaceous Maestrichtian. The material on which this study was based was supplied by the Anglo Transvaal Consolidated Investment Company of South Africa. It belongs to part of this company's oil prospecting project and is from Borehole ‘A’ drilled near Lake Sibaya, Zululand, South Africa. The method used to extract and prepare the coccoliths for observation under both the optical as well as the electron microscope are similar to those described by Pienaar (1966). All samples are housed in the Bernard Price Institute for Palaeontological Research, University of the Witwatersrand, while the prepared slides and electron micrographs are housed in the Department of Plant Biology, University of Natal, Durban, South Africa. TAXONOMIC PROBLEMS It has been shown by numerous workers in this field that the systematics of this group of algae is in a state of turmoil. This has arisen mainly out of the fact that there have been many approaches to this field of study. The use of phase contrast, polarizing, and ordinary optical microscope as well as the electron microscope have resulted in con- fusion within the group because these algal remains appear so completely different when viewed under the different microscopes. It has been shown that the only satisfactory way to study this group systematically, is by utilizing the electron microscope which with its high power of resolution reveals the detailed structure of the coccoliths. This is vitally important as the systematics of the Coccolithophoridae is based on hard-part morphology; I do not suggest that the optical microscope be abandoned, and have in fact stressed (Pienaar 1966) that it is invaluable in the stratigraphic application of the group. In the present report no attempt has been made to place the coccoliths under any hierarchical systematic framework. I consider it premature to do so until there is a good collection of electron micrographs available, and for this reason the coccoliths are described in alphabetical order. [Palaeontology, Vol. 11, Part 3, 1968, pp. 361-7, pis. 69-71.] C 5586 B b 362 PALAEONTOLOGY, VOLUME 11 DESCRIPTIVE TERMINOLOGY Distal shield. Shield which is the furthest away from the point of attachment of the coccolith to the coccosphere, and is always convex. Longitudinal axis. Line joining the two longitudinal poles. Longitudinal poles. Situated at each end of the longitudinal axis of the theoretical ellipse that contains the coccolith. Plates. Elements which comprise the proximal and distal shields; they vary from being wedge-shaped to rectangular in outline. Pore. Region where there are no crystals of calcium carbonate, usually within the central area of the coccolith. Proximal shield. Part of the coccolith which is in direct contact with the coccosphere; always concave. Shield. Main structural element of a coccolith, always composed of plates. Shield area. Width of the shields when the coccolith is seen in plan view. Transverse axis. Line joining the two transverse poles. Transverse poles. Situated at each end of the transverse axis of the theoretical ellipse that can contain the coccolith. SYSTEMATIC DESCRIPTIONS Genus coccolithus Schwarz 1894 Coccolithus cribosphaerella sp. nov. Holotype. Plate 70, figs. 4, 5. Locus typicus. Borehole ‘A’, Lake Sibaya, Zululand, South Africa. Assemblage 1382, depth 280 ft.; Cretaceous. Diagnosis. Elliptical placoliths composed of two well-developed shields; distal shield larger, convex, composed of 21-22 plates. The exact detail of the smaller proximal shield is not known. The central area is sculptured with numerous pores aligned parallel to the longitudinal axis of the ellipse. Description. The scalloped outline is due to the overlapping plates of the distal shield. The proximal shield is presumed to be similarly composed. The central area is sculptured EXPLANATION TO PLATE 69 Figs. 1, 5. Maslovelta btackii sp. nov. non-replicated. 1, X 28,000. 5, X 10,700. Fig. 2. Discolithus cristallinus sp. nov., non-replicated; X 21,660. Fig. 3. Maslovelta pulchra sp. nov., non-replicated; X 27,000. Fig. 4. Cyclolithus zulua sp. nov. non-replicated; X 13,330. Figs. 6, 7. Coccolithus zuluensium sp. nov., non-replicated. 6, X 13,330. 7, X 13,750. Fig. 8. Maslovelta africana sp. nov., non-replicated; X 12,850. Fig. 9. Discolithus rhabdosphericus sp. nov., non-replicated; X 13,125. Palaeontology, Vol. 11 PLATE 69 PIENAAR, Cretaceous coccolithophorids PIENAAR: UPPER CRETACEOUS COCCOLITHOPHORIDS FROM ZULULAND 363 with 3-5 rows of pores aligned parallel to the longitudinal axis of the ellipse; the pores vary from circular to hexagonal in shape. Size. Longitudinal axis, 4-5-6-0 p. Transverse axis, 3-1-4-8 p. Width of shield area, 0-75-10 p. Remarks. This coccolith is well represented in the assemblages studied. The pore size is variable. Stradner (1961) described Coecolithus opacus which resembles the South African species and could well be a related form. Coecolithus zuluensium sp. nov. Holotype. Plate 69, figs. 6, 7. Locus typicus. Borehole ‘A’, Lake Sibaya, Zululand, South Africa. Assemblage 1388, depth 330 ft.; Cretaceous. Diagnosis. Elliptical coccolith made up of two well-developed shields; distal shield larger. The central area is spanned by two cross-bars which widen centrally and corre- spond to the axes of the ellipse. Description. The distal shield is composed of 27-29 overlapping plates. The longitudinal polar plates are wedge-shaped and are larger than the remaining rectangular plates. The proximal shield is similar in construction to the distal shield. The central area is the characteristic feature of the species; the cross-bars are com- posed of calcium carbonate crystals arranged in an irregular manner. Size. Longitudinal axis, 3-25-4-4 p. Transverse axis, 2-3-3-7 p. Width of shield area, 0-8 1-0 p. Remarks. The coccoliths grouped together in this species resemble Disco/iihus rhabdo- sphaericus sp. nov. but differ in that they have a proximal and a distal shield and lack a central boss. The species is assigned to the genus Coecolithus because of the two- shielded nature. It has been noticed that during the course of the study of the South African Coccolithophorids that the members belonging to the genus Coecolithus usually possess large wedge-shaped polar plates. Genus cyclolithus Kamptner 1948 Cyclolithus zulua sp. nov. Holotype. Plate 69, fig. 4; Plate 70, fig. 1. Locus typicus. Borehole ‘A’, Lake Sibaya, Zululand, South Africa. Assemblage 1382, depth 280 ft-; Cretaceous. Diagnosis. Elliptical coccolith composed of a single shield abutting against a raised central rim. The shield is made up of 16-17 plates. Description. The 16-17 plates are rectangular and overlap slightly. Towards the central area is a raised rim, against which the plates abutt. The raised portion is levigate and devoid of any sculpture. Size. Longitudinal axis, 4-4-6 0 p. Transverse axis, 3-6-4T p. Width of shield area, 0-9-10 p. 364 PALAEONTOLOGY, VOLUME 11 Remarks. On first impression this specimen appears to have two shields but on closer examination the second shield is seen to be a slightly raised rim with practically no plate structure. Because of its one-shielded nature and the central area devoid of any sculp- ture, the specimens are placed in the genus Cyclolithus Kamptner 1948. Remarks. Under the Botanical Code of Nomenclature, the name Discolithus is retained for these algal fossils (cf. Loeblich and Tappan 1963). Genus discolithus Kamptner 1948 Discolithus cristallinus sp. nov. Holotype. Plate 69, fig. 2; Plate 70, fig. 2. Locus typicus. Borehole ‘A’, Lake Sibaya, Zululand, South Africa. Assemblage 1382, depth 280 ft.; Cretaceous. Diagnosis. Elliptical one-shielded coccolith composed of 24-40 overlapping plates. The central area is infilled with crystals of calcium carbonate arranged in an irregular order. Description. The average number of plates is between 28 and 30. Size. Longitudinal axis, 2-4-2-75 p. Transverse axis, 1 -4-7-0 p. Width of shield area, 0-25-0-45 p. Remarks. Occasionally specimens of Discolithus cristalinus were found with a distinct row of crystals following the outline of the central area. Their appearance was almost like the beginning of a second shield. In addition some forms have larger crystals of calcium carbonate covering the smaller crystals. This species is easily recognized by its single shield and the infilled central area. Discolithus rhabdosphaericus. sp. nov. Holotype. Plate 69, fig. 9; Plate 71, figs. 1, 2, and 6. Locus typicus. Borehole ‘A’, Lake Sibaya, Zululand, South Africa. Assemblage 1382, depth 280 ft.; Cretaceous. Diagnosis. Elliptical one-shielded coccolith composed of 35 overlapping plates. The central area is spanned by a solid parallelogram-like structure, the diagonals of which correspond with the axes of the ellipse. Description. The sides of the parallelogram are concave and each composed of six plates. The diagonals are raised and unite in the centre to form a boss which is perforated by an axial pore. The proximal surface of the coccolith is markedly concave and no central EXPLANATION TO PLATE 70 Fig. 1. Cyclolithus zulua sp. nov., replicated; X 30,000. Fig. 2. Discolithus cristallinus sp. nov., non-replicated ; X 22,000. Fig. 3. Discolithus spiralis sp. nov., replicated; X 12,000. Figs. 4, 5. Coccolithus cribosphaerella sp. nov., replicated. 4, X 14,000. 5, X 12,000. Fig. 6. Maslovella af icana sp. nov., replicated; X 17,500. Palaeontology, Vol. 11 PLATE 70 PIENAAR, Cretaceous coccolithophorids PIENAAR: UPPER CRETACEOUS COCCOLITHOPHORIDS FROM ZULULAND 365 boss is observed when the coccolith is found in this position. The distal surface is markedly convex in comparison with the proximal surface. Size. Longitudinal axis, 3-5-4-2 p. Transverse axis, 2-7-3-2 p. Width of shield area, 04-0-6 /a. Remarks. This is a very common and distinctive species and was assigned to the genus Discolithus Kamptner 1948 because of its one-shielded nature and the infilled central area. Discolithus spiralis sp. nov. Holotype. Plate 70, fig. 3; Plate 71, fig. 4. Locus typicus. Borehole ‘A’, Lake Sibaya, Zululand, South Africa. Assemblage 1387, depth 320 ft.; Cretaceous. Diagnosis. Ellipitical one-shielded coccolith with the central area infilled with crystals of calcium carbonate arranged in a sigmoid pattern. Description. Shield composed of approximately 56 overlapping plates. The central area is large and elliptical and completely infilled. Size. Longitudinal axis, 5-4-5-7 /a. Transverse axis, 3-6-3-7 p. Width of shield area, 0-3-0-4 p. Remarks. This is a very distinctive coccolith and may always be recognized by the sigmoid arrangement of the crystals infilling the central area. Genus maslovella Tappan and Loeblich 1966 Synonym. Covillea Black 1964. Maslovella africana sp. nov. Holotype. Plate 69, fig. 8; Plate 70, fig. 6; Plate 71, figs. 3, 5. Locus typicus. Borehole ‘A’, Lake Sibaya, Zululand, South Africa. Assemblage 1387, depth 320 ft.; Cretaceous. Diagnosis. Circular to subcircular asymmetrical coccolith composed of two well- developed shields. At the locality of the axial pore the central area is infilled with crystals of calcium carbonate. The distal shield is larger than the proximal shield and is placed asymmetrically on top of it. Description. The distal shield is composed of 25-29 overlapping plates, the average number of plates being 25. The plates are wedge-shaped and those situated at the longitudinal polar regions are larger and more markedly wedge-shaped than the re- maining plates. The proximal shield is smaller than the distal shield and composed of 25-29 plates of variable size. In the region of the one longitudinal pole are wedge-shaped plates which are only a little smaller than the distal shield plates. At the opposite longi- tudinal pole all the plates are smaller and less than half the size of the distal shield plates. The central area is infilled with irregularly arranged crystals of calcium carbonate. Size. Longitudinal axis, 3T-3-2 p. Transverse axis, 2-5-2-7 p. Proximal shield, 1 -9—2-6 p X 2-75-2-9 p. Distal shield, 3T-2-5 /ax4-1-3-7 p. Remarks. Maslovella africana is common in most of the assemblages studied and characterized by the asymmetrically placed shields. It is tentatively placed in the genus 366 PALAEONTOLOGY, VOLUME 11 Maslovella Tappan and Loeblich 1966 which it most closely resembles. Black (1964), however, did not mention in his description of the type specimen any asymmetry in the genus, and thus a new genus might have to be erected. This form has also been found by the author in Type Maestrichtian material sent to him by Dr. E. Martini. Maslovella blackii sp. nov. Holotype. Plate 69, figs. 1, 5. Locus typicas. Borehole ‘A’, Lake Sibaya, Zululand, South Africa. Assemblage 1382, depth 280 ft.; Cretaceous. Diagnosis. Elliptical two-shielded coccolith. The distal shield is larger than the proximal shield and is composed of 17-27 plates. The central area is completely infilled with crystals of calcium carbonate. Description. The plates situated at the longitudinal polar region are larger and wedge- shaped while the remaining plates are rectangular in shape and overlap to varying degrees. The structure of the proximal shield is not known but it is thought to be similar to that of the distal shield. The central area is large, elliptical, and completely infilled. Size. Longitudinal axis, 1-75-2-8 p. Transverse axis, 1 -5-2-0 p. Width of shield area, 0-5-0-6 p. Remarks. During the study of the South African sediments a number of coccoliths were found which were similar to the type description of the genus Maslovella (Black) Tappan and Loeblich 1966, except for a few minor differences. Some had only fourteen plates in the distal shield; others had only one well-developed shield and a robust rim in the place of the second shield. The patterning of the central area also varies by having slightly regular to irregular crystals infilling the central area or with two rows of crystals meeting along the longitudinal axis of the ellipse. These forms are thought to be intermediates or broken forms and are all grouped together into Maslovella blackii. Maslovella blackii differs from M. africana in that the latter form possesses a distinct asymmetry of the two shields, whereas the former species is symmetrical. Maslovella pulclira sp. nov. Holotype. Plate 69, fig. 3. Locus typicus. Borehole ‘A’, Lake Sibaya, Zululand, South Africa. Assemblage 1387, depth 320 ft.; Cretaceous. Diagnosis. Elliptical coccolith composed of two well-developed shields. The distal shield and the smaller proximal shield are both composed of seventeen plates. The central area is elliptical and infilled with prolongations of the proximal shield plates. EXPLANATION TO PLATE 71 Figs. 1 , 2, 6. Discolithus rhabdosphericus sp. nov., replicated. 1, Distal view, X 17,500. 2, Proximal view, x 15,000. 6, Proximal view, x 20,000. Figs. 3, 5. Maslovella africana sp. nov. replicated; 3, Proximal view, x 25,000. 5, Distal view, X 16,000. Fig. 4. Discolithus spiralis sp. nov., replicated; Proximal view, X 12,000. Palaeontology, Vol. 11 PLATE 71 PIENAAR, Cretaceous coccolithophorids PIENAAR: UPPER CRETACEOUS COCCOLITHOPHORIDS FROM ZULULAND 367 Description. The three longitudinal polar plates at each end are larger than the remaining plates and are distinctly wedge-shaped. The proximal shield plates have prolongations which dip inwards towards the central area and meet along the longitudinal axis of the ellipse. The prolongations are rectangular and are on all proximal plates except the six longitudinal polar plates. Size. Longitudinal axis, 2 0 /x. Transverse axis, L4 /x. Width of shield area, 0-5 /x. Remarks. This form is tentatively placed in the genus Maslovella (Black) Tappan and Loeblich 1966 because of the two distinctive shields and the infilled central area. It differs from all the previously described species belonging to this genus in the detail and delicate construction of the central area. CONCLUSION The South African material studied is very rich in the remains of these algae and affords excellent opportunity for a more detailed investigation of the coccolith flora of both earlier and later horizons. The work reported in this paper is only a portion of the work done on the Upper Cretaceous of Zululand, South Africa. The remainder of the new species and variations of existing species will be described in later papers. At present a detailed account of the microstratigraphy of the Cretaceous of Zululand is being prepared. The age of the strata from which the coccoliths were described was determined by the presence of Maestrichtian foraminifera (De Gasparis 1967); this conclusion was supported by the occurrence of Maestrichtian coccoliths. Acknowledgements. The author is indebted to Anglo-Transvaal Consolidated Investment Company who sponsored this study and provided the material. To Dr. G. F. Hart the author is grateful for his advice and help. REFERENCES black, m. 1964. Cretaceous and tertiary coccoliths from Atlantic seamounts. Palaeontology, 7, 307-16. bramlette, m. m., and MARTINI, E. 1964. The great change in the calcareous nannoplankton fossils between the Maestrichtian and the Danian. Micropalaeontology, 10, 291-322. de gasparis, m. l. 1967. The microstratigraphy of the Cretaceous System of Zululand. M.Sc. Thesis, Univ. Witwatersrand. loeblich, a. r. jr., and tappan, h. 1963. Type fixation and validation of certain calcareous nanno- plankton genera. Proc. biol. Soc. Washington, 76, 191-6. pienaar, r. n. 1966. Microfossils from the Cretaceous System of Zululand as studied with the electron microscope. S. Afr. J. Sci. 62, 147-57. stradner, h. 1961. Vorkommen van Nannofossilien im Mesozoikum und alttertiar. Erdoel-Ztschr. 77, 3, 77-88. tappan, h., and loeblich, a. r. jr. 1966. Maslovella nom. nov. Taxon, 15, 43. RICHARD N. PIENAAR Department of Plant Biology University of Natal Durban South Africa Manuscript received from author 10 July 1967 A NEW EOCENE C ASSIG ERINELLA FROM FLORIDA by W. G. GORDEY Abstract. A new species of Cassigerinella (C. eocaenica ) is proposed for specimens obtained from an Eocene sample taken off Blake Plateau, Florida. The species is similar, but quite distinct from the Oligocene marker C. chipolensis (Cushman and Ponton 1932). Therefore the stratigraphical value of the overlap in ranges of C. chipolensis with Pseudohastigerina micra (Cole 1927) as indicators of basal Oligocene is not affected. The joint occurrence of the species Pseudohastigerina micra (Cole 1927) and Cassi- gerinella chipolensis (Cushman and Ponton) has been regarded as a means of recognizing the basal Oligocene, at least within the tropical and semi-tropical belts. This overlap was first recognized by Blow and Banner (in Eames et ah 1962, p. 68) as a further criterion for the recognition of their zone of Globigerina sellii Borsetti 1959 (= G. oiigocaenica Blow and Banner 1962). Saunders and Cordey (1965) also recognized this overlap in a study of the Oceanic Formation of Barbados. It occurred in samples immediately overlying deposits of definite Eocene age containing such forms as Hantkenina spp., Globorotalia centralis, and G. cerro-azulensis. Bolli (1966), in a general review of world-wide plank- tonic zonations, added further support to the stratigraphic value of this observed overlap. Saito and Be (1963) established an Oligocene age (sensu Eames et al., op. cit.) for the Vicksburg group of the Gulf Coast region. However, they stated (p. 704) that C. chipo- lensis and P. micra occurred together with Eocene forms (e.g. Hantkenina alabamensis, H. primitiva, Globorotalia cerro-azulensis) from a core taken off the Florida coast (Lamont Core A167-21, 29° 49' N., 79° 39' W.). This record therefore cast considerable doubt on the stratigraphic value of this overlap as far as the recognition of the Eocene- Oligocene boundary or basal Oligocene was concerned. Blow examined these cassi- gerinellids and concluded (pers. comm.) that they were different from C. chipolensis. Through the kindness of Drs. Be and Saito the writer obtained material which contained the same form of Cassigerinella with Hantkenina spp., but from a location to the north of the Lamont Core, at Blake Plateau, 30° 04-8' N., 79° 14-5' W., Sample J.6B. A careful comparison of the cassigerinellids in the Blake Plateau material with C. chipolensis from both the Globigerina ampliapertura zone of Trinidad and the Oceanic Formation of Barbados indicates that the Eocene specimens differ from C. chipolensis. In view of the stratigraphic significance of C. chipolensis , it is desirable that a new species should be erected for these Eocene occurrences. SYSTEMATIC DESCRIPTIONS Genus cassigerinella Pokorny 1955 Cassigerinella eocaenica sp. nov. Text-fig. 1 a-e Description. Test very small, calcareous, perforate; 8-9 chambers in the final whorl, showing a gradual and uniform increase in size, moderately inflated. Initial two (possibly [Palaeontology, Vol. 11, Part 3, 1968, pp. 368-70.] CORDEY: A NEW EOCENE C ASS IG ERIN ELL A FROM FLORIDA 369 three) chambers planispirally arranged, later becoming biserial alternating, but coiled in the same plane. Periphery lobulate, initially sub-acute, becoming sub-rounded, sutures distinct, depressed, curved. Aperture a latero-marginal, extra-umbilical arch. TEXT-FIG. 1. a-e, Cassigerinella eocaenica sp. nov. Sample J-6b, 30° 04-8' N., 79° 14-5' W., 215' from top. a-c, Holotype, BMNH P46838; d, e, Paratype, BMNH P46839; both X 300. f-m, Cassigerinella chipolensis (Cushman and Ponton), f-h. Dissected hypotype showing penultimate whorl, early chambers planispirally coiled, later chambers alternating; G. ampliapertura Zone, Cipero Coast, Trinidad; BMNH P46840, X 213. i-k. After Cushman and Ponton, X 150. /, m, Hypotype, G. ampliapertura zone, Cipero Coast, Trinidad; BMNH P46841, x275. Remarks. C. eocaenica is similar to C. chipolensis, but differs in being consistently smaller, the greatest breadth of the final whorl varying from 0-1 to 0-12 mm., the average being 0-1 mm. Measurement of 45 specimens of C. chipolensis shows a variation in the breadth of the final whorl from 0T3 to 0T9 mm., the average being 0-16 mm. Secondly, the chambers in the final whorl are less inflated than in C. chipolensis. Specimens of C. chipolensis from Barbados (Oceanic Formation), from the Cipero Formation of Trinidad and from the Oligocene of Lindi (Blow and Banner in Eames et al. 1962, pi. 15, figs, m-n), show a more rapid increase in the size of the chambers of the final whorl 370 PALAEONTOLOGY, VOLUME 11 than in C. eocaenica. C. chipolensis also has a more rounded periphery throughout. Finally, the chamber arrangement in the final whorl of C. eocaenica varies from an initial planispiral arrangement to alternating, whereas the chamber arrangement in C. chipolensis is entirely alternating throughout the final whorl. The planispiral arrange- ment is only developed in the first two or three chambers of the penultimate whorl of C. chipolensis (text-fig. 1 ,f-h). The only other species which shows any morphological similarity to C. eocaenica is C. globolocula Ivanova 1958. Pokorny (in Eames et al.) agreed that his species C. boudecensis was probably conspecific with chipolensis , and considered that globolocula was ‘certainly conspecific with boudecensis ’ (op. cit., p. 83). The writer would agree with this conclusion and therefore the above remarks on C. chipolensis and C. eocaenica apply equally to C. globolocula (and also C. boudecensis). Deposition of types. Holotype and paratype specimens are deposited in the British Museum (Natural History). An unfigured paratype is deposited in the United States National Museum, No. 643514. Material. Fifteen specimens of C. eocaenica', forty-five specimens of C. chipolensis. Acknowledgements. The author is grateful to Drs. Be and Saito (Lamont Geological Observatory, Columbia University, New York) for the donation of the samples upon which this study is based; Dr. R. Lagaaij and J. A. Postuma (Bataafse Internationale Petroleum Maatschappij N.V., The Hague), for their critical reading of the manuscript; and the Bataafse Internationale Petroleum Maatschappij N.V., for permission to publish this paper. REFERENCES bolli, h. m. 1966. Zonation of Cretaceous to Pliocene marine sediments based on planktonic Fora- minifera. Bo In inf. Asoc. Venezolana de Geologia, Mineraria y Petroleo, 9(1), 3-32. eames, F. e. etal. 1962. Fundamentals of Mid-Tertiary Stratigraphical Correlation. Cambridge University Press. ivanova, l. G. in bykova, n. k. 1958. Novye Rody i Vidy Foraminifera. Trudy vnigri, no. 115, Mikrofauna SSR, 9, 4-81. saito, t. and be, a. w. h. 1963. Planktonic Foraminifera from the American Oligocene. Science, 145, 703-4. saunders, j. b. and cordey, w. g. 1965. The biostratigraphy of the Oceanic Formation in the Bath Cliff section, Barbados. Proc. 4th Caribbean Geol. Congress, Port of Spain, 1965. (In press.) w. G. CORDEY Bataafse Int. Petr. Mij. EP/12, Carel van Bylandtlaan 30 The Hague Typescript received from author 29 June 1967 Netherlands MORPHOLOGY AND PHYLOGENY OF ORBULINOIDES BECKMANNII (SAITO 1962) by w. G. CORDEY Abstract. An examination of dissected specimens of the Eocene species Orbulinoides beckmannii (Saito 1962), formerly Porticulasphaera mexicana (Cushman 1925), reveals certain morphological features not hitherto described. A study of the ontogeny of O. beckmannii and a comparison with the contemporaneous species Globigerapsis kugleri Belli, Loeblich and Tappan, and Globigerinatheka barri Bolli, Loeblich and Tappan, indi- cate that O. beckmannii is unlikely to be related to either of these species. It appears to have developed from some globorotaloid ancestor. This study is based on specimens obtained from a block of extraneous Eocene in the Miocene Nariva Formation of Trinidad. This same material was used by Bolli (19576, p. 1 59) in his monograph on the Eocene faunas of Trinidad. Bolli, Loeblich and Tappan (1957, p. 34) gave a full and accurate description of their new genus Porticulasphaera, and Bolli (1957u, p. 116) in placing the Miocene species Globigerinoides glomerosa Blow 1965 in this new Eocene genus commented that detailed comparative studies would probably \ . . reveal differences between the Eocene and Miocene forms . . (op. cit., p. 115). Bolli, Loeblich and Tappan designated Globigerina mexicana Cushman 1925 as the type species of their new genus Porticulasphaera. Saito (1962) considered that the Cushman type should be referred to the genus Globigerapsis Bolli, Loeblich and Tappan. Further examination of G. mexicana by Saito, and independently by Blow, has led to the conclusion that G. mexicana Cushman is conspecific with Globigerapsis semiinvo/uta (Keijzer 1945). The writer has also examined the holotype of G. mexicana and agrees with their conclusion. Blow and Saito have therefore assigned the specimens which Bolli ( 1957a) previously referred to as Porticulasphaera mexicana (Cushman 1925), and upon which this study is based, to the new genus Orbulinoides as O. beckmannii (Saito 1962). Olsson (1965) subsequently erected the genus Praeorbulina for the Miocene forms (i.e. Porticulasphaera of Bolli 1957(7, p. 115) with the type species Globigerinoides glomerosa glomerosa Blow 1956. The object of this paper is to discuss certain morpho- logical features observed in O. beckmannii, and which have not previously been described. Morphology. Externally O. beckmannii shows numerous apertures at the base of the final chamber. The early trochospiral part of the test shows between one and four supplementary apertures; occasionally no such apertures were present. When present they are usually located at the junction of the spiral and intercameral sutures. The specimen illustrated by Bolli (1957a, pi. 37, fig. 1 a) showing eight such apertures is considered atypical (cf. Bolli, Loeblich and Tappan, 1957, pi. 6, figs. 9a, b). A removal of the final chamber and part of the walls of the last three chambers of the initial trochospire (text-fig. lc) reveals the numerous supplementary apertures described by Bolli, Loeblich and Tappan (1957) and Olsson (1965). However, a closer examination [Palaeontology, Vol. 11, Part 3, 1968, pp. 371-5.] 372 PALAEONTOLOGY, VOLUME 11 text-fig. 1. a-d, Orbulinoides beckmannii (Saito 1962). a. Dissected specimen showing earliest whorl with an umbilical-extra-umbilical aperture, BMNH P46842, x 76. b, c. Dissected specimens, BMNH P46843, 46844, x 105. d. Dissected specimen showing trochospiral supplementary aperture (ts) opening into the vestibule (v), BMNH P46846, x93. e-g, Globigerapsis kugleri Bolli, Loeblich and Tappan 1957. Dissected specimens showing globigerinid primary aperture ; e. Specimen with Bulla partly removed, X 1 20 ;/, Bulla and two chambers removed, x 120; g. Bulla and five chambers removed; BMNH P46847, X 105. h-j, Globigerinatheka barri Bolli, Loeblich and Tappan 1957. Dissected specimens showing a globi- gerinid primary aperture; h, Bulla partly removed, X 100; i, Bulla and two chambers removed, x 100; j. Bulla and five chambers removed; BMNH P46848, X 73. All specimens from the O. beckmannii Zone, Navet Formation, Eocene, Point-a-Pierre, Trinidad. CORDEY: O RB ULINO ID ES BECKMANNII (SAIT O 1962) 373 of these apertures shows that in no instance are they directly connected with the outside of the test. They open into a small cavity (here termed vestibule) between the thick outer wall and the delicate wall of the initial trochospirally arranged chambers (text-fig. le). This vestibule is usually situated at the junction of the spiral and inter- cameral sutures, but probably also extends some way along the intercameral suture. It is also seen that the external supplementary apertures on the initial trochospire were never aligned with the internal apertures. Therefore, there is only an indirect connexion, that is via the vestibule, between the inside and outside of the test, as far as the early supplementary apertures are concerned. Bolli’s comparison (op. cit., p. 115) of the supplementary apertures of Orbu/inoides and Globigerinoides cannot be upheld. There is a basic difference in that, in the latter genus, these apertures are never covered by subsequent thickening of the test and are in direct communication with the inside of the test. It would seem that early supplementary apertures became vestigial structures, in the sense that the degree of direct connexion between the inside and outside of the test in this area was much reduced. This appears to be in contrast with species of Globigerinoides (and also Orbulina suturalis Bronnimann 1951) where a direct connexion is maintained for each of the apertures. The method of test thickening in Orbulinoides also appears to be in contrast with certain fossil species of Globigerinoides (e.g. G. subquadratus Bronnimann 1954 = G. ruber (d’Orbigny 1826) of Bolli 1957a). In this genus thickening is progressive rather than occurring only after the adult stage has been reached. A consideration of the morphology of the initial trochospire of Orbulinoides (e.g. its large primary aperture, inflated spinose chambers, numerous large supplementary apertures, and thin wall) strongly suggests that during the trochospiral stage the animal inhabited the epipelagic zone ( sensu Hedgepeth 1957, p. 18, fig. 1). The subsequent test thickening, and the reduction in the number of supplementary apertures, might be correlated with its migration to deeper levels. The work of Be and Ericson (1963) and Be (1965) offers some support for this view, since a correlation between the thickness of the test with depth in the Recent species Globorotalia truncatulinoides and Globigeri- noides sacculifera (Brady) was observed. The reduction in the number of ‘functional’ supplementary apertures may have been one means of increasing its weight in order to occupy lower levels. It is equally possible that the thickening of the test and reduction in supplementary apertures are associated with reproduction. The final inflated chamber may represent a type of brood pouch. The presence of such an inflated final chamber in G/obigerapsis and Globigerinatheka may indicate a similar mode of reproduction in these genera. Le Calvez(1936) showed that in Orbulina minima (d’Orbigny) there was a gradual reduction in the globigerine part of the test with depth. Furthermore, in specimens from the deepest levels (about 300 m.) the globigerine chambers had completely disappeared, and all that remained was the spheroidal final chamber which was frequently found to be filled with gametes. There- fore, there is some support for the view that the inflated final chambers of these Eocene genera might be connected with reproduction. Phylogeny. Bolli (1957a, p. 160) stated that ‘. . . Globigerapsis, Globigerinatheka and Porticulasphaera obviously represent a related group’. He considered that Globigerapsis index (Finlay) gave rise to G. kugleri Bolli, Loeblich and Tappan, from which O. 374 PALAEONTOLOGY, VOLUME 11 beckmannii developed. His further support for this view was the fact that the three genera showed a 90 per cent, tendency to coil dextrally. Bolli is correct in the case of Globiger- apsis and Globigerinatheka, and Eckert’s (1963) study of these genera supports this con- clusion. However, in the writer’s opinion, Orbulinoides is unrelated to either of these genera. The gross morphological similarity of these three species is more likely to be a function of convergent evolution than any genetic affinity (at least as far as beckmannii is concerned). This view is based on a comparison of the ontogeny of G. kugleri , G. barri , and O. beckmannii (text-fig. 1). It is clear that both kugleri and barri have arisen from a globigerinid ancestral form, while O. beckmannii shows an unmistakable umbilical- extra-umbilical primary aperture initially, and therefore is derived from a globorota- loid ancestor. The striking differences in the morphology of the initial whorls, parti- cularly the distinctive shape of the later trochospiral chambers of beckmannii , and the large primary aperture, adds further support to this view (text-fig. 1, cf. figs. a , b with f-j). Bolli considered the gross similarity of the adult tests of these three species the most important phylogenetic factors. The writer, however, considers that the evidence of the ontogeny of these species outweighs the adult similarity in gross morphology. Acknowledgements. The writer is indebted to J. B. Saunders (Texaco Trinidad Inc.) for the material on which this study is based; Dr. R. Lagaaij and J. A. Postuma (Bataafse Internationale Petroleum Maatschappij N.V., The Hague) and Dr. J. F. Noorthoorn van der Kruijff (Koninklijke/Shell Explo- ratie en Produktie Laboratorium, Rijswijk) for much helpful discussion and criticism; Dr. W. H. Blow for providing a copy of the Blow and Saito MS.; Dr. R. Cifelli (U.S. National Museum) for permission to examine the types; and Bataafse Internationale Petroleum Maatschappij N.V. for per- mission to publish this paper. REFERENCES bandy, o. l. 1965. Restrictions of the ‘Orbulina’ datum. Micropaleontology, 12, 77-86, pi. 1. be, a. w. h. 1965. The influence of depth on shell growth in Globigerinoides sacculifera (Brady). Ibid. 11, 81-97. and ericson, d. b. 1963. Aspects of calcification in planktonic foraminifera (Sarcodina). Ann. N.Y. Acad. Sci. 109, 65-81. blow, w. h. 1956. Origin and evolution of the foraminiferal genus Orbidina d’Orbigny. Micropaleonto- logv, 2, 57-70. and saito, t. 1967. The morphology and taxonomy of Globigerina mexicana Cushman. (In press.) bolli, h. m. 1957a. Planktonic Foraminifera from the Oligocene-Miocene Cipero and Lengua Forma- tions of Trinidad, B.W.I. Bull. U.S. natn. Mus. 215, 97-123, pis. 22-29. 19576. Planktonic Foraminifera from the Eocene Navet and San Fernando Formations, Trinidad, B.W.I. Ibid. 215, 155-72, pis. 35-39. loeblich, a. r., and tappan, h. 1957. Planktonic foraminiferal families Hantkeninidae, Orbu- linidae, Globorotaliidae, and Globotruncanidae. Ibid. 215, 3-50. eckert, h. r. 1963. Die obereozanen G/obigerinen- Schiefer (Stad- und Schimbergschiefer) zwischen Pilatus und Schrattenfluh. Eclog. geol. Helv. 56, 1001-72. hedgepeth, w. 1957. In ‘Treatise on Marine Ecology and Paleoecology’. Mem. geol. Soc. Am. 67. le calvez, j. 1936. Modifications du test des Foraminiferes pelagiques en rapport avec la reproduction Orbidina universa d’Orb. et Tretomphalus bulloides d'Orb. Ann/s Protist. 5. CORDEY: ORBULINOIDES BECKMANNII (SAITO 1962) 375 loeblich, a. r., and tappan, H. 1964. Treatise on Invertebrate Paleontology, ed. r. c. moore, Part C, Protista 2, 2. Geol. Soc. Am. and Univ. Kansas Press. olsson, r. k. 1965. Praeorbulina Olsson, a new foraminiferal genus. J. Paleont. 38, 770-1. saito, t. 1962. Eocene Planktonic Foraminifera from Hahajima (Hillsborough Island). Trans. Proc. palaeont. Soc. Japan , n.s. no. 45, 209-25. w. G. CORDEY Bataafse Int. Petr. Mij. EP/12, Carel van Bylandtlaan 30 The Hague Typescript received from author 29 June 1967 Netherlands THE GASTRIC CONTENTS OF AN ICHTHYOSAUR FROM THE LOWER LIAS OF LYME REGIS, DORSET by JOHN E. POLLARD Abstract. The partial skeleton of a small ichthyosaur associated with the gastric mass is described from the Lower Lias of Lyme Regis. The gastric mass was oval in shape and composed of minute dibranchiate cephalopod hooklets in random orientation. Four distinct types of hooklet are recognized in these gastric contents. Examination of published records and museum specimens suggests that gastric contents composed of cepha- lopod remains are more commonly preserved than those of fish remains. A study of ichthyosaur coprolites shows a predominance of defecated fish remains and an absence of hooklets from these structures. The diet, mode of feeding, and digestive mechanism of Liassic ichthyosaurs, in comparison with a teuthophagous cetacean, the sperm whale, are considered. In April 1963 the skeleton of a small ichthyosaur was found in the shales of the Lower Lias on the foreshore west of Lyme Regis, Dorset. Unfortunately due to the exposed location of this specimen, in soft shaly-mudstone at about the half-tide level, only a short time was available for its extraction. Only the anterior part of the skeleton could be recovered consisting of parts of the skull, pectoral girdle, vertebral column, and ribs. Careful preparation of this material showed that the skeleton was crushed and slightly dismembered, but that the stomach contents were preserved as a dark discrete area under- neath the vertebral column and ribs. Such occurrences are fairly well known, but the good state of preservation and lack of dispersal of the stomach contents of this specimen make them worthy of detailed description, quantitative analysis, and discussion in terms of the feeding habits and digestive mechanism of the Liassic ichthyosaurs. DESCRIPTION Horizon and locality. The specimen was collected from the shales of the lower part of the Psiloceras planorbis Zone of the Lower Lias (Woodward and Ussher 1911, p. 38), at the south-east corner of Pinhay Bay, two miles west of Lyme Regis (National Grid Reference: SY 325907). The enclosing sediment was a poorly fossiliferous silty and shaly mudstone, which was interbedded with thin lime- stones and shales containing Liostrea liassica, Hemicidaris spines, and rarely P/agiostoma gigantea and Psiloceras planorbis. No other vertebrate remains or coprolites were observed at this horizon. Skeletal remains. The partial skeleton of the ichthyosaur extracted was 2 ft. (60 cm.) in length and consisted of the skull and parts of the vertebral column, pectoral and pelvic girdles, rib cage, and a paddle. (The specimen is now preserved in the collections of the Geology Department, University of Manchester, registration number SF.l.) Text-fig. 1 is drawn from a field photograph of the specimen in situ, and shows the relative positions of the bones and the gastric mass from the dorsal aspect. The prepared skeleton can be examined both dorsally and ventrally, Plate 72, figs. 1 and 2, and enables the individual bones to be identified. [Palaeontology, Vol. 11, Part 3, 1968, pp. 376 88, pis. 72-73.] J. E. POLLARD: GASTRIC CONTENTS OF AN ICHTHYOSAUR 377 The skull is 10 in. (25 cm.) long, crushed dorso-ventrally and twisted sinistrally. The anterior part of the snout is missing, but the premaxilla and eleven upper jaw teeth and thirteen lower jaw teeth on the right side of the mouth are visible on the upper surface (PI. 72, fig. 1). On the under surface of the skull both dentary bones are present and twenty-three upper jaw teeth, and sixteen lower jaw teeth, from the left side of the mouth (PI. 72, fig. 2). The teeth appear to be well formed typical ichthyosaur teeth, up to 13 mm. in length exposed, with smooth apices and bifurcating grooves on the crown. The form of the tooth crown is close to that of Ichthyosaurus communis Conybeare as figured by Owen (1881, pi. 24, fig. 5). The anterior part of the right orbit was present dorsally 0 6 1 r 1 r 0 10 20 12 + 30 40 18 _L_ ~r 50 24 in. 60 cm. text-fig. 1. Ichthyosaur skeleton in situ in the Lower Lias west of Lyme Regis, showing the relation- ship of the various bones to the gastric mass. Widely spaced fine stippling represents the shale matrix, while the closely spaced coarser stippling represents the gastric contents. (text-fig. 1) but no sclerotic plates were seen. The hind part of the skull is badly crushed, and the only other bones clearly recognizable are displaced fragments of the articular and basioccipital (PI. 72, fig. 1). The post-cranial skeleton is represented by a total of twenty-eight vertebrae and numerous fragments of single and double ribs. On the upper (dorsal) surface of the specimen, PI. 72, fig. 1, seven thoracic vertebrae occur in a row from 4 to 8 in. (10-12 cm.) behind the skull, and bound the gastric mass dextrally. The dorsal left boundary of the gastric mass is formed by a series of parallel double ribs (text-fig. 1 and PI. 72, fig. 1), while a complex of complete and broken double and single ribs are elsewhere compressed into the gastric mass dorsally. The anterior boundary of the gastric mass on the ventral side of the specimen is formed by the bones of the pectoral girdle (PI. 72, fig. 2). Parts of the left coracoid, humerus, and scapula are clearly recognizable and are impressed into the gastric mass ventrally. The interclavicle is present, and fragments of fifteen phalangeal bones of the left anterior paddle were found just beyond the humerus. The only other recognizable bones collected were a displaced pubis and three phalangeal bones of a posterior paddle, Plate 72, fig. 1, all occurring postero-dextrally of the gastric mass. The nature and arrangement of these skeletal remains suggest that there had been a fair amount of displacement of the bones during burial and that the gastric mass must c c C 5586 378 PALAEONTOLOGY, VOLUME 11 have been trapped in an unusually anterior position, crushed between the anterior- dorsal thoracic rib cage and the pectoral girdle. Gastric mass. The gastric mass of this specimen, text-fig. 1, was broadly oval in shape, compressed dorso-ventrally, 13-5 cm. long from anterior to posterior, and 8 -5-9-0 cm. wide from right to left of the skeleton. Only the anterior part of the mass (8 cm. anterior to posterior by 7 cm. right to left) was collected and prepared for further study (PI. 72, fig. 1 ; PI. 73, fig. 1). Estimates of the dorsal area of the stomach mass, measured from the field photograph vary from 85-5 to 95-5 sq. cm. or approximately 90±5 sq. cm. The depth or thickness of the dorso-ventral cross section of the gastric mass was measured accurately on the prepared specimen (PI. 73, fig. 1), using a travelling microscope, and varied from 0-25 to 0-75 cm., with a mean value of about 0-33 cm. The cleaned and prepared dorsal and ven- tral surfaces of the gastric mass, Plate 72, fig. 2; Plate 73, figs. 1 and 2, show that the stomach contents preserved consist of a densely packed mass of dibranchiate cepha- lopod hooklets and rare large quartz grains. These hooklets are packed in random orien- tation (PI. 73, fig. 2) in a matrix of finely crystalline calcite. The quartz grains are sub- angular or sub-rounded in shape from 0-25 to 1 -40 mm. in diameter, sparsely distributed on the dorsal surface, but occurring in considerable concentration in patches of the ventral surface of the mass (i.e. at point X on PI. 72, fig. 2). Three, or possibly four, distinctly shaped types of hooklet can be recognized in these contents, types A, B, C, and D of text-fig. 2. Type A is relatively short straight spinose form with a strongly bifid base, rather like an odontaspid shark’s tooth in shape. Type B is longer and more slender than type A, with a less pronounced base and a gentle curve along its length. Type C is broader bladed than types A and B, has distinct lateral flattening, and a strong, nearly 90°, hook. The base of type C is much less pronounced than on types A or B, but this character may be suppressed due to lateral flattening. Type D of text-fig. 2 is extremely rare in the gastric contents, about 1 mm. or less in size, and a specimen from a different horizon and locality is figured here for comparative purposes, the significance of which will be discussed later. Each of these hooklet types text-fig. 2. Cephalopod hooklets from the Lias. Types A, B, and C are all drawn from hooklets in the gastric contents shown on Plate 73, figs. 1 and 2, while type D is drawn from specimen OUM. J. 14800, in mudstone from the Upper Lias at Dumbleton, Gloucestershire. EXPLANATION OF PLATE 72 Fig. 1. Dorsal view of prepared ichthyosaur skeleton preserved with gastric contents. Lower Lias, Planorbis Zone, Lyme Regis, Dorset. SF.l. Geology Dept. Collections, University of Manchester. a., articular; v., vertebra; rib; sea., scapula; pd., paddle; pub., pubis. Fig. 2. Ventral view of ichthyosaur specimen SF.l. Symbol ‘X’ indicates the region of the ventral surface of the gastric mass with a concentration of quartz grains, d., dentary; sa, supra-articular; icl., interclavicle; cor., coracoid; hum., humerus. Palaeontology, Vol. 11 PLATE 72 2 POLLARD, Ichthyosaur gastric contents J. E. POLLARD: GASTRIC CONTENTS OF AN ICHTHYOSAUR 379 in the gastric contents shows considerable size range, for instance in terms of length, type A varies from 1-0 to 1-90 mm. (10 measured), type B from 0-70 to 3-00 mm. (16 measured), and type C from 0-9 to 2-90 mm. (22 measured). Types B and C appear to be commoner than type A in the mass. All these hooklets seem to be composed of a jet black and very brittle organic, possibly chitinous, material. They are all hollow although sometimes filled with crystal- line calcite. Due to their brittle nature and hollow centre, most of the hooklets are cracked and partially crushed and splinter if any attempt is made to separate them from the mass. In an attempt to determine the approximate number of hooklets on the dorsal surface of the gastric mass, the distribution of the hooklets in an area 2 cm. square was plotted from the enlarged photograph, Plate 73, fig. 2. The frequency of hooklets on this surface varied between 450 and 540 per sq. cm., with a mean of about 500 per sq. cm. This number represents only those hooklets that could be clearly identified and is, therefore, a minimal estimate. The total number of hooklets on the dorsal surface of the gastric mass, area 90±5 sq. cm., is about 45,000±7,000 (i.e. 90±5 X 500±50). It has proved very difficult to estimate the total number of hooklets in the gastric mass due to their being crushed and randomly orientated. The cross-sectional diameter of a number of hooklets of various sizes, uncrushed on the dorsal surface, varied from 0-20 to 0-50 mm. The mean depth of the gastric mass is 0-33 cm., so that allowing for parallel packing and no crushing, the hooklets would be from approximately 16 (3*30/0*2) to 7 (3-30/0-5) layers deep. Making an allowance for crushing and random packing, from 6 to 14 or 10±4 layers deep, would seem to be a reasonable estimate. Therefore, the total number of hooklets in the gastric mass is 45,000^7,000 x 10±4 = 478,000^ 250,000 or 478,000±53 per cent. Such a large error is unavoidable in such approximate calculations, but the figure gives some idea of the correct order of magnitude. DISCUSSION In order to understand the signifiance of the gastric contents previously described, and the precise nature of the dibranchiate remains they contain, a search has been made in the literature and other specimens have been examined in several British museums. The author does not intend this as an exhaustive treatment of the subject, but more as a spur to examination and comment by other workers. Other Liassic ichthyosaurs with gastric contents. Ichthyosaur remains with associated gastric contents preserved have been known for more than a hundred years from the Liassic shales of Lyme Regis and Whitby in England, and Holzmaden in Germany. Buckland in the Bridgewater Treatise (1836) is among the earliest English records. He described and figured (pi. 13 and 14) two ichthyosaur specimens from Lyme Regis that contained a coprolite mass with fish scales, preserved within the abdominal cavity. These specimens are in the collections of the Oxford University Museum and will be discussed later in this paper. Probably the earliest description of the preserved cephalopod hooklets associated with ichthyosaur bones is that of Coles (1853). He describes a layer of carbonaceous material made up of ‘ minute black points ’ — hollow and filled with calcite, that was found 380 PALAEONTOLOGY, VOLUME 11 adhering to an ichthyosaur vertebra from the Lias of the Tewkesbury district. His excellent figures (pi. 5, figs. 2 and 13) show shape, size, and crack patterns identical to the hooklets described and figured in this paper (PI. 73, fig. 2 and text-fig. 2). This material was wrongly identified by Coles (1853, p. 81) as ‘setiform or bristly scales’ of the ichthyosaur integument, and was reported by him to be known associated with ichthyosaur skeletons from the Lias of Lyme Regis and Ilminster as well as Tewkesbury. Cole’s error was corrected by Moore (1856), who reported finding stomach contents composed of cephalopod arm hooklets in sixteen out of twenty-three Liassic ichthyosaur skeletons he had prepared for his museum. Moore examined the gastric contents further, and suggested that they consisted of the desiccated ink and arm hooklets of naked Jurassic cuttle-fish allied to Onyehoteutliis. Buckman (1879), when describing a new species of fossil dibranchiate Belemnoteuthis montefiorei from the Lower Lias of Charmouth, mentions the frequent occurrence of ichthyosaur stomach contents and coprolites full of cephalopod arm hooklets. Similar general statements recording gastric contents composed largely of cephalopod hooklets have been made by several workers studying ichthyosaurs from the Holzmaden Lias (Seeley 1880, Branca 1908, Drevermann 1914, Huene 1922, Hofmann 1958, and Augusta 1964). Wurstemburger (1876) described a Holzmaden specimen of Stenopterygius quadriscissus, with head 50 cm. long, vertebral column 240 cm. long, where a large stomach mass of fish and cephalopod remains was found only 20 cm. behind the head. This unusually anterior thoracic position of the stomach is very similar to that of the specimen described here. Williston (1914, p. 123) refers to an ichthyosaur skeleton in the Stuttgart Museum that has preserved in its stomach contents the remains of more than 200 belemnites. Dr. K. D. Adam (pers. comm.) informs me that no such specimen exists in the Stuttgart Museum, but Williston’s comment is probably a mistaken reference to a well preserved specimen of the shark Hybodus from the Upper Lias of Holzmaden described by Brown (1900) and later Shimanskiy (1949). The gastric contents of this shark contain over 250 belemnite rostra. Many British museums possess in their collections ichthyosaur skeletons with well preserved gastric contents. On other specimens the gastric contents have obviously been cleaned off in the preparation of the skeleton, and so it would appear that these contents are of much commoner occurrence than the literature would suggest. The ichthyosaurs figured by Buckland (1836, pis. 13 and 14) are preserved in the Oxford University Museum, numbered specimens J. 13587 and J. 13593 respectively. Re-examination of these specimens by the author confirms that Buckland’s figures and descriptions are extremely accurate and that the gastric contents consist largely of scales and spines of the Liassic fish Pholidophorus sp., set in a matrix of a pale buff coprolitic clay. The larger specimen J.13593 (Buckland 1836, pi. 14) does have a very sparse scattering of type C hooklets over the whole dorsal surface of the gastric mass. Three other specimens EXPLANATION OF PLATE 73 Fig. 1 . Dorsal view of the gastric mass and associated bones of ichthyosaur specimen SF. 1 (compare with PI. 72, fig. 1). Scale of 1 cm. Fig. 2. Magnified view of part of the dorsal surface of the gastric mass of specimen SF. 1 showing the various types of cephalopod hooklets present. (This field of view may be orientated on PI. 73, fig. 1, by the arcuate row of five large quartz grains in the lower half of the picture.) Scale of 1 cm. Palaeontology, Vol. 11 PLATE 73 POLLARD, Ichthyosaur gastric contents J. E. POLLARD: GASTRIC CONTENTS OF AN ICHTHYOSAUR 381 of Liassic ichthyosaurs from Lyme Regis in Oxford University Museum, J. 12125, J. 13592, and J. 10348, all contain patches of gastric material, composed of types A, B, and C hooklets, in the thoracic or anterior abdominal regions. The gastric mass in each case is identical to the specimen described here, just hooklets without any matrix of the coprolitic clay seen in Buckland’s specimens. Specimens of various species of ichthyosaur from the Lower Lias at Lyme Regis on display in the public galleries of the British Museum (Natural History) show gastric contents of densely packed hooklets devoid of matrix (e.g. BMNH 36256, BMNH R1614, BMNH R1072, BMNH 38523, BMNH 43006, and BMNH R1896). In the Manchester Museum an excellent specimen from the Upper Lias at Whitby has a large gastric mass containing A, B, and C type hooklets, just posterior to the pectoral girdle. The conclusion to be derived from a study of these listed, and other specimens, is that gastric contents of densely packed dibranchiate cephalopod hooklets are much commoner in prepared specimens than the fish remains in a matrix of coprolite clay described by Buckland. The gastric contents of many Jurassic plesiosaurs are also known to be composed largely of dibranchiate hooklets (Juravlev 1943, Hekker and Hekker 1955, and Tarlo 1959), but here large gastroliths usually occur as well (Seeley 1877, Brown 1904, and Williston 1904). Gastroliths have not been found preserved in ichthyosaur stomach contents. Contents of coprolites from Lower Lias. Well-preserved coprolites, usually assigned to ichthyosaurs or plesiosaurs, have been known from the Lower Lias at Lyme Regis since before Buckland's classic paper of 1835. Most of these coprolites are assumed to have been formed by ichthyosaurs, on account of their similarity in composition to material described by Buckland (1836) from within the ichthyosaur abdominal cavity. Other workers (Fraas 1891 and Woodward 1917) have questioned the assignment of these coprolites to ichthyosaurs. They argue that spirally folded coprolites are rarely found associated with ichthyosaur skeletons and are more likely to have been formed by the spiral intestine of Liassic sharks than by the typical reptilian intestines which the ichthyosaurs probably possessed. The following analysis shows that the majority of Liassic coprolites do not have well-formed spiral folds but have faunal, lithological and chemical features identical to the ichthyosaur gastric contents described by Buck- land. Moreover, the hybodont sharks, suggested by Woodward (1917) as probable producers of the coprolites, are believed on account of their dentition (Romer 1966, p. 40) to have been benthonic or necto-benthonic scavenger feeders, and not nectonic fish feeders as were the producers of the Liassic coprolites. Buckland (1835) showed that Lower Lias coprolites contain fish remains, bones of young ichthyosaurs, and possibly the sucker rings of fossil cuttle-fish. He did not observe or describe any dibranchiate hooklets. The matrix of these coprolites, which I have called lbuff coprolitic clay’, was shown by Buckland to be phosphatic material derived from the digestion of fish and reptile bones. Buckland suggested that the strong spiral involutions frequently seen on coprolites indicate that the small intestine of the ichthyo- saur was ribbon like and twisted into a spiral. Firtion (1938) analysed the contents of coprolites from the Lower Lias of Alsace. He found that the undigested contents were mainly crinoids, gastropods, or pelecypods with less abundant foraminifera, ostracods, fish remains, and brachiopod shells. These coprolites rarely had spiral folds and were 382 PALAEONTOLOGY, VOLUME 11 obviously formed by a benthonic feeding animal. However, he assigned them to ichthyo- saurs, although stomach contents of the above composition are unknown in ichthyo- saurs. Such coprolites could as easily belong to teleosaurs or plesiosaurs (Drevermann 1914). Ager (1963, p. 120) mentions that a study of coprolites suggests that Mesozoic ichthyosaurs included belemnites in their diet, but he does not refer to any actual records of this fact. Fifty well-preserved coprolites from Buckland’s collection in the Oxford University Museum have been examined by the author. These specimens vary from 1 to 6 in. in length, mainly 2-3 in. long, and their contents can be examined on the cleaned surface, or internally where they have been sectioned and polished. Forty-five specimens contain recognizable fish remains, mainly scales, fin rays, and spines of Pholidophorus sp., less commonly remains of Dapedium sp. and Lepidotus sp. Two specimens, those figured by Buckland (1835, pi. 29, figs. 2, 3, 4, and 5), contain reptilian bones and fifteen specimens have well-formed spiral involutions. None of these coprolites contain visible remains of dibranchiate hooklets and the possible sucker rings figured by Buckland (1835, pi. 30, figs. 1 , 2, and 3) are considered to be transverse sections of fin rays and small vertebrae of fish. Examination of sixteen well-preserved coprolites in Manchester Museum and Geology Department, University of Manchester, shows that all these contain fish scales and spines, none contain reptilian bones, and only four specimens have well-formed spiral folds. One of the Manchester Museum specimens has a small patch of shale matrix with type B and C hooklets adhering to its surface, but they are not contained in the adjacent coprolite material. Lydekker (1889, pp. 1 1 4—17) lists sixty-six coprolites from the Lower Lias in the British Museum collections, but only mentions fifteen con- taining fish scales, only two with reptilian bones, and only one showing well-formed vascular impressions. From this survey it is suggested that ichthyosaurs from the British Lower Lias primarily defecated fish remains in their coprolites. Undigested cephalopod remains have not been seen in these coprolites, so it may be inferred that they accumulated in the stomach, as their predominance in gastric contents would suggest. The commonest form of fish eaten, Pholidophorus , is presumed to have been a nectonic or necto- benthonic form, not a deep bodied benthonic fish like Dapedium or Lepidotus. The possible significance of this latter observation will be discussed later. Nature of the cephalopod remains. Throughout the preceding part of this paper the hooklets found in the gastric contents of the ichthyosaur have been broadly described as belonging to Liassic dibranchiate cephalopods. It is of some importance to establish the precise nature of the dibranch iates possessing these hooklets before discussing their relationship to their ichthyosaur predators. As well as occurring in gastric contents of ichthyosaurs, these hooklets are known preserved in their life position on the arms of predominently soft bodied dibranchiates that are rarely found in Liassic and other Jurassic argillaceous sediments. Pearce (1842) named one of these soft-bodied dibranchiates with arm hooklets from the Oxford Clay as Belemnoteuthis. The arm hooklets of Belemnoteuthis were figured by Owen (1844, pi. 6, fig. 2) and Mantell (1852, fig. 4) as all possessing elongated pointed bases, and one form similar in shape to type D of text-fig. 2. It has already been mentioned that Coles (1853) seems to have been the first person J. E. POLLARD: GASTRIC CONTENTS OF AN ICHTHYOSAUR 383 to describe the hooklets from the Lias, although he was in error about their nature. The suggestion of Moore (1856), that these hooklets belonged to naked dibranchiates allied to Onychoteuthis, does not seem to be supported by any earlier published records of specimens of this genus from the Liassic rocks of Britain. Huxley (1864) was the first person to unquestionably associate these forms with belemnoid arm hooklets. He figured (1864, pi. 1, fig. 5a) forms identical with the types A, B, and C of text-fig. 2 when he described two specimens of belemnites from the Lias (BMNH 74106 and BMNH 39855), where soft parts were preserved in association with the guard, phragmocone, and proostracum. One of these specimens, BMNH 74106, must be interpreted with some caution as it has been restored in preparation. Dr. K. A. Joysey (pers. comm.) has informed me that one such specimen (J42835) in the Sedgwick Museum, Cambridge, is a ‘well intended’ forgery by a preparator, for the belemnite guard has been artificially shaped to improve its appearance before being set in an artificial matrix in association with a genuine group of hooks. However, Huxley’s interpretation appears to be correct and has been accepted by such later workers as Crick (1907), Naef (1922), and Jeletsky (1966). Buckman (1879) described a specimen of a head of hooked arms from the Lower Lias, which he named Belenuioteuthis montefiorei. This specimen is in the collection of the British Museum (BMNH C5026) and the hooklets are identical with the types A, B, and C described here. However, in affinity this specimen seems to be Belemnites as suggested by Crick (1902) and not the Belenmoteuthis of Pearce (1842). The most complete study of the arms of Liassic and other Jurassic dibranchiates is undoubtedly that of Crick (1907). He examined seventeen specimens of ‘belemnite’ arms in the British Museum collections and described and figured six of these specimens in detail in his paper. Crick concluded that the belemnites possessed six arms bearing rows of hooklets with swollen bases, as types A, B, and C, while Belenmoteuthis (= Acan- thoteuthis) had eight or ten arms bearing hooklets with pointed bases. This latter form of hooklet is characteristic of fossil dibranchiates known from the Upper Jurassic Oxford and Kimmeridge Clays in Britain, and the lithographic stone of Solenhofen in Germany (Pearce 1842, Owen 1844, Mantell 1852, Crick 1897 and 1907). Several of the standard textbooks on palaeontology (Zittel 1913, Woods 1946, and Piveteau 1953) figure dibranchiate cephalopod hooklets from Mesozoic sediments but give little idea of possible affinities of the various forms. Naef (1922) in his authoritative work on fossil dibranchiates figures and discusses various forms, including types from the Upper Lias of Holzmaden similar to types B and C of this paper, but is uncertain of any definite correlation of hooklet form with type of dibranchiate. Both Naef (1922, p. 219) and Jeletsky (1966, p. 138) disagree with Crick’s (1907) interpretation of six- arm belemnites and suggest that they had eight or ten arms in common with the belem- noteuthids. Jeletsky (1966, p. 138) believes that all members of the order Belemnitida possessed arm hooklets. From a detailed study of the literature and museum specimens it is here suggested that some broad association of hooklet form with three major groups of Mesozoic dibranchiate cephalopod may be possible. Members of the Belemnitidae may have had hooklet types A, B, and C as described here, characterized by a gentle curved shape and a swollen bifid base. Such forms are known mainly from the Lias (Huxley 1864, Crick 1907, Naef 1922, and Jeletsky 1966). Dibranchiates of the family Belemnoteuthidae may 384 PALAEONTOLOGY, VOLUME 11 have had gently curved or recurved hooklets with an elongate obliquely pointed base. This hooklet type is known from Upper Jurassic specimens (Pearce 1842, Owen 1844, Mantell 1852, Crick 1897, 1907, Hekker and Hekker 1955). Genera of the order Phragmoteuthida probably had hooklets of a belemnitid type (Jeletsky 1966, p. 31). Separated hooklets of a variety of shapes are frequently found in microfaunas of Jurassic and Cretaceous age and are known as ‘ Onychites' sp. (Quenstedt 1885, Naef 1922, Piveteau 1953, Hekker and Hekker 1955). Such hooklets probably belonged to other little-known members of the order Belemnitida as fossil teuthoid squids and sepioid cuttle fish seem to have been devoid of arm hooklets in Mesozoic seas (Jeletsky 1966). Therefore, the types of hooklets described in the earlier part of this paper would seem to have belonged to dibranchiate cephalopods of the family Belemnitidae, and not the family Belemnoteuthidae or the order Phragmoteuthida. TABLE 1 Aims and arm hooklets of fossil belemnoids from the Lower Lias. Details from Woods (1946, fig. 169) and Crick (1907, pi. 23, figs. 1, 3, and 5). Number Total no. Hooklets Number of hooklets per arm of of of per arm lengths Specimens arms hooklets mean max. > 40 mm. > 30 mm. > 20 mm. SM. J37812 8 211 29 34 32 (5) 0 17 (3) BM. C3007 6 143 24 28 26(3) 21 (3) 0 BM. 82895 4 123 31 36 31 (4) 0 0 BM. 47020 6 156 25 30 26 (2) 26 (2) 24 (2) Approx, ‘mean’ 6 c. 150 27 32 29 23 20 Examination of published figures and museum specimens of ‘belemnite’ arms with hooklets (e.g. Woods 1946, fig. 169, and Crick 1907, pi. 23, figs. 1 to 6) has enabled observations to be made regarding the number of hooklets per arm, the total number of hooklets per belemnite individual, and the arrangement of hooklet types along the arms. The varied state of preservation of specimens with arms from the Lower Lias makes detailed analysis very difficult. Crick (1907) showed that of the seventeen specimens in the British Museum collections only six were worthy of description, and of these six only three are considered sufficiently complete by the present author for detailed analysis (Table 1). Crick (op. cit.) showed that the belemnite arms varied in length, and that the hooklets were arranged in two parallel rows on the inner surface of the arms. In many of the known specimens the arms are either incompletely preserved or superimposed, so that it is impossible to be certain of the original arrangement of the arm hooklets. The specimens listed in Table 1, although varying in the number of arms, all possess arms that appear to be complete, as the hooklets are arranged in parallel rows, and are largest in the mid-length of each arm, gradually diminishing in size towards each end (Crick 1907, p. 271). From Table 1 it appears that the number of hooklets per arm varies with the length of the arm, but about thirty (fifteen pairs) hooklets per arm is an average number. Specimen SM. J37812 possesses eight distinct arms, and therefore confirms that there must have been eight or more arms in the belemnites in agreement with Naef (1922) and Jeletsky (1966). There must have been, therefore, at least 300 hooklets on an individual ten-armed belemnite in Liassic times. J. E. POLLARD: GASTRIC CONTENTS OP AN ICHTHYOSAUR 385 The detailed arrangement of the hooklets along the arms can only really be seen on two of the specimens included in Table 1, Wood’s fig. 169 (SM. J37812) and BMNH 82895 (Crick 1907, pi. 23, fig. 3). On these specimens the proximal hooklets are seen to be small examples of types A and B. The hooklets at mid-length of the arms are large examples of types B and C which decrease in size towards the distal end. This broad pattern of hooklet arrangement can be seen to a lesser extent on other less well preserved specimens (e.g. Crick 1907, pi. 23). Significance of the gastric contents (a) Feeding habits Diet. The evidence of the gastric contents and coprolites suggests that Liassic ichthyo- saurs fed mainly on fish and/or dibranchiate cephalopods. Among fish-eaters stomach contents were rarely preserved while coprolites are commonly found. A reverse situation possibly exists regarding dibranchiate eating forms, stomach contents being commonly preserved but coprolites rarely. Both these diets suggest that ichthyosaurs in Liassic seas were predators on nectonic not benthonic animals. If at least two distinct dietary habits were established amongst Liassic ichthyosaurs, they could be explained by either selective predation, or feeding at different levels in the sea as in the sperm whale (Clarke 1962, p. 186). Drevermann (1914, p. 42) has suggested that in Upper Jurassic seas, virtually toothless ichthyosaurs like Ophthahnosaurus may have fed exclusively on naked cuttle fish. The suggestion put forward in an earlier section of this paper that some Liassic ichthyosaurs fed mainly on belemnoids is made with some reservation, as the known fossil belemnoids from the Lias where arm hooklets and hard parts occur in association are few and none too well preserved. The similarity of the ichthyosaurs in mode of life and diet to odontocete Cetacea, especially the sperm whale, has been sug- gested by several workers (e.g. Buckland 1835, p. 227, Kukenthal 1892, Branca 1908, and Wiman 1946). Volume of food eaten. In the previous sections it has been shown that the gastric contents contain 478,000±250,000 hooklets, and that each belemnite probably had about 300 arm hooklets. The gastric contents described here, therefore, could represent between 760 and 2,430 digested individual belemnites. The undigested organic hard parts of these belemnites accumulated in the stomach, much in the same way as arm hooklets and beaks of modern dibranchiates accumulate in the stomach of the sperm whale. Akimuskin (1955) records 28,000 squid beaks, representing 14,000 squids, and Clarke (1962) records 4,000 beaks plus 28 undigested squids, representing 2,160 squids, found in the stomachs of sperm whales caught in the North Pacific and Atlantic respectively. The length of time of accumulation of these belemnoid hooklets in the ichthyosaur stomach is impossible to ascertain. It could represent several meals or a lifetime’s accumulation. Judging from the length of skull and relative size of the various bones it is probable that the ichthyosaur described was only a young specimen. In this particular specimen, therefore, the gastric contents might represent a lifetime’s accumulation. Mode of feeding. Several workers (i.e. Buckland 1836 and Seeley 1880) have suggested that the ichthyosaurs were voracious feeders, where the prey was swallowed whole without mastication, as in the sperm whale (Clarke 1956 and Clarke 1962). In the 386 PALAEONTOLOGY, VOLUME 11 teuthophagous whales the presence or absence of teeth makes very little difference to the efficiency of feeding (Clarke 1956). If a parallel situation existed amongst dibranchiate- eating Jurassic ichthyosaurs, the presence or absence of teeth could not be explained purely in terms of diet as suggested by Drevermann (1914). The sperm whale frequently has facial scars from squid tentacles (Clarke 1956), suggesting that the squids were swallowed head first. If belemnoids were eaten in this manner by Liassic ichthyosaurs, biting of the heads would cause separation of the hooked arms from the body with hard parts. Further significance of this comment is discussed below. (b) Digestive mechanism The ichthyosaurs appear to have swallowed their prey whole into a large expandable stomach where all the digestive breakdown took place (Buckland 1836). Chyme, including softened fish and reptile bones, passed into the intestine, and then the indigestible mater- ial was defecated as coprolites. In outline this digestive process is similar to that of the sperm whale (M. R. Clarke, pers. comm.). Undigested dibranchiate remains, hooklets, and possibly beaks accumulated in the stomach as they could not be passed on. The reasons for this accumulation are difficult to understand. Possibly the process was a defence mechanism on the part of the ichthyosaur to protect the delicate tissues of the posterior part of the digestive system from damage by the sharp undigested hooklets. The gastric contents could represent a gravity accumulation of indigestible material on the ventral side of the stomach in a very fluid chyme, produced by the digestive break- down of cephalopod tissue. Such an explanation would account for the pockets of quartz grains found on the ventral side of the gastric mass (PI. 72, fig. 2). A very fluid chyme might not have been able to transport the undigested matter through the pyrolic valve, as undoubtedly occurred with the viscous chyme produced from the digested fish remains, now preserved as coprolites. A third possibility is that the hooklets gripped in the ventral stomach wall and formed an interlocking network, trapping the quartz grains. This last suggestion could explain why the gastric contents of this specimen were not dispersed before burial, despite the slight dismembering of the skeleton. The accumulation of such remains raises a number of problems for the ichthyosaur which can only be answered by further analogies with the sperm whale. Did these remains accumulate to the detriment of the animal, perhaps blocking the digestive tract and causing death? In the sperm whale the accumulated squid beaks are periodically vomited from the stomach (M. R. Clarke, pers. comm.). Ambergris frequently contains squid beaks and so may also aid this regurgatory process. Such mechanisms as these may have existed in the ichthyosaurs to clear inconvenient accumulations of hooklets from the stomach. If Liassic ichthyosaurs frequently ate belemnoid dibranchiates, as has been suggested, another problem arises concerning the digestion of the crystalline calcite guards. Modern teuthophagous cetaceans have little difficulty in digestion of the conchiolinic or weakly calcified ‘pens’ of squids by solution in the stomach, but densely crystalline guards of the belemnites are much more difficult to destroy, as witnessed by their abundance in Mesozoic clastic sediments. As gastric contents with belemnoid guards are rare, or unknown, in ichthyosaurs three possible explanations are suggested: 1. Heads were bitten off and so guards were not swallowed. This seems an unlikely mechanism as in reptiles and modern teuthophagous cetaceans prey is swallowed whole. J. E. POLLARD: GASTRIC CONTENTS OF AN ICHTHYOSAUR 387 2. The body with guard, phragmocone, and proostracum separated from the head in the stomach and was regurgitated. The habits of the sperm whale give some support for this suggestion as regurgitation of squids frequently occurs on capture of the whales, and in the digestive process in the stomach the bodies and heads of the squids separate at an early stage (Clarke 1956 and Clarke 1962). 3. The guards were dissolved, or broken up by gastroliths and then dissolved, by the chemical environment of the stomach. Difficulties with this possibility are that gastro- liths are unknown in ichthyosaurs, and a very acid stomach environment would be necessary to have dissolved dense primary crystalline calcite. However, of these suggestions processes 2 and 3 would seem to be the more likely. Other possible explanations may simply be that the ichthyosaurs, as modern cetaceans, primarily ate naked dibranchiates, or chemical solution of guard posed no problem for the digestive mechanism of the Liassic ichthyosaur. Acknowledgements. I thank Dr. F. M. Broadhurst for help in extraction of the specimen and critically reading the manuscript, Dr. K. D. Adam, Dr. K. A. Joysey, and Dr. M. R. Clarke for offering many helpful suggestions, and Mr. P. A. Washbourne for help with photography. Mr. J. M. Edmunds and Mr. P. Powell of the Oxford University Museum gave access to Buckland’s original material; the Assistant Keeper and Dr. M. K. Howarth of the British Museum (Natural History) and Dr. R. M. C. Eager of the Manchester Museum allowed me to examine specimens in their collections. REFERENCES ager, d. v. 1963. Principles of Palaeoecology. New York. 371 pp. Akimushkin, i. i. 1955. [The feeding of the Cachalot.] Dokl. Akad. Nauk S.S.S.R. 101, 1139-40. (In Russian.) AUGUSTA, j. 1964. Prehistoric Sea Monsters. London. 65 pp., 21 pi. branca, w. 1908. Nachtrag zur Embryonenfragen bei Ichthyosaurus. Sber. preuss. Akad. IT7vv(1908), 1 , 392-6. brown, b. 1904. Stomach stones and food of plesiosaurs. Science, 20, 184-5. brown, c. 1900. Uber das Genus Hybodus und seine systematische Stellung. Palaeontographica, 46, pp. 163 ff. buckland, w. 1835. On the discovery of Coprolites, or fossil faeces, in the Lias at Lyme Regis, and in other formations. Trans, geol. Soc. Load. (2), 3, 223-38, pis. 28-31. 1836. The Bridgewater Treatises', Geology and Mineralogy. London. 2 vols. buckman, j. 1879. On the Belemnoteuthis montefiorei. Proc. Dorset nat. Hist, antiq. Fid. Club. 3, 141-3, pi. 1. clarke, m. r. 1962. Stomach contents of a Sperm Whale caught off Madeira in 1959. Norsk. Hval- f angst t id. 5, 173-91. clarke, r. 1956. Sperm Whales of the Azores. ‘ Discovery ’ Rep. 28, 237-98. coles, h. 1853. On the skin of Ichthyosaurus. Q. JI. geol. Soc. Lond. 9, 79-82, pi. 5. crick, g. c. 1897. Acanthoteuthis speciosa Munster. Geol. Mag. (4), 4, 1-4, pi. 1. 1902. Note on the type specimen of Belemnoteuthis montefiorei. Proc. malac. Soc. Lond. 5, 13-16, pi. 1. 1907. On the arms of the belemnite. Proc. malac. Soc. Lond. 7, 269-79, pi. 23. drevermann, f. i. 1914. Die Meersaurier im Senkenberg Museum. Senkenberg Naturforsch. Ges. Ber. 45, 35-48, 12 figs. firtion, f. 1938. Coprolithes du Lias Lnferieur d’Alsace et de Lorraine. Bull. Serv. Carte, geol. Als. Lorr. 5, 27-43, pis. 4-8. fraas, e. 1891. Die Ichthyosaurier der sud-deutschen Trias und jura-ablagerunge. Tubingen. hekker, e. l., and hekker, r. f. 1955. Ostatki Teuthoidea iz verkhnei yuryi nizhnego mela Povolzhya. Vop. Paleont. 2, 36-44, pis. 1 and 2. (In Russian.) 388 PALAEONTOLOGY, VOLUME 11 hofmann, j. 1958. Einbettung und Zerfall der Ichthyosaurier im Lias von Holzmaden. Meyniana , 6, 10-55. 30 figs. huene, F. von. 1922. Die ichthyosaurier ties Lias und ihre Zusammenhange. Berlin. 114 pp., 22 pis. huxley, t. h. 1864. On the structure of the Belemnitidae; with a description of a more complete Belemnites than any hitherto known, and an account of a new genus of Belemnitidae Xiphoteuthis. Mem. Geol. Surv. U.K. Monogr. II. 22 pp., 3 pis. jeletsky, j. a. 1966. Comparative morphology, phylogeny and classification of fossil Coleoidea. Paleont. Contr. Univ. Kans., Mollusca 7, 1-162, pis. 1-25. juravlev, k. j. 1943. The remains of Upper Jurassic sea reptiles at the Seveljevra shale mine Acad. Sci. U.S.S.R. Biol. 5, 293-306, 4 figs. (In Russian, English summary.) kukenthal, w. 1892. Ichthyosaurier und Wale. Neues. Jb. Miner. Geol. Paleont. Beil Bd. 1, 161-6. lydekker, R. 1889. Catalogue of fossil Reptilia and Amphibia in the British Museum. Part II. Ichthyo- pterygia and Sauropterygia. London. 303 pp. mantell, g. a. 1852. A few notes on the structure of the Belemnite. Ann. Mag. nat. Hist. (2), 10, 14-19. moore, c. 1856. On the skin and food of ichthyosauri and teleosauri. Rep. Br. Ass. Advmt. Sci. (1856), 69. naef, a. 1922. Die fossilen Tintenfische. Jena. 322 pp. owen, r. 1844. A description of certain Belemnites preserved with a great proportion of their soft parts in the Oxford Clay. Phil. Trans. R. Soc. 134, 65-85, pis. 2-8. 1881. The reptilia of the Liassic formations. Part III. Ichthyosaurus. Palaeontogr. Soc. {Mono- graph). pp. 83-130, pis. 21-33. pearce, c. 1842. On mouths of ammonites and fossils contained in laminated beds of the Oxford Clay, discovered in cutting of G.W.R. near Christian Malford in Wiltshire. Proc. Geol. Soc. Loud. 3, 592-4. piveteau, j. 1953. Trade de paleontologie. Vol. 2. Paris. quenstedt, f. a. 1885. Handbuch der Petr efaktenkunde. Tubingen. romer, a. s. 1966. Vertebrate Paleontology. Chicago. 468 pp. seeley, h. g. 1877. Mauisaurus gardneri, an elasmosaurian from the base of the Gault at Folkestone. Q. Jl. Geol. Soc. Land. 33, 541-7, pi. 23. 1880. Report on the mode of reproduction of certain species of ichthyosaurs from the Lias of England and Wurtemburg. Rep. Br. Advmt. Sci. (1880), 68-76, pi. 1. shimanskiy, v. n. 1949. O sistematicheskom polozhenii rinkolitov. Trudy Paleont. Inst. 20, 199-208. tarlo, l. b. 1959. Pliosaurus brachyspondylus from the Kimmeridge Clay. Palaeontology, 1, 283-91, pis. 51, 52. williston, s. w. 1904. The stomach stones of the plesiosaurs. Science, 20, 565. 1914. Water reptiles of past and present. Chicago. 251 pp. wiman, c. 1946. Uber Ichthyosaurier und Wale. Senckenbergiana, 27, 1-11, pis. 1-3. woods, h. 1946. Palaeontology. Invertebrate. 8th edn. Cambridge. 477 pp. woodward, a. s. 1917. The so-called coprolites of ichthyosaurs and labrinthodonts. Geol. Mag. (6). 4, 540-2, pi. 34. woodward, h. b., and ussher, w. a. e. 1911. The geology of the country near Sidmouth and Lyme Regis. Mem. geol. Surv. U.K. wurstemberger, A. R. c. von. 1876. Ueber Lias Epsilon. Jh. Ver. vaterl. Naturk. Wurtt. 32, 193-358, pi. 4, figs. 1-16. zittel, k. a. von. 1913. Textbook of Palaeontology. 2nd English edn. London. JOHN E. POLLARD Department of Geology The University Manchester 13 Typescript received 19 December 1966 MANTLE CANAL PATTERNS IN SCHIZOPHO RIA (BRACHIOPODA) FROM THE LOWER CARBONIFEROUS OF NEW SOUTH WALES by JOHN ROBERTS Abstract. The brachiopod Schizophoria verulamensis Cvancara, from the Lower Carboniferous of the Gresford district, New South Wales, has distinctive mantle canal patterns in small, medium, and large specimens. The different patterns are interpreted as stages in the ontogeny and are thought to result mainly from the growth and enlargement of the genitalia. Changes in the morphology of the canals in pedicle valves throughout six horizons in the Lower Carboniferous sequence show a trend, probably genetically controlled, towards the earlier maturity of the genitalia. The super- imposition of larger or mature mantle canal systems on smaller or immature systems indicates that the canals are impressed by periodic resorption, and it is argued that this probably took place during the winter when there was a stable mantle canal system and little shell deposition. In large specimens there are connexions between the vascula genitalia, vascula media, and lateral canals in the pedicle valve, and between the vascula genitalia and vascula myaria in the brachial valve. It is suggested that the connexions may have increased the amount of oxygenated coelomic fluid circulating over the gonads, possibly causing them to ripen earlier. Pathological variations in the mantle canals are described, and it is shown that the canals are able to adjust to injuries or defects in the shell. The mantle canals in living brachiopods are extensions of the body cavity into the mantle. Their function is mainly respiratory and in articulates they also act as receptacles for the gonads. In both living and fossil brachiopods the pattern of canals is frequently impressed on the inner surface of the shell, particularly that of older specimens. In fossil brachiopods impressions of the mantle canals are common in orthoids and strophomenoids, are rarely seen in the spiriferoids and chonetoids, and only recently have been observed in two species of productoids from the Lower Carboniferous of the Bonaparte Gulf Basin, northwestern Australia. As a result their use in systematics is limited. Schuchert and Cooper (1932) described the morphology of the mantle canals in the orthoids and used the canal patterns as a basis for the separation of a new subfamily. Opik (1934) further demonstrated their use in his study of the Clitambonitidina, and proposed a detailed terminology for the various parts of the mantle canal system. Williams (1956) was the first to present a comprehensive review of the mantle canal patterns in articulate brachiopods. He modified Opik’s terminology and showed that the mantle canal patterns in articulates could be referred to standard types which were probably derived from Cambrian forms. Williams (1956) and Williams and Rowell (1965) illustrated many different types of mantle canal patterns. The present study deals with the morphology, development, and probable genetic changes in the mantle canals of the Lower Carboniferous orthoid Schizophoria verula- mensis Cvancara (1958). Extremely well-preserved material has made possible the recognition of a sequence of young, mature, and old stages in the mantle canal system of 5. verulamensis. These stages have in turn led to the recognition of a trend towards an earlier maturity of the reproductive system during the Lower Carboniferous, and a [Palaeontology, Vol. XI, Part 3, 1968, pp. 389-405, pis. 74-75.] 390 PALAEONTOLOGY, VOLUME 11 method of impression or superimposition of the mantle canals which had not pre- viously been recognised in the Brachiopoda. Because of the exceptionally fine preserva- tion it has also been possible to recognize an intricate interconnexion between the various parts of the mantle canal system, a feature not recognized in Williams’s and Rowell’s treatment of mantle canal patterns. Schizophoria verulamensis Cvancara is one of the most common brachiopods in the Lower Carboniferous (Tournaisian and Visean) of New South Wales, and has been collected from every known fossil horizon in the Gresford district, about 35 miles north-west of Newcastle (text-fig. 1). The majority of the specimens described in this paper are preserved in indurated grey siltstone; the rest occur in medium-grained lithic sandstone. They are exceptionally well preserved, and have very finely detailed impres- sions of the mantle canals and setal follicles. Internal moulds only are figured in the plates because the impressions of the canals, represented by ridges, are particularly prominent when illuminated by an oblique light source. Schizophoria verulamensis is especially well represented in collections from the following localities (with grid references): L. 53 Greenhills (46609790), L. 206 Lewins- J. ROBERTS: MANTLE CANAL PATTERNS IN SCHIZOPHORIA 391 brook syncline (46279968), L. 270 Trevallyn (45729864), L. 218 Gresford Quarry (45809912), L. 50 Gresford Quarry (45789913), L. 215 Lewinsbrook (46089882). With the exception of L. 53 Greenhills, which is on the Paterson One-Mile Sheet, the grid references are taken from the Dungog One-Mile Sheet. The stratigraphic positions 1 1 TlZ'one' [I_.l Sands'°n° |f Conglomerol, | Crinoidol limestone Oolitic limestone text-fig. 2. Diagrammatic stratigraphic sections of the marine rocks at Trevallyn and in the Lewins- brook Syncline. of the horizons are shown in text-fig. 2 on diagrammatic stratigraphic sections of the marine rocks in the north-western part of the Gresford district. All of the collections are housed in the Geology Department, University of New England, Armidale, N.S.W. The stratigraphy of the Gresford district has been described by Roberts (1961 and 1965c) and the faunas by Roberts (1963, 1964, and 1965n and b ), Campbell and Roberts (1964), and Brown, Campbell, and Roberts (1965). The specimens of Schizophoria verulamensis were collected from a single horizon at each fossil locality except L. 53, where two horizons were sampled. An attempt was 392 PALAEONTOLOGY, VOLUME 11 made to collect randomly at each locality, and all the specimens collected were examined in the laboratory. Width-frequency plots for pedicle and brachial valves are given in text-fig. 3. The samples are characterized by the almost total absence of small specimens, and the presence of a large number of disarticulated valves, features which suggest that the populations had been affected by current action and sorting prior to burial. 20 30 40 50 20 30 40 50 60 WIDTH OF VALVES IN MM • Pedicle valves * Brachial valves text-fig. 3. Width-frequency polygons for Schizophoria verulamensis Cvancara specimens from six localities in the Gresford district. The post-mortem modification is also borne out by the unequal numbers of pedicle and brachial valves and small number of articulated shells collected from each locality. These figures are given in Table 1. The collection from L. 270 is the only sample having a comparable number of pedicle and brachial valves and also contains the greatest number of articulated shells, suggesting that it is perhaps the closest to a life assemblage. Further evidence that most of the assemblages are death assemblages is provided by the occurrence, at L. 218, L. 270, and L. 206, of shell beds containing large numbers of J. ROBERTS: MANTLE CANAL PATTERNS IN SCHIZOPHORIA 393 mature specimens of Schizophoria to the virtual exclusion of smaller specimens and other small species. The L. 53 horizon contains several other large species as well as Schizophoria , but also has very few small species. Smaller species and polyzoan debris accompany the Schizophoria specimens at L. 215 and L. 50. TABLE 1 Disarticulated valves Locality Pedicle valve Brachial valve Articulated shells L. 53 25 6 — L. 206 28 18 1 L. 270 31 30 10 L. 218 30 23 1 L. 50 25 20 5 L. 215 54 31 7 MANTLE CANALS The mantle canals in living brachiopods are situated in the mesenchyme layer of the mantle and are lined with ciliated epithelium. They are filled with coelomic fluid, which is rapidly circulated by the ciliated epithelium (Hyman 1959, p. 593), the circulation in some species being facilitated by a median ridge dividing the in- and out-flowing currents. The coelomic fluid contains free coelomocytes, including a spherical type which has a respiratory function and may be equivalent to a red blood corpuscle (Hyman 1959, p. 557). The pigment in the ‘respiratory cells’ absorbs oxygen and on reduction releases it to oxygenate the tissues. The pattern of the mantle canals varies in different brachiopod groups (Williams 1956), and there is usually a different pattern in the pedicle and brachial valves. In most cases the canals originate as wide branches from the body cavity; some of these divide repeatedly until they extend to the proximal ends of the setal follicles around the edge of the mantle; other form sac-like receptacles and act as gonocoeles. Most living brachiopods have a complicated system of mantle canals extending throughout the mantle, providing a large surface area through which oxygenation of the coelomic fluid can take place. An adaptation to increase the surface area of the mantle canals is illustrated by Williams and Rowell (1965, fig. 24) in which the mantle canals in the inarticulate Glottidia subtend gill ampullae into the mantle cavity. In some fossil groups, however, the mantle canals have a relatively small surface area, and respiration may have been largely carried out by the lophophore. The lophophore in modern articulates has a plentiful supply of coelomic fluid provided by the coelomic canal and the small brachial canal (Williams and Rowell 1965, fig. 30). The small brachial canal gives off a branch to each of the filaments on the lophophore which have a large surface area and are ideally suited to respiration. Spiriferoids, for example, had narrow pinnate mantle canals with a small surface area combined with a large spiral lophophore. The lophophore probably had a large surface area suitable for oxygen exchange, was bathed in a stream of fresh water brought in by the moving filaments, and was an ideal organ for respiration as well as food gathering. Mantle canals appear to be poorly developed in the productoids, which probably also had a spiral lophophore d d C 5586 394 PALAEONTOLOGY, VOLUME 11 (Williams 1956, fig. 5 (6); Brunton 1966, figs. 8 and 9), and the lophophore may have acted as the main respiratory organ. Chonetoids occasionally exhibit mantle canals (Muir-Wood 1962, pi. 6, fig. 7), but they too probably used a spiral lophophore for respiration as well as for food gathering. The mantle canals are also the receptacles for the gonads in living articulate brachio- pods, and so the gonads are continuously bathed in oxygenated coelomic fluid. Blood vessels, which are distinct from the mantle canals, extend from the body cavity into the mantle canals and run along the margins nearest the mantle cavity (i.e. along the inner margins of the mantle canals). These vessels, which are only seen in living brachiopods, follow all the branches of the mantle canals and extend to the margin of the mantle, and also form anastomosing blood sinuses to the gonads (Hancock 1859, pi. 55, fig. 1; pi. 63, fig. 1; and pi. 56, fig. 4). The blood system in brachiopods is an open system, the blood returning to the body cavity by way of tissue spaces. The blood itself is generally free of cells (Hyman 1959, p. 558) and consists of a colourless lymph-like fluid. Because the blood seeps back to the coelomic cavity before being recirculated it is probably coelomic fluid from which the coelomocytes have been filtered off. The blood system is responsible for the supply of nutriment to the organs within the body cavity, the gonads, and to the mantle, especially to the cells at the grow- ing edge of the mantle. Waste products are probably transported by the blood as it filters through the tissues back to the coelome, where they are removed by the nephridia. The blood must also contain dissolved minerals, mainly calcium carbonate, for deposi- tion as shell material by the cells at the growing edge of the mantle. MANTLE CANALS IN SCHIZOPHORIA VERULAMENSIS The morphology of the mantle canal impressions in both valves of Schizophoria vendamensis is shown in text-fig. 4. In the following description of the mantle canal system I propose to use the term ‘canals’ for the impressions of the mantle canals on the inner surfaces of the valves so as to avoid another descriptive term. The mantle canals of small and apparently sexually immature specimens of Schizo- phoria vendamensis consist of small canals of vascula media and vascula myaria as well as a large number of narrower linear canals originating from the anterior and antero- lateral margins of the muscle field and extending to the margins of the valve (text-fig. 5 a , b). There is no differentiation of vascula genitalia. In the pedicle valve the impressions of the vascula media arise from the anterior margins of the diductor muscle scars and extend with only minor branching to the front of the valve. An arcuate lateral canal such as that in older individuals (text-fig. 5c) does not appear to have been developed and the lateral margins of the valve were presumably supplied with coelomic fluid by the narrow linear canals extending from the body cavity. Impressions of the mantle canals are less frequently preserved in brachial valves. Small brachial valves have canals of vascula media originating from between the inner margins of the anterior adductor muscle scars or, more rarely, from the inner parts of the adductor scars, and extending towards the front of the valve. The vascula myaria are simple, straight canals arising from between the anterior and posterior adductor muscle scars and extending towards the postero-lateral margins. J. ROBERTS: MANTLE CANAL PATTERNS IN SCHIZOPHO RIA 395 Adductor muscle scars Posterior adductor muscle scar PEDICLE VALVE BRACHIAL VALVE text-fig. 4. Morphology of the mantle canal impressions in Schizophoria verulamensis Cvancara. A text-fig. 5. Mantle canal patterns in Schizophoria verulamensis Cvancara showing the development from small to large individuals. The pedicle valve is on the left and the brachial valve on the right. 396 PALAEONTOLOGY, VOLUME 11 Larger, apparently sexually mature individuals (text-fig. 5 c, d) have an interconnected system of vascula genitalia in the pedicle valve (PI. 74, fig. 3), an incipient system of vascula genitalia in the brachial valve (PI. 74, fig. 6), and more complex vascula media and vascula myaria canals. In the pedicle valve of mature individuals the main canals of vascula media are well separated from one another and are much broader and more deeply impressed than in the younger form. Each of the main canals branches distally, giving off a number of canals which run to the front of the valve, and an arcuate lateral canal which extends parallel with the margin of the valve to the hinge. The lateral canal has a crenulate pattern and at the peak of each crenulation gives off a major branch, which usually sub- divides several times before reaching the margin of the valve. Towards the hinge the lateral canal is irregularly crenulate and terminates in a group of diverse branches. The vascula genitalia cover an area enclosed by the muscle field, the vascula media, and the inner margins of the arcuate lateral canals, and are connected by numerous branches with the vascula media and lateral canals; the significance of these connexions will be dealt with below. The vascula genitalia in brachial valves having the same size and presumably the same age as mature pedicle valves do not appear to be as strongly impressed as those in the pedicle valve. The genital markings consist of pit-like depressions or incipient inter- connected canals on the postero-median parts of the valve, and give rise to numerous branching canals, similar to those of the vascula media and vascula myaria, which extend to the lateral margins (PI. 74, fig. 6). There are up to ten canals of vascula media, and ten canals of vascula myaria extending with variable bifurcation towards the anterior margin. Both the vascula media and vascula myaria leave tracks on the anterior adductor muscle scars. In the largest and probably gerontic specimens (text-fig. 5 e, f ) the mantle canal system is frequently more strongly impressed on the inner surface of the shell and is characterized by even broader main canals and an expansion of the vascula genitalia. i EXPLANATION OF PLATE 74 Schizophoria verulamensis Cvancara Fig. 1. X 1. F.7145a. Internal mould of pedicle valve showing the connexions between the vascula media and vascula genitalia; L. 215 Lewinsbrook. Fig. 2. X 1. F.7145r. Internal mould of pedicle valve; note the truncation of the narrow canals by the main canals of vascula media; L. 270 Trevallyn. Fig. 3. X 2. F.4677a. Internal mould of pedicle valve showing a sexually mature mantle canal system superimposed over an immature system; L. 215 Lewinsbrook. Fig. 4. X 1. F.8036. Internal mould of pedicle valve showing a third canal of vascula media extending to the front of the valve; L. 50 Gresford Quarry. Fig. 5. x 1. F.7147q. Internal mould of brachial valve showing a sexually immature pattern of narrow linear mantle canals; L. 270 Trevallyn. Fig. 6. X 1. F.7141a. Internal mould of brachial valve showing the early development of the vascula genitalia; L. 215 Lewinsbrook. Fig. 7. xl. F.7153b. Internal mould of mature brachial valve showing the connexion between the vascula genitalia and the vascula myaria; L. 50 Gresford Quarry. Fig. 8. X 1. F.7153c. Internal mould of a mature brachial valve; note the connexion between the vascula genitalia and the vascula myaria; L. 50 Gresford Quarry. Fig. 9. x2. F.7153a. Internal mould of an old brachial valve; note the narrow canals emerging from beneath the wide superimposed canals of vascula media; L. 50 Gresford Quarry. Palaeontology, Vol. 11 PLATE 74 ROBERTS, Schizophoria J. ROBERTS: MANTLE CANAL PATTERNS IN SCHIZOPHORIA 397 In the pedicle valve the main vascula media canals are close together and branch distally, the arcuate lateral canals are closer to the margin than in younger specimens, and the vascula genitalia have the same ramifying canals as those in mature specimens, covering almost the entire inner surface of the valve (PI. 75, figs. 4-7). The canals of vascula media in the brachial valve are usually wider than in smaller specimens but do not appear to have significantly altered their positions. The vascula myaria are also broader and run straight to the antero-lateral margins, or are bent inwards at their distal extremities. Many small branches along the outer margins of the vascula myaria provide a connexion with the vascula genitalia. The genitalia have the same morphology as those in the pedicle valve. They cover broad flabel late areas on the lateral and postero-lateral parts of the valve, but have a smaller area than the genitalia on the pedicle valve because they are confined by the vascula myaria. A system of narrow linear canals at the margins of mature shells (PI. 74, figs. 1-3, 9; PI. 75, figs. 1, 2) are continuous with small branching canals in the middle part of the shell (viz. the small canals between the vascula media in PI. 74, fig. 3). These small branching canals are interpreted as impressions of a juvenile mantle canal system, and hence their distal parts are regarded as impressions of mantle canals, probably impres- sions of the termination of the canals at the proximal ends of the setal follicles. These impressions were apparently continuously recorded by a process of differential thicken- ing at the margin. The linear canals are joined by branches of a superimposed mantle canal system (see below), indicating their function as a canal rather than the trace of a setal follicle. Had they been merely traces of setal follicles they would intercalate as new setae were introduced around the margin; when they increase, these canals branch (PI. 74, fig. 3; PI. 75, fig. 3). Changes in the mantle canals. In the pedicle valve the main canals of vascula media extending in front of the muscle field can be divided into three intergrading morpho- logical types (text-fig. 6): A. Canals convex outwards and widely separated. B. Canals straight and widely separated. C. Canals straight and close together. It has been shown above that the A and B types of canals typify small- to medium- sized individuals and C type canals larger individuals, and hence the different shapes of the vascula media can be explained as stages in the ontogenetic development of the individual. Accompanying the change from the A to C types of canals is the enlargement of the complex network of the vascula genitalia, and it is suggested that this enlarge- ment within the mesenchyme layer of the mantle gradually forced the vascula media canals together and the lateral canals to move towards the margin as the shell became older. Specimens having each of these types of canals are present in all of the horizons in the Gresford district. This explanation of the development of the canals does not account for their distribu- tion as outlined in Table 2, unless evolution has taken place. An analysis of Table 2 is simplified if it is realized firstly that the C type canal is that seen in the largest specimens, and secondly that the L. 53 sample is dominated by large specimens (text-fig. 7). Table 2 shows a decrease in A type canals and an associated 398 PALAEONTOLOGY, VOLUME 11 increase in B type canals from the oldest to the youngest horizons. This trend could be explained in at least two ways: as the expression of the ontogeny, as outlined above, in which sampling errors gave the impression of a genetic change, or as a true genetic change towards the earlier maturity of the genitalia. An examination of the width-frequency polygons (text-fig. 3) shows that the samples are essentially the same in terms of size distribution, except for L. 53. Also, the specimens from the highest stratigraphic horizon with B type canals are about the same size as specimens from the lower horizons characterized by A type canals. In the absence of genetic change the specimens having the B type canals should have been larger. TABLE 2 A Canal types B C L. 53 Greenhills Occasional 30% 70% L. 206 Lewinsbrook Syncline 25% 75% occasional L. 270 Trevallyn 65% 35% occasional ^ j Gresford Quarry 65% 35% occasional L. 215 Lewinsbrook 70% 30% occasional More conclusive evidence in favour of the genetic change is available from younger horizons (probably middle to upper Visean) at Barrington, N.S.W., in which moderate to large specimens of S. verulamensis have mainly straight, sub-parallel (or B type) canals (Cvancara 1958). The earlier maturity of the genitalia would have increased the number of larvae pro- duced by each individual and probably significantly increased the S. verulamensis popula- tion to one of even greater dominance during the Visean. Superimposition of mantle canals. In mature specimens of Schizophoria verulamensis, particularly in pedicle valves, the greater part of the mantle canal system is super- imposed over small radial canals. The small canals originate from the body cavity frequently branch, and run irregularly towards the front of the shell. In mature valves the posterior parts of these canals (viz. the narrow branching canals between the main vascula media in PI. 74, fig. 3) are interpreted as the traces of juvenile mantle canals. EXPLANATION OF PLATE 75 Schizophoria verulamensis Cvancara Fig. 1. x2. F.7152a. Internal mould of pedicle valve showing narrow canals extending from the front of the muscle field and being truncated by the superimposed mature mantle canal system; L. 50 Gresford Quarry. Fig. 2. x2. F.7146k. Internal mould of pedicle valve; L. 270 Trevallyn. Fig. 3. x2. F.4677b. Internal mould of pedicle valve; L. 215 Lewinsbrook. Fig. 4. X 1. F.7144o. Internal mould of an old pedicle valve; L. 53 Greenhills. Fig. 5. X 1. F.4784. Internal mould of an old pedicle valve with C-type vascula media and large areas of vascula genitalia; L. 53 Greenhills. Fig. 6. X 1. F.7145b. Internal mould of an old pedicle valve; L. 270 Trevallyn. Fig. 7. xl. F.7146c. Internal mould of an old pedicle valve with large vascula genitalia, C-type vascula media and the lateral canals situated close to the margin; L. 270 Trevallyn. Palaeontology , Vol. 11 PLATE 75 ROBERTS, Schizophoria J. ROBERTS: MANTLE CANAL PATTERNS IN SCHIZOPHORIA 399 Because all the canals were connected to the bases of the setal follicles (Williams and Rowell 1965) these juvenile canals are continuous with the traces, recorded at the margin, of the termination of the mantle canals at the setal follicles. text-fig. 6. The three types of vascula media canals in the pedicle valve of Schizophoria verulamensis Cvancara. so r 5 UJ 40 - > _l < > • L53 o L206 + L270 x L2I5 O Q LU CL I o z LU 20 - o o • ® t* x + f» • * liVt0 x x#+x x x °x x n 4° <»XX+* o X +» 4- + + , *4** x xx o X . ° 8 o o o t n —I 1 I 30 40 50 WIDTH PEDICLE VALVE MM 60 text-fig. 7. Scatter diagram of length plotted against width for pedicle valves of Schizophoria verula- mensis Cvancara from L. 215 Lewinsbrook, L. 270 Trevallyn, L. 206 Lewinsbrook Syncline, and L. 53 Greenhills. In mature pedicle valves impressions of the small canals extend from the front of the muscle field between the large canals of vascula media, branching as they run towards the front of the valve (PI. 74, fig. 3; PI. 75, figs. 1, 2). In the middle and mid-anterior 400 PALAEONTOLOGY, VOLUME 11 parts of the valve they are truncated by superimposed larger canals: viz. the inner branch of the right-hand vasculum medium on the specimen in Plate 74, fig. 3 in which small canals extend on either side of a branch of the superimposed larger canal. Near the margins, the large canals join with the narrow linear traces left by the mantle canals at the ends of the setal follicles, and extend to the depressions left by the setal follicles. The junction between these canals results from both systems supplying the same setal follicles at the margin; the small linear marks were impressed as the shell grew, and the larger canals, which had been formed by the modification of the smaller ones, were impressed at a later time. The superimposition of the mantle canal systems is less clear in the brachial valve. One large specimen, however, has a system of narrow linear canals beneath broad canals of vascula media (PI. 74, fig. 9). The U-shaped branch on the left hand side of the vascula media truncates at least six of the narrow canals, and other narrow canals emerge from beneath the sites of bifurcation of the larger canals and are apparently continuous with the traces of the canals at the bases of the setal follicles. The two main systems (i.e. the younger and older systems) described above are inter- preted as impressions of the mantle canals at two different periods during the life of the shell, probable before and after sexual maturity. Because of the impression of a large or mature system on top of a smaller and probably immature system it is inferred that the major part of the canal system was impressed at certain well-spaced intervals instead of being continuously recorded on the inside of the shell. Williams (1956, p. 273) thought that mantle canals were impressed on the inner surface of the shell by differential secretion of shell material by the outer epithelium covering the canals. In S. verulamensis this mechanism may have been responsible for the continuous impression of the distal parts of the canals at the margin, but it could not have been responsible for the impression of the main part of the mantle canal system, because there are two and sometimes three systems superimposed over one another, unless rapid differential secretion took place at specific times; this is unlikely because it means that rather than being gradually thickened the shell would have been periodically rapidly thickened. A more convincing explanation is that the outer epithelium over the mantle canals resorbed the shell material at specific times. Resorption is most likely to have taken place during the periods of limited shell growth when the mantle canal system had a stable morphology. During periods of rapid growth there would have been continuous reorganization of the mantle canals at the mantle edge and the development of addi- tional setal follicles at the growing margin. To maintain efficient circulation to the margin the main canals of vascula media and vascula myaria would have enlarged by reorganizing the connective tissue in the mantle and coalescing with smaller canals. Many modern shallow-water marine organisms, particularly molluscs, cease depositing shell during the colder months of the year (Epstein and Lowenstam 1954), and hence it is reasonable to assume that the winter was the most likely time for the retardation of shell deposition and the impression of the mantle canals in 5. verulamensis. Abnormal mantle canals. In the pedicle valve figured in Plate 74, fig. 4, the left-hand canal of vascula media divides a short distance in front of the muscle field, giving a third main canal extending to the median anterior part of the valve. All three canals J. ROBERTS: MANTLE CANAL PATTERNS IN SCHIZOPHORIA 401 branch in a similar manner to those in other valves. Functionally this additional large branch may have provided a greater supply of coelomic fluid to the growing edge of the front of the mantle. The two outer canals of vascula media give off arcuate lateral canals. Another specimen showing an abnormal development of the vascula media (text-fig. 8) has impressions of the two main canals of vascula media arising from the anterior margin of the left diductor muscle scar and running, parallel with one another, to the antero-lateral part of the valve. There is no canal originating from near the right diductor scar, presumably because of some injury or functional defect to that side of the body cavity, and the vascula genitalia extend in front of the muscle scar. The text-fig. 8. Pathological pedicle valve of Schizophoria verala- mensis Cvancara showing the two main canals of vascula media extending from the left diductor muscle scar, and the vascula genitalia extending in front of the right diductor muscle scar. preservation is too poor to trace the junction of the right-hand canal with the right lateral canal. This variation indicates a certain adaptability in the organization of the mesenchyme layer of the mantle, and supports the arguments put forward above that major reorganizations took place within the mantle during the development of the species. Other features probably associated with this deformity are a deep furrow on the antero-median part of the left diductor muscle scar and an exceptionally narrow median ridge which does not expand anteriorly. In a large pedicle valve (PI. 75, fig. 6) branches from the arcuate lateral canal in the posterior part of the valve, instead of extending normal to the margin of the valve, turn posteriorly and run almost parallel with the margins, diverging only slightly from the main lateral canal. In other valves (PI. 74 ,fig. 3 jonly the posterior-most branches of the lateral canal turn and run more or less parallel with the canal. Connexion between the mantle canals. The description and figures of mantle canals of fossil brachiopods given by Williams (1956) and Williams and Rowell (1965) suggest that each of the three pairs of extensions of the coelomic cavity into the mantle — the vascula media, vascula myaria, and vascula genitalia — is a discrete unit, unconnected with other canals. 402 PALAEONTOLOGY, VOLUME 11 Larger specimens of Schizophoria verulamensis with well-preserved mantle canals have vascula media and arcuate lateral canals possessing well defined inner and outer margins. The vascula genitalia contained within the arc of the lateral canals is connected by many fine branches with the main vascula media canals and the arcuate lateral canals, the connexions varying from narrow constricted branches (the branches on the right-hand vasculum medium of the specimen in PI. 74, fig. 3) to broad canals (the connexion with the right-hand lateral canal on the same specimen). In another specimen (PI. 74, fig. 1) the branches into the vascula media canals are more pronounced and have a herringbone pattern. As well as being connected with the vascula media, the vascula genitalia are presumably connected directly with the coelomic cavity adjacent to the muscle-field. There is no connexion with the vascula media in the brachial valve, but instead the vascula genitalia are linked with the main canals of vascula myaria (PI. 74, figs. 7, 8). The branches are comparable with those in the pedicle valve and appear to extend along the entire outer margin of the vascula myaria canals. Similar connexions occur in Ordovician orthid brachiopods described by Cooper (1956) (see ChciulistomeUa magna (Schuchert and Cooper), pi. 70, fig. 21; and Mimella globosa (Willard), pi. 89, figs. 10, 12, 14). One other possible example of an intercon- nected system in Schizophoria is given by Pocock (1966, fig. 17), who figured a specimen of S. striatula (Schlotheim) in which lateral branches arise from the vascula myaria and extend towards the vascula genitalia. This type of interconnexion between the genitalia and the other mantle canals is apparently unknown in modern brachiopods. The interiors of modern brachiopods figured by Hancock (1859) show that the genitalia are either situated in the mid-part of a sac-like mantle canal, such as that in Terebratulina caputserpentis Linne (Hancock 1859, pi. 53, figs. 1, 2), or lie within the narrow canals of vascula genitalia or vascula media as in Macandrevia cranium (Muller) (Hancock 1859, pi. 53, fig. 4). In specimens with sac-like mantle canals, closest to those in Schizophoria , the vascula media and lateral canals are simply the lateral extremities of the gonocoel; they have no inner walls, and hence differ from the discrete canals in S. verulamensis. There are no con- nexions with the network of genitalia, but the genitalia are bathed in coelomic fluid. In S. verulamensis , coelomic fluid must have actually passed along the tubular network of vascula genitalia. The connexion of the vascula genitalia with other canals in S. verulamensis probably increased the circulation of oxygenated fluid over the gonads, the fluid coming from the vascula media and lateral canals (or the vascula myaria in the brachial valve) as well as from within the vascula genitalia itself. The gonads in brachiopods without intercon- nected canals are oxygenated only by the coelomic fluid contained within the vascula genitalia, and sometimes the vascula media, meaning that there is probably less oxy- genated fluid passing over them. Additional aeration could be responsible for the forma- tion of larger gonads and possibly for a faster rate of production of sex cells, and it may also cause them to ripen earlier. In the modern brachiopod Terebratulina caputserpentis, already mentioned above, the gonads are bathed in the coelomic fluid of a large sac-like gonocoel, which is probably an even more efficient method of aeration. The interconnexion of the canals seems to be a favourable attribute, and on a purely morphological basis it would seem that a system, such as that in S. verulamensis, would J. ROBERTS: MANTLE CANAL PATTERNS IN SCHIZOPHORIA 403 have been more efficient for respiration and reproduction than, for example, a more restricted, discrete system such as that in the recent brachiopod Fal/ax (Williams and Rowell 1965, fig. 23d). Some of the respiratory functions of the mantle could, however, have been taken over by the lophophore in other brachiopod groups, so that there was less need to develop an interconnected system and the mantle canals acted as repositories for the gonads and as an accessory respiratory system. The present work shows that a closer examination of living brachiopods is desirable to see if any species have an interconnected mantle canal system like that in S. verula- mensis and other orthid brachiopods. Interpretation of the mantle canals. The development of the mantle canals in S. verula- mensis can be interpreted as follows. During the first period of growth, probably about one year, the shell was sexually immature and had a relatively undifferentiated system of mantle canals. In the first winter, when there was little shell growth, narrow linear canals including vascula media and vascula myaria were impressed on the inner surface of the shell, probably by resorption (PI. 74, fig. 5). During the following summer the shell grew rapidly and became sexually mature. The rapid growth and the introduction of many setal follicles at the mantle edge, combined with the growth of the gonads, resulted in the reorganization of the mantle canal system. In the pedicle valve the vascula media became wider and developed complex lateral canals, and the vascula genitalia changed from narrow linear canals to a network of ramifying channels which branched with one another and with the main branch and lateral canal of the vascula media (PI. 74, fig. 3). The changes in the brachial valve were not as marked; the canals of vascula media and vascula myaria becoming wider, and vascula genitalia becoming pitted near the postero- median part of the valve and also having wider canals extending to the lateral margins (PI. 74, fig. 6). These canals were impressed during the second winter, when the shell was about two years old. The changes in mature individuals, especially the development of the lateral canals and the network of vascula genitalia in the pedicle valve, must have resulted in a com- plete reorganization of the connective tissue in the mantle as well as the introduction of many blood vessels to the new canals and the genitalia. According to Hyman (1959, p. 533) the connective tissue of living brachiopods consists of a ‘homogeneous material containing stellate mesenchymal cells that may unite into loose networks’; the mesen- chyme was thus probably easily modified to allow the introduction of new or larger canals. Further growth in the following summer resulted in an increase in size, the enlarge- ment of the muscle-field, and a thicker shell, especially at the posterior. In the pedicle valve the vascula genitalia increased enormously in size, forcing the main vascula media canals close together (PI. 75, figs. 4-7) and the lateral canals towards the margin, so that the impressions of the mantle canals covered almost the whole inside of the valve. In the brachial valve the vascula media and vascula myaria canals became wider and the vascula genitalia developed a ramifying network similar to that in the pedicle valve (PI. 74, figs. 7, 8, 9). These enlarged canals were probably impressed during the third winter. Little is known about the rate of growth and maturity of modern brachiopods. Rudwick (1962, p. 334; 1965, p. H209) suggested that modern articulate brachiopods 404 PALAEONTOLOGY, VOLUME 11 can reach sexual maturity after two or three years, and that they live for as long as seven or eight years or even longer. Amongst the inarticulates Lingula is sexually mature after a year or a year and a half (Chuang 1959), and Glottidia after only six to nine months (Paine 1963). It is therefore reasonable to suggest that S. verulamensis became sexually mature at about two years old, and that the mature mantle canal system was impressed during its second winter. Mantle canals of older individuals cover almost the entire surface of the shell and leave no room for the expansion of the vascula genitalia, so that after about the third year the mantle canals became stable and could not be modified during further growth of the shell. There is no indication of how long the largest shell lived. SUMMARY AND CONCLUSIONS The morphology of the mantle canals in Schizophoria verulamensis Cvancara changes during ontogeny, mainly because of the growth of the gonads. The canals start in immature specimens as narrow linear canals over the inner surface of the valves and, with the growth of the gonads, are later modified in larger specimens into several major feeder canals which distribute coelomic fluid from the body cavity to the margins of the shell, around the areas of genitalia. In the largest shells the genitalia cover the greater part of the inner surface of the valves and have enlarged to such an extent that the large distributive canals have been forced together in the middle of the valve, or moved outwards towards the margins. The pedicle valve is affected earlier by the growth of the genitalia than the brachial valve. An analysis of the distortion to the main vascula media caused by the enlargement of the vascula genitalia in pedicle valves from six horizons throughout the Lower Carboni- ferous sequence indicates a trend, probably genetically controlled, towards the earlier maturity of the genitalia. The main part of the mantle canal system was impressed periodically by the resorption of shell beneath the mantle canals rather than by differential secretion as suggested by Williams (1956). Resorption most likely took place during the winter when there was limited shell growth and a stable mantle canal system, recording several distinct mantle canal patterns during the development of the individual. Larger specimens of S. verulamensis have the vascula genitalia connected by numerous small branches with the vascula media and lateral canals in the pedicle valve and the vascula myaria in the brachial valve. This type of connexion has not been observed in recent brachiopods. It is thought to have increased the circulation of oxygenated coelomic fluid over the gonads, and may have been responsible for their earlier ripening. The mantle canals in pathological specimens indicate that not only was modification possible during ontogeny, but the mantle canal system could also be adapted to the effects of injuries or defects in the shell during life. Acknowledgements. This work was begun at the University of Western Australia and finished at the Bureau of Mineral Resources. I am indebted to Drs. K. S. W. Campbell and P. J. Coleman for their interested discussion of the problem, and to Drs. J. J. Veevers, J. M. Dickins, and K. S. W. Campbell for reading the manuscript. This paper is published with the permission of the Director, Bureau of Mineral Resources, Geology, and Geophysics. J. ROBERTS: MANTLE CANAL PATTERNS IN SCHIZOPHORIA 405 REFERENCES brown, d. a., Campbell, k. s. w., and Roberts, j. 1965. A Visean cephalopod fauna from New South Wales. Palaeontology, 7, 682-92. brunton, c. h. c. 1966. Silicified productoids from the Visean of County Fermanagh. Bull. Brit. Mas. (Nat. Hist.) Geol. 12, 175-243. Campbell, K. s. w., and Roberts, j. 1964. Two species of Delepinea from New South Wales. Palaeon- tology, 7, 514-24. chuang, s. h. 1959. The breeding season of the brachiopod Lingula unguis (L). Biol. Bull. 117, 202-7. cooper, G. a. 1956. Chazyan and related brachiopods. Smithson, misc. Coll. 127, 1-1245. cvancara, a. m. 1958. Invertebrate fossils from the Lower Carboniferous of New South Wales. J. Paleont. 32, 846-88. epstein, s., and lowenstam, h. a. 1953. Temperature-shell-grown relations of Recent and interglacial Pleistocene shoal-water biota from Bermuda. J. Geol. 61, 424-38. Hancock, a. 1859. On the organization of the Brachiopoda. Phil. Trans. Roy. Soc. Lond. 148, 791-869. hyman, l. H. 1959. The Invertebrates: Smaller coelomate groups. 5, 1-783. McGraw-Hill, New York. muir-wood, H. M. 1962. On the morphology and classification of the brachiopod Suborder Chonetoidea. Brit. Mus. (Nat. Hist.), London, 1-132. opik, a. a. 1934. Uber Klitamboniten. Acta et Commentationes, Univ. Tartu, 26, 1-239. Paine, R. T. 1963. Ecology of the brachiopod Glottidia pyramidata. Ecological Monographs, 33, 187-213. pocock, y. p. 1966. Devonian Schizophoriid brachiopods from Western Europe. Palaeontology , 9, 381-412. Roberts, j. 1961. The geology of the Gresford district, N.S.W. /. Proc. roy. Soc. N.S.W. 95, 77-91. 1963. A Lower Carboniferous fauna from Lewinsbrook, New South Wales. Ibid. 97, 1-31. 1964. Lower Carboniferous brachiopods from Greenhills, New South Wales. J. geol. Soc. Aust. 11, 173-94. 1965u. Lower Carboniferous faunas from Wiragulla and Dungog, New South Wales. J. Proc. roy. Soc. N.S.W. 97, 193-215. 19656. A Lower Carboniferous fauna from Trevallyn, New South Wales. Palaeontology, 8, 54-81 . 1965c. Lower Carboniferous zones and correlations based on faunas from the Gresford-Dungog district. New South Wales. J. geol. Soc. Aust. 12, 105-22. rudwick, m. j. s. 1962. Filter-feeding mechanism in some brachiopods from New Zealand. J. Linn. Soc. Lond. Zoology, 44, No. 300, 592-615. 1965. In Treatise on Invertebrate Paleontology. Part H, Brachiopoda. Ed. R. C. Moore. Univ. Kansas Press. schuchert, c., and cooper, g. a. 1932. Brachiopod genera of the Suborders Orthoidea and Pentame- roidea. Mem. Peabody Mus. Nat. Hist. 4(1) 1-270. williams, A. 1956. The calcareous shell of the Brachiopoda and its importance to their classification. Biol. Rev. 31, 243-87. and rowell, a. j. 1965. In Treatise on Invertebrate Paleontology. Part H, Brachiopoda. Ed. R. C. Moore. Univ. Kansas Press. JOHN ROBERTS Bureau of Mineral Resources, Geology and Geophysics Canberra City, A.C.T. Typescript received from author 1 May 1967 Australia ON ‘ DENDROCRINUS ’ CAM BRIENSIS HICKS, THE EARLIEST KNOWN CRINOID by DENIS E. B. BATES Abstract. Dendrocrinus cambriensis is re-described from new material and referred to Ramseyocrinus gen. nov., and placed in the family Eustenocrinidae. The crinoid remains described by Hicks from the lower Ordovician rocks of Ramsey Island and the adjacent mainland of Wales are the earliest yet reported, and, as such, might be expected to yield significant information on the origin of the class. Several writers (Moore 1950, p. 32, Ubaghs 1953, p. 744, Regnell 1960, p. 172) have referred to their early stratigraphical occurrence, and Ramsbottom (1960, p. 5) has re-investigated Hicks’s specimens, without being able to add significantly to Hicks’s information. Unfortunately none of Hicks’s specimens shows the posterior side, and hence the anal tube was unknown. The writer has had the opportunity of examining the extensive collection of material from Ramsey Island made by Dr. F. J. North for the National Museum of Wales, kindly made available by Dr. D. A. Bassett, and has based his description on specimens from it. Class crinoidea J. S. Miller 1824 Subclass inadunata Wachsmuth and Springer 1881 Order disparata Moore and Laudon 1943 Family eustenocrinidae Ulrich 1924 Monocyclic crinoids with a slender cylindrical crown about equal in diameter to the stem; calyx composed of three (?) to five basals, four radials, and an equal-sized infer- radianal; four arms, branching isotomously; infer-radianal bearing a series of un- branched quadrate plates presumably supporting an anal sac. Genus Ramseyocrinus gen. nov. Diagnosis. A genus of the Eustenocrinidae with a four-lobed stem composed of colum- nals of irregular thickness; basals probably three in number, separated by well-developed sutures. Type species. Dendrocrinus cambriensis Hicks 1873. Discussion. The described species of Eustenocrinus have a round smooth column, with two alternating sets of thin columnals, one set twice as thick as the other. The five basal plates are all imperfectly fused one to another. In Ramseyocrinus there are, apparently, fewer basals, and the sutures between them are as well defined as those between the radials. [Palaeontology, Yol. 11, Part 3, 1968, pp. 406-9, pi. 76.] D. E. B. BATES: ON ‘ D EN D ROCRI N US' CAMBRIENSIS HICKS 407 Ramseyocrinus cambriensis (Hicks) Plate 76, figs. 1-5 1873 Dendrocrinus cambriensis Hicks, p. 50, pi. 4, figs. 17-20. 1960 Jocrinus ? cambriensis (Hicks) Ramsbottom, pp. 5-6, pi. 3, figs. 9-1 1. Diagnosis. Basals low, over twice as wide as high; rectangular radials wider than high, and highly convex; nine or more primibrachs, arms branching at least three times. Type specimens. The syntypes (Ramsbottom 1960, p. 5) are preserved in the Sedgwick Museum, with some gutta-percha impressions in the Geological Survey Museum. Ramsbottom’s description is mainly based on a specimen preserved in the British Museum (BMNH E3). The author has also discovered in the Sedgwick Museum the counterpart mould of one of Hicks’s syntypes, and the specimen now bears the numbers SM 16739a-b. The description below is mainly based on a specimen from the National Museum of Wales, preserved as counterpart moulds (29 308 G296 and G318). Type locality. All the specimens figured by Hicks come from the Porth Gain Beds of Bay Ogof Hen, Ramsey Island (Pringle 1930, p. 12), although specimens are recorded from near Llanveran Farm, on the mainland near Pen Berry. No specimens have been seen or found by the writer from the latter locality. Description. Cup slightly wider than high, sides flaring slightly upwards, surface of plates smooth. Basals apparently not more than four in number, and probably only three, just less than half the height of the calyx and over twice as wide as high, penta- gonal, with the upper surfaces concave, and the lower surfaces sharply bent inwards between the petals of the stem. Radials four in number, quadrangular, wider than high, with convex lower and concave upper surfaces, highly convex, especially towards their lateral margins. Anal series commencing with an infer-radianal similar in width and convexity to the radials, but only half their height, the succeeding three plates of the anal series similar in size and height. The radials and first primibrachials almost equal in height to the infer-radianal and the succeeding two plates of the anal series. Lower- most brachials not interlocking and probably not included in the cup. Primibrachs nine to twelve, wider than high, almost circular in cross-section, with a deep narrow ventral groove, secundibrachs four to nine, as high as wide; tertibrachs seven or more; quadribrachs at least eight; arms branching isotomously at least three times. Stem wide, petaloid, four lobed, the lobes proximally almost parallel-sided with deep indentations, distally becoming less marked so that the stem becomes rectangular in cross-section; columnals of irregular height. Discussion. Moore and Laudon (1943, p. 26) described the two lowermost plates in each ray of Eustenocrinus as infer-radial and super-radial, homologizing them with the com- pound radials of genera such as Ectenocrinus (op. cit., fig. 3). However, in a later paper (1950, p. 30) Moore has proposed a simpler nomenclature, regarding the lowermost plate in each ray (excepting that beneath the anal tube) as radial, and the plates above and incorporated in the cup as fixed brachials. The latter procedure is followed here. The chief differences from the described species of Eustenocrinus are in the form of the stem and the number of basals, which warrant separation of the Welsh species into a new genus. There are at least three basals, one of which, on the anterior side, may be formed 408 PALAEONTOLOGY, VOLUME 11 from two plates fused together and showing traces of the suture between them (PI. 76, fig. 1). In addition E. springeri (Ulrich 1924, p. 99) has arms which bifurcate only once, and its diameter at the top of the basals appears to be less than at either the base or top of the cup (op. cit., fig. 14a). E. milled (Wetherby 1880, p. 153) has the radial plates higher than wide, with a thickening of their upper extremity which gives the species a somewhat swollen appearance at the top of the calyx. Eustenocrinus was considered by Moore and Laudon (1943, p. 25) as the most primitive member of the homo-synbathrocrinids, the major stock of the Disparata. Stratigraphically, however, it post-dates the earliest members of that stock, and hence is not itself the ancestor. The features considered by Moore to be primitive include the steep-sided conical cup, the convex base showing the lowest plates, isotomous arm branching, and the lack of any fused super and infer-radial plates. The occurrence of Ramseyocrinus at the base of the Ordovician, and its similar structure, further strengthens Moore’s claim for this stock to be the ancestor of the Disparata. The most striking features of Ramseyocrinus are the four-lobed stem and the four arms. It was hoped that Ramseyocrinus would prove to be a possible ancestor for the whole class Crinoidea, but these features indicate that it post-dates the splitting up of the ancestral crinoid stock into at least the sub-classes Inadunata and Flexibilia. Most other contemporary or older pelmatozoans have a well-developed five-fold symmetry, and thus it is likely that the Eustenocrinidae arose from an ancestor having five arms, one arm being replaced by an anal tube. Thus the Eustenocrinidae must first have lost the right posterior arm during the late Cambrian and then regained it during the early Ordovician, if they were, in fact, ancestral to the bulk of the Disparata. That this hap- pened is questionable; it seems more probable that the Eustenocrinidae constitute an early simplified offshoot from the Disparata. REFERENCES hicks, h. 1873. On the Tremadoc Rocks in the neighbourhood of St. David’s, South Wales, and their fossil contents. Quart. J. geol. Soc. Loud. 29, 39-52, pis. 3-5. moore, R. c. 1950. Evolution of the Crinoidea in relation to major Paleogeographic changes in Earth History. Int. Geol. Congress, 18th Session, London, 1948, pt. XII, pp. 27-53. and laudon, l. r. 1943. Evolution and Classification of Paleozoic Crinoids. Spec. Pap. geol. Soc. Amer. 46, 1-153, pis. 1-14. pringle, j. 1930. The geology of Ramsey Island (Pembrokeshire). Proc. Geol. Ass. Lond. 41, 1-31, pis. 1-3. ramsbottom, w. h. c. 1961. A monograph of British Ordovician Crinoidea. Palaeontogr. Soc. (Monogr). 1-37, pis. 1-8. EXPLANATION OF PLATE 76 Ramseyocrinus cambriensis (Hicks) from the Porth Gain Beds of Bay Ogof Hen, Ramsey Island. All photographs are of latex casts, whitened with ammonium chloride. None has been retouched. Figs. 1-3. Specimen 29 308 G296/G318 (Nat. Museum of Wales). 1. Anterior view, showing stem of irregular columnals, and two (?) basal plates partially fused together, X 5-8. 2. Posterior view, showing the anal tube, two of the four (?) basals, and food grooves on two of the arms, X 5-5. 3. Stem viewed from below, showing the four lobed cross-section, x4-5. Figs. 4-5. Syntype, specimen SM. 16739a-b (Sedgwick Museum, Cambridge). 4. Posterior (?) view, x 2-0. 5. Anterior (?) view, X 31. The basal and radial plates cannot be positively identified, nor can the anal series, which may be missing. Palaeontology, Vol. II PLATE 76 BATES, Ramseyocrinus cambrensis (Hicks) 1 D. E. B. BATES: ON ‘ D END RO C RIN US' CAMBRIENSIS HICKS 409 regnell, G. 1960. The lower Palaeozoic Echinoderm Faunas of the British Isles and Balto-Scandia. Palaeontology, 2, 161-79. ubaghs, g. 1953. Classe des Crinoides in Trade de Paleontologie, 3, 658-773. Ed. J. Piveteau. Paris. ulrich, e. o. 1924. New classification of the Heterocrinidae in Foerste, A. F., Mem.geol. Surv. Can. 138, 82-104. wetherby, a. g. 1880. Remarks on the Trenton Limestone of Kentucky with descriptions of new fossils from that formation and the Kaskaskia (Chester) group, Sub-carboniferous. /. Cincinnati Soc. Nat. Hist. 3, 144-60, pi. 5. DENIS E. B. BATES Department of Geology University College of Wales Typescript received 4 February 1967 Aberystwyth E e C 5586 REVISION OF TWO UPPER CAMBRIAN TRILOBITES by A. W. A. RUSHTON Abstract. Conocoryphe ? bucephala Belt, 1868, from the Upper Ffestiniog and Lower Dolgelly Beds of Wales, and possibly England, is re-illustrated; the pygidium is described for the first time; the species is transferred from the Olenid genus Beltella to which it was referred by Lake (1919) to Parabolinoides Frederickson, a genus widely known in the approximately contemporaneous Conaspis Zone in the U.S.A. Sphaerophthalmus major Lake, 1913, though regarded by Henningsmoen (1957) as of questionable validity, is distinguishable from other species of Sphaerophthalmus ; S. major occurs in England, Wales, and Sweden, and probably in Norway and eastern Canada, in the Zone of Peltura scarabaeoides. In his monograph on the Olenidae, Henningsmoen (1957) commented on the two Upper Cambrian trilobite species Conocoryphe ? bucephala Belt and Sphaerophthalmus major Lake, but for lack of data was unable to frame a diagnosis for either. In connexion with work on the Cambrian faunas from the Institute of Geological Sciences Merevale No. 1 borehole (Rushton, 1966), 1 examined the type-material of both these species including material from the following museums: British Museum (Natural History) (BM), Geological Survey Museum (GSM), Oxford University Museum (Ox), and the Sedgwick Museum, Cambridge (SM). The terminology used in the description below is that of Henningsmoen (1957). Family parabolinoididae Lochman Genus parabolinoides Frederickson 1949 Type species (by original designation). Parabolinoides contractus Frederickson 1949. Parabolinoides bucephalus (Belt 1868) Text-fig. 1 ; Plate 77, figs. 1-10, 1 1 ? 1868 Conocoryphe ? bucephala Belt, p. 10, pi. 2, figs. 1-6. 1873 Conocoryphe Willicunsoni Salter, p. 12 [Not Conocoryphe Williamsonii Belt, 1868], 1919 Beltella bucephala (Belt); Lake, p. 106, pi. 12, figs. 1 1-15. 1919 Beltella verisimilis (Salter) [part im]; Lake, p. 107, pi. 13, figs. 4, 5 only. ? 1927 Beltella ? sp. nov. Cobbold, p. 557. ? 1930 Beltella cf. bucephala (Belt); Stubblefield, p. 57. 1957 Olenusl bucephalus (Belt 1868); Henningsmoen, p. Ill (with further synonymy). Lectotype (selected by Henningsmoen 1957, p. 111). The original of Belt’s plate 2, fig. 1. The only specimen extant which can be considered as this original is BM I 7578 from the Upper Ffestiniog Beds near Dolgelly, shown here on Plate 77, fig. 7 ; it is the only relatively large and complete specimen in Belt’s collection, is the same size as his figure, and also agrees in being rather obscure in the posterior part of the shield. It is assumed that the figure is restored. Remarks on the material studied. The specimens of Conocoryphe ? bucephala collected by Belt and now in the British Museum (Natural History) and the Geological Survey [Palaeontology, Vol. II, Part 3, 1968, pp. 410-20, pis. 77-78.] RUSHTON: REVISION OF TWO UPPER CAMBRIAN TRILOBITES 411 Museum show the effects of flattening, crushing, and deformation to varying extents: the lectotype is obliquely compressed and the preglabellar field is crushed on to the anterior extension of the doublure of the slightly displaced free cheek; the cranidia in Plate 77, figs. 3 and 4, are laterally and frontally compressed respectively, and that shown in fig. 1 is crushed flat but seems to be relatively undistorted. The glabella of the type specimen of Conocoryphe williamsoni Salter non Belt (PI. 77, figs. 10a and b) is crushed on to the hypostoma, and the preglabellar field on to part of the free cheek, as in the lectotype of Parabolinoides bucephalus ; the free cheeks are displaced backwards and inwards relative to the cranidium and the pygidium is steeply inclined relative to the thorax, but otherwise the specimen is good, being but little flattened and deformed. In an attempt to determine the original proportions of P. bucephalus the method suggested by Henningsmoen (1960, p. 207) was applied to two mature cranidia on one of Belt’s slabs of Upper Ffestiniog Beds (GSM Za 4843) which satisfy the necessary con- ditions except that one is somewhat larger than the other. The calculation indicates that the original ratio of length to breadth of the cephalic axis was 4:3, which is in good agreement (less than 5 per cent, difference) with the ratios obtained from the apparently undeformed specimens shown in Plate 77, figs. 1, 6, and 10a; in contrast, direct measure- ment of the obviously deformed specimens in figs. 3 and 4 gives ratios of 2-0:1 and 0-9 : 1 respectively. Description. Lake’s description of the species is mainly correct but it requires some amplification, particularly in the account of the cranidium. Cephalic axis prominent, tapered, rounded in front, breadth three-quarters of length; SI weak or distinct, long, oblique backwards, straight or slightly convex; S2 weak, short, less oblique; a few specimens show traces of S3, as stated by Belt (1868, p. 10); the appearance of the glabellar furrows varies much according to the state of pre- servation. Occipital furrow deepest medially, composite; occipital ring apparently without node or spine. Frontal area one-third to nearly half as long as cephalic axis; border-furrow slightly bowed forwards and, in some specimens, with a weak median in-bend; anterior border crescentic or subtriangular, longer at the mid-line than the preglabellar field. Eye-ridges distinct or weak. Eyes opposite S2 or L2. Inter-ocular cheeks half or less than half as wide as the glabella at the eye-line; post-ocular cheeks rounded postero-laterally, slightly narrower than the occipital ring. Pre-ocular facial sutures straight, divergent forwards to the border-furrow, strongly convergent over anterior border; post-ocular facial sutures divergent, slightly convex. Free cheek with broad border and long anterior extension of the doublure (PI. 77, fig. 8). Ventral sutures unknown. There are twelve thoracic segments in the type-specimen of Conocoryphe williamsoni Salter and twelve or thirteen in the lectotype of C. ? bucepha/a, in which the joint between the thorax and pygidium is not very clear. The pleural tips are pointed and the axis lacks nodes except in young specimens. Pygidium poorly known; pygidial axis with about four rings (including terminal); pleural regions with about three pleural and two interpleural grooves, three pairs of marginal spines and a striated doublure (PI. 77, fig. 106). Young specimens in Belt’s collection have a convex glabella which, compared with adult specimens, is larger in proportion to the rest of the cranidium (PI. 77, fig. 9). 412 PALAEONTOLOGY, VOLUME 11 Measurements in mm. SM A3095 BM 59284 BM It 2585 Length of cranidium 110 100 6-1 Width of cranidium 17-2 16T (est.) 90 Length of cephalic axis c. 7-6 7-2 4-2 Width of cephalic axis 5-8 5-3 3-2 The lectotype is about 34 mm. long; SM A3095 is about 30 mm. long (restored). Remarks. I agree with Lake (1919) in regarding Salter’s Conocoryphe williamsoni as a synonym of Belt’s C. ? bucephala. Compared with C. ? bucephala the glabella of C. williamsoni seems to be slightly smaller in proportion to the rest of the cranidium but this may be because in C. williamsoni it is preserved convex (except where crushed on to the hypostoma), whereas in the relatively undeformed specimens of C. ? bucephala it is ‘spread’ by flattening and crushing. Otherwise, so far as the species can be com- pared, they agree quite well. The small cranidium collected by Stubblefield (1930) from Bentleyford Brook in Shropshire is slightly flattened compared with young specimens in Belt’s material and has a less prominent anterior border, but is otherwise very similar. The specimens collected by Stubblefield from concretions in Comley Brook, Shropshire (Cobbold 1927, p. 557) and considered by him to be the same species as the Bentleyford cranidium are very small but may provisionally be retained as Parabolinoides cf. bucephalus. I have not seen the specimens from the Ffestiniog Beds illustrated by Lake (1919) as belonging to the Tremadoc species Beltella verisimilis (Salter), but from the illustrations they appear to be laterally compressed specimens of Parabolinoides bucephalus. Generic position. The chief features of Conocoryphe ? bucephala — the tapered glabella with oblique furrows and rounded front, wide anterior border, small eyes close to the glabella, the course of the facial suture and the spinose pygidium — are those of Para- bolinoides Frederickson. Belt's original reference to Conocoryphe ? followed Salter’s example in using the name for what would now be called ptychoparioid trilobites. Subsequent writers (Reed 1900; Lake 1919; Henningsmoen 1957) agreed in assigning P. bucephalus to the Olenidae, but the wide cephalic border and divergent pre-ocular sutures are not typical of that family. Nevertheless, P. bucephalus otherwise resembles late Olenus species such as O. cata- ractes Salter which, however, has more thoracic segments and stronger glabellar furrows ; EXPLANATION OF PLATE 77 Figs. 1-10. Parabolinoides bucephalus (Belt). 1. Cranidium, Ffestiniog Beds near Penmaenpool, Merio- neth, X 2\, BM 59284. 2-5. Deformed cranidia, Upper Ffestiniog Beds near Dolgelly (Belt Collection), x 2\\ GSM Za4844, BM I 7580, BM I 7588, BM I 7590 respectively. 6. Crushed cranidium with part of thorax; Ffestiniog Beds? X 2£, BM It 2585. 7, 8. Lectotype and free cheek, Upper Ffestiniog Beds near Dolgelly (Belt Collection), x 2\, BM I 7578, BM I 7583. 9. Convex immature cranidium, prob- ably this species, X 5, BM 58499. 10. Complete specimen (type specimen of Conocoryphe williamsoni Salter), Lower Dolgelly Beds, Rhiwfelyn SM A3095; lOu, X24; 106, view of pygidium, X 5. 11. P. cf. bucephalus (Belt); immature cranidium from Orusia Shales (Lower Dolgelly Beds) in tributary of Bentleyford Brook, Shropshire, X 5, GSM D2655. Photographs by Mr. C. Friend. Palaeontology, Vol. II PLATE 77 RUSHTON, Upper Cambrian trilobites RUSHTON: REVISION OF TWO UPPER CAMBRIAN TRILOBITES 413 species of the (typically Tremadoc) olenid genus Beltella Lake differ in having truncate pleural tips and an entire pygidial margin. Leptoplcistides Raw, which is related to Beltella (Henningsmoen 1957, p. 265) but has spinose pleurae, differs from P. bucephalus in having axial spines and a different type of free cheek. Beltella solitaria Westergard (1922, pi. 14, fig. 1) has a less prominent anterior border and a truncate glabella, and seems to be a true Olenid. text-fig. 1. Diagrammatic drawings of cranidia of Parabolinoides species, about X 2-|. ( a ) P. contractus Frederickson (after Frederickson 1948). (b) P. hebe Frederickson (after Frederickson 1948). (c) P. palatus Berg (after Berg 1953). (d-f )P. bucephalus (Belt), restored from various specimens. P. bucephalus is placed in Parabolinoides rather than any other genus of the Para- bolinoididae such as Maustonia Raasch in Lochman or Taenicephalus Ulrich and Resser in Walcott because the post-ocular sutures are convex rather than sinuous (see Loch- man in Harrington et al. 1959, p. 0272). Bernia Frederickson which has convex post- ocular sutures is considered a synonym of Parabolinoides by Bell and Ellinwood (1962, p. 398). Comparison with other species (text-fig. 1). Parabolinoides bucephalus differs from other species of Parabolinoides in having the inter-ocular cheeks about half as wide as the glabella, whereas in other species they are only about one-third as wide. The most similar species are P. contractus Frederickson (1949) which, however, also differs in having a longer frontal area and an occipital node, and P. hebe Frederickson which has a relatively short anterior border and a more truncate glabella. P. palatus Berg (1953) has a relatively short, square glabella and more transverse post-ocular sutures. P. ? cordillerensis Lochman (1950) has relatively very narrow inter-ocular cheeks and some- what sinuous post-ocular sutures. 414 PALAEONTOLOGY, VOLUME 11 Occurrence. Parabolinoides bucephalus has been found in North Wales and possibly in Shropshire at horizons at or near the top of the Olenus Zone and in the Parabolina spinulosa Zone (Upper Ffestiniog and Lower Dolgelly Beds). The specimens in Belt’s collection are all labelled ‘Upper Ffestiniog, Dolgelly’ and in his account of the species Belt (1868) mentions the localities Gwern-y-barcud (near Penmaenpool), Mynydd Gader, and near Craig-y-Dinas. G. J. Williams collected P. bucephalus at the same horizon in Nant Cistfaen and on Trinant, near Arenig (Lake 1919, pp. 107, 109). Mr. S. W. Hester of the Geological Survey has collected a specimen identified by Stubblefield as Beltella cf. bucephala from the basal Lower Dolgelly Beds at Ogof Ddu, one mile east of Criccieth, Caernarvonshire. The type of Conocoryphe mUiamsoni Salter is stated by Lake (op. cit., p. 107) to be from the ‘Upper Lingula Flags’ (= Dolgelly Beds) at Rhiwfelyn in the Upper Mawddach Valley; a variety of horizons is exposed near Rhiwfelyn (Wells 1925, map, pi. 32) but the lithology of the specimen agrees with that of the Lower Dolgelly Beds which yield Parabolina spinulosa (Wahlenberg). Immature cranidia of Parabolinoides cf. bucephalus have been collected from a sep- tarian nodule from Comley Brook, south Shropshire (Cobbold 1927, pp. 556-7) where they were associated with Parabolina spinulosa and Orusia lenticularis (Wahlenberg). In a section at Bentleyford Brook a single immature cranidium was collected near the base of a 63-ft. thickness of micaceous shales which, at higher horizons, yielded also Orusia lenticularis and Parabolinites ? [Parabolinella] aff. williamsonii (Belt). Apart from P. bucephalus , recorded species of Parabolinoides are confined to the Con- aspis Zone in the central parts of the United States of America (Lochman and Wilson 1958, see text-figs. 9, 10, 1 1). If, as has been suggested, the top of the Elvinia Zone is not higher than the top of the Olenus Zone (Rushton 1967), P. bucephalus and the American species of Parabolinoides occur at about the same horizon. North Welsh European North American succession Zones Zones Lower Doglelly Parabolina spinulosa Conaspis Ffestiniog Upper Maentwrog f Olenus e (6 sub-zones a-f) d Elvinia Family olenidae Burmeister Subfamily leptoplastinae Angelin Genus sphaerophthalmus Angelin 1854 Type species (subsequently designated by Linnarsson 1880). Trilobites a/atus Boeck, 1838. According to Henningsmoen (1957, p. 21 1) there are four species of Sphaerophthalmus, all occurring in the Upper Cambrian Peltura Zones of the North Atlantic Province. He RUSHTON: REVISION OF TWO UPPER CAMBRIAN TRILOBITES 415 lists S. alatus (Boeck), Olenus humilis Phillips 1848, S. majusculus Linnarsson 1880, and S', major Lake 1913. Henningsmoen discussed the nomenclature of the species; his con- clusions may be summarized as follows : 1. In 1880 Linnarsson stated that Olenus humilis Phillips was a junior synonym of Sphaerophthalmus alatus (Boeck). Henningsmoen showed that this was a mistake, the species being distinct. 2. Other workers accepted Linnarsson’s statement; one such was Lake (1913) who figured and described material from Phillips’s type-area for Olenus humilis under the name S. alatus (Boeck). Henningsmoen suggested that the name humilis should be revived for S. alatus of Lake (not Boeck). Lake (op. cit.) also described a new species, S. major. text-fig. 2. Reconstruction of cephalon of Sphaerophthalmus humilis (Phillips) showing attitude of free cheek, X 10, anterior and lateral views. Cranidium after Westergard. 3. Westergard (1922) likewise figured Phillips’s species under the name S. alatus and he tentatively assigned a second species to S. major Lake; he also re-illustrated S. maju- sculus Linnarsson. Henningsmoen pointed out that although the Scandinavian specimens of ‘S’, major ’ should be referred to S. alatus (Boeck) the same was not necessarily true of the type-material of S. major. The morphology of Sphaerophthalmus. The cranidia of species of Sphaerophthalmus are convex so that the differences between specimens preserved in limestone and those in shale are quite important, as can be seen in the figures given by Westergard (1922, pi. 13, figs. 9-35). For example, the transversely arched anterior border, when flattened, be- comes an emargination in dorsal view (Westergard 1922, pi. 13, figs. 20 and 23), and the post-ocular cheeks are bent up from their original steeply down-sloping attitude so that the width of the cranidium is increased. When the cranidium is flattened the facial suture describes an elongated S-shape, but when preserved convex the course is, in dorsal view, curved anteriorly and nearly straight behind and, in side view, arched up at the palpebral lobe. The free cheek of Sphaerophtha/musbears, a hemispherical eye which projects beyond the general line of the facial suture; it appears that this can only fit under the arched palpe- bral lobe if the free cheeks slope steeply, as I have attempted to show in text-fig. 2. As the cheek spine is bent down in relation to the general plane of the margin of the free cheek it seems that, in life, these spines curved under the cephalon (text-fig. 2), in which position they would prevent the trilobite from crawling on the sea floor; Henningsmoen (1957, p. 78) has suggested that Sphaerophthalmus may have been a strong swimmer. 416 PALAEONTOLOGY, VOLUME 11 Sphaerophthalmus major Lake 1913 Text-fig. 3b; Plate 78, figs. 1-8 1913 Sphaerophthalmus major Lake, p. 77, pi. 8, figs. 7, 9-13; ? non fig. 8. 1949 Sphaerophthalmus major Lake; Edmonds, p. 60. 1952 Sphaerophthalmus major Lake; Hutchinson, p. 90, pi. 4, fig. 17?; non fig. 16 [= Ctenopyge fletcheri (Matthew)]. 1957 Sphaerophthalmus major Lake 1913; Henningsmoen, p. 217. 1957 Sphaerophthalmus humilis (Phillips 1848) \partim\ ; Henningsmoen, pi. 22, figs. 7, 11?, 15 only. 1957 Sphaerophthalmus minor [.y/c] ; Henningsmoen, p. 218 (error for S. major). Lectotype. As lectotype I select the cranidium GSM 8903 (PI. 78, fig. 2) illustrated by Lake (1913, pi. 8, fig. 7) which was collected from the White Leaved Oak Shales at White Leaved Oak, Malvern, and presented to the Geological Survey by Miss M. Lowe. Other material. Apart from BM I 14849 (the original of Lake’s pi. 8, fig. 10) which has not been seen, all of Lake’s syntypes have been examined and are considered to be conspecific except possibly I 14851 (the original of Lake’s pi. 8, fig. 8); this consists of two free cheeks, probably of a Ctenopyge species, and a thorax, possibly of S. major, upside-down and back-to-front in relation to them. Topotype material in the collections of the Geological Survey Museum, the Sedgwick Museum, the Oxford University Museum, and the British Museum (Natural History) has also been examined. Diagnosis. A species of Sphaerophthalmus with palpebral lobes which do not project laterally and whose centres are opposite or slightly behind the anterior ends of SI ; eye- ridges weak or absent; inter-ocular cheeks more than half as wide as glabella (about two-thirds as wide as glabella in shale-preserved specimens) at eye-line ; occipital spine relatively long; free cheek semicircular with long spine springing from the middle of the lateral margin; thorax with axial nodes and pleural spines, pleural regions about two- thirds as wide as the axis; pygidial outline rounded behind, pleural regions about two- thirds as wide as axis which is composed of about three rings. Description. The general features of the cranidium are very like those of S. humilis. Glabella prominent but commonly flattened in shale-preserved specimens, in which cases the axial furrow may be deepened. SI oblique backwards, concave, joined across glabella in an even curve; S2 very short or indistinct impressions in the sides of the glabella. Occipital spine long but in many specimens only a tubercle is preserved. EXPLANATION OF PLATE 78 Sphaerophthalmus species, X 6. Pigs. 1 and 5 are of specimens from the Upper Dolgelly Beds near Dolgelly, Merioneth (Belt Collec- tion); the remainder are from the White Leaved Oak Shales of the Malvern Hills. Pigs. 1-8. S. major Lake. 1. Small cranidium showing long occipital spine, BM I 7619 partim (see Fig. 5). 2. Lectotype cranidium, GSM 8903. 3. Cranidium showing mould of occipital spine, SM A53115. 4. Cranidium, SM A53088. 5. Cephalon and thorax, paratype, BM I 7619 partim. 6, 7. Free cheeks, SM A52564, 52562. 8. Thorax and pygidium, paratype, SM A509. Figs. 9, 10. S. cf. a/atus (Boeck), external moulds of cranidium and free cheek, SM A53130. Figs. 1 1-15. S. humilis (Phillips). 11,12. Cranidia, SM A53093,GSM 8963. 13. Cranidium from Phillips’s Collection, Ox A287a. 14. Free cheek, SM A52546. 15. Thorax and pygidium figured by Lake, GSM 8913. Photographs by Mr. J. M. Pulsford. Palaeontology, Vol. 11 PLATE 78 RUSHTON, Upper Cambrian trilobites RUSHTON: REVISION OF TWO UPPER CAMBRIAN TRILOBITES 417 Anterior border short (sagittally), strongly arched in anterior view but probably only slightly emarginate in dorsal view; in the lectotype it is strongly emarginate and up- turned because compressed. Eye-ridges very weak on both internal and external moulds, moderately oblique. Palpebral lobe narrow, arched up in lateral view, emarginate in dorsal view; centres opposite or just behind the anterior ends of SI. Inter-ocular cheeks about two-thirds of the glabellar width at the eye-line, but probably little more than half in specimens preserved with their full convexity; post-ocular cheeks as wide as to four- fifths the width of the occipital ring, but probably about two-thirds in convex specimens. c d text-fig. 3. Diagrammatic drawings of flattened cranidia, free cheeks and pygidia of Sphaerophthalmus species, about X 4. (a) S. humilis (Phillips), partly after Westergard. ( b ) S. major Lake (the pygidium is partly conjectural), (c) S. majusculus Linnarsson; pygidium, associated cranidium (after Wester- gard) and doubtfully attributed free cheek (after Henningsmoen). (d) S. alatus (Boeck), after Wester- gard (pygidium unknown). Free cheek semicircular with a curved spine up to twice as long as the genal field springing from the middle of the lateral margin. Border of about even width. The hemispherical eye is placed far back but distinctly forward of the border-furrow. The thorax and pygidium have been adequately described by Lake, but it should be noted that the pleural and axial spines are generally not preserved, and that the pygidial flanks may bear more than one pair of furrows. Comparison with other species (text-fig. 3). Compared with S. major S. humilis (text-fig. 3a; PI. 78, figs. 11-15) has narrower fixed cheeks; the inter-ocular cheeks in convex specimens are less than half as wide as the glabella at the eye-line and even when 418 PALAEONTOLOGY, VOLUME 11 flattened they are only slightly more than half. The eyes are further back in S. humilis and the eye-ridges are consequently more oblique, the occipital and genal spines are shorter, the border of the free cheek tends to widen forwards, the pleural regions of thorax and pygidium are narrower and the pygidium is triangular rather than rounded behind. Measurements made on several specimens from the White Leaved Oak Shales have shown that the width of the inter-ocular cheeks of S. humilis is 0-53 or less of the width of the glabella, whereas that of S’, major is 0-63 or more. S. alatus (text-fig. 3d; cf. PI. 78, figs. 9, 10) resembles S. major in having the palpebral lobes opposite SI and has fixed cheeks of comparable width, but it differs in having palpebral lobes that project laterally and are joined to the glabella by distinct eye-ridges; the anterior ends of SI are further forward in relation to the length of the cephalic axis in S. alatus than S. major , and the palpebral lobes are correspondingly further forward on the cranidium. In S. alatus the anterior part of the fixed cheeks narrows forwards; the occipital spine is shorter than that of S. major and the cheek spine is set further back; the posterior part of the margin of the free cheek is nearly straight, whereas in S. major it is curved. S. majusculus (text-fig. 3c) has wider fixed cheeks than S. major, distinct eye-ridges, palpebral lobes that project slightly laterally, and a node in place of the occipital spine. The free cheek tentatively assigned to the species by Henningsmoen (1957, p. 219) has a shorter spine set further back than that of S. major. The pygidium of S. majusculus is relatively wider than that of S. major and is triangular rather than rounded behind. Measurements in mm. Length of cranidium GSM 8903 (lectotype) 4-2 SM A53088 c. 3-4 Width of cranidium at eyes 5 0 4-1 Width of cranidium posteriorly 6-6 5-6 Length of cephalic axis 3-5 2-9 Width of cephalic axis 2-6 1-9 Sphaerophtlialmiis major probably reached a total length of about 10 mm. Occurrence. Sphaerophthalmus major is found within the Zone of Peltura scarabaeoides in England, Wales, Sweden, possibly in Nova Scotia (Canada), and apparently also in Norway; such evidence as there is suggests that it occurs in the Subzone of Ctenopyge linnarssoni. In discussing these occurrences reference is made to Henningsmoen’s (1957, p. 39) subdivisions of the P. scarabaeoides Zone: (Peltura par ado xa Subzone Parabolina lobata Subzone Ctenopyge linnarssoni Subzone Ctenopyge bisulcata Subzone In Scandinavia Ctenopyge bisulcata (Phillips) and C. linnarssoni Westergard are known only from their own subzones, whereas C.pecten (Salter), P. scarabaeoides (Wahlenberg), and Sphaerophthalmus humilis (Phillips) occur in both subzones. England. The lectotype of S. major is from the White Leaved Oak Shales of the Malvern Hills. The detailed stratigraphy of these Shales is unknown and though the presence of RUSHTON: REVISION OF TWO UPPER CAMBRIAN TRILOBITES 419 the C. bisulcata Subzone is indicated by the association on single pieces of shale of C. bisulcata, C. pecten, S. humilis , and P. scarabaeoides, the presence of the C. linnarssoni Subzone is not proved. S. major has not been seen in association with undoubted frag- ments of C. bisulcata, but occurs with C. pecten and fragments of a Ctenopyge species recalling C. bisulcata, C. linnarssoni, or C. falcifera Lake. In the Geological Survey Merevale No. 1 Borehole, which penetrated the Monks Park Shales (Upper Stockingford Shales) in Warwickshire (Kellaway 1966), poorly preserved specimens of S. major were collected at depths of 190 ft. 6 in. and 193 ft. 0 in., from horizons slightly higher than those yielding S’, humilis and C. bisulcata (depths of 205 ft. 6 in. and 206 ft. 6 in.). Wales. One paratype of S. major is from the Upper Dolgelly Beds of the Dolgelly area in North Wales; the precise locality and zonal position are unknown. Lake records S. major from Moel Gron, near Dolgelly. Sweden. A block of Upper Cambrian bituminous limestone in the Sedgwick Museum, collected by Mr. S. C. A. Holmes from Vilske-Klefua, Vastergotland, Sweden (SM A56378-87), carries an association of S. major with Ctenopyge fletcheri (Matthew) and Peltura scarabaeoides westergaardi Henningsmoen, suggesting a horizon near the junc- tion of the Subzones of C. linnarssoni and Parabolinci lobatci (see Henningsmoen 1957, p. 39). Norway. Henningsmoen (1957, pi. 22, figs. 7, 15) illustrated two cranidia from the Sub- zone of C. linnarssoni at Viul, Ringerike, Norway, under the name of Sphaerophthalmus humilis ; the fixed cheeks are too wide and the occipital spine rather too large for that species and they are more probably cranidia of S. major. Nova Scotia. Hutchinson (1952, p. 90) records S. major from the upper part of the MacNiel formation ( Peltura Zone), and figures two specimens on his plate 4; one (fig. 16) is probably C. fletcheri incorrectly identified as S. major but the other (fig. 1 7) may be S. major (Henningsmoen 1957, pp. 205, 218). Acknowledgements. I thank Mr. A. G. Brighton (Sedgwick Museum, Cambridge), Dr. W. T. Dean (British Museum [Natural History]), and Mr. H. P. Powell (Oxford University Museum) for lending me specimens in their care. Sir James Stubblefield, F.R.S., and Mr. R. V. Melville kindly read the manuscript and made many helpful suggestions. This paper is published with the permission of the Director, Institute of Geological Sciences. REFERENCES bell, w. c., and ellinwood, h. l. 1962. Upper Franconian and Lower Trempealeauan Cambrian trilobites and brachiopods, Wilberns Formation, Central Texas. J. Paleont. 36, 385-423, pis. 51-64. belt, t. 1868. On the ‘Lingula Flags' or ‘Festiniog Group’ of the Dolgelly District. Part III. Geol. Mag. 5, 5-11, pi. 2. berg, r. r. 1953. Franconian trilobites from Minnesota and Wisconsin. J. Paleont. 27, 553-68, pis. 59-61. boeck, c. 1838. Uebersicht der bisher in Norwegen gefundenen Formen der Trilobiten-Familie; in Keilhau: Gaea Norvegica, I, 138-45. Christiania. cobbold, e. s. 1927. The Stratigraphy and Geological Structure of the Cambrian Area of Comley (Shropshire). Quart. J. geol. Soc. Lond. 83, 551-73, pi. 43. 420 PALAEONTOLOGY, VOLUME 11 edmonds, j. e. 1949. Types and figured specimens of Lower Palaeozoic Trilobites in the University Museum, Oxford. Geol. Mag. 86, 57-66. frederickson, E. a. 1949. Trilobite fauna of the Upper Cambrian Honey Creek Formation. J. Paleont. 23, 341-63, pis. 68-72. Harrington, H. j., et al. 1959. Treatise on Invertebrate Paleontology, Part O, Arthropoda 1. Univ. Kansas Press. henningsmoen, G. 1957. The Trilobite family Olenidae. Skr. norske Vidensk.-Akad., Mat.-naturv. Kl. I, 1-303, pi. 1-31. 1960. The Middle Ordovician of the Oslo Region, Norway, 13. Trilobites of the family Asaphidae. Norsk geol. Tidsskr. 40, 203-57, pis. 1-14. Hutchinson, r. d. 1952. The stratigraphy and trilobite faunas of the Cambrian sedimentary rocks of Cape Breton Island, Nova Scotia. Mem. geol. Surv. Canada, 263, 1-124, pis. 1-7. kellaway, g. a. 1966. In Ann. Rep. Inst. geol. Sci. Part 1, Summ. Prog. geol. Surv. G.B. for 1965, 49. lake, p. 1906-46. A Monograph of the British Cambrian Trilobites. Palaeont. Soc. [Monogr.]: (4), 1913, 65-88, pis. 7-10; (5), 1919, 89-120, pis. 11-14. linnarsson, j. g. o. 1880. Om forsteningarne i de svenska lagren med Peltura och Sphaerophthalmus. Sver. geol. Unders. Avh. Ser. C, 43, 1-31, pis. 1, 2. lochman, Christina. 1950. Upper Cambrian faunas of the Little Rocky Mountains, Montana. J. Paleont. 24, 322-49, pis. 46-51. and wilson, j. l. 1958. Cambrian biostratigraphy in North America. J. Paleont. 32, 312-50. Phillips, j. 1848. The Malvern Hills compared with the Palaeozoic Districts of Abberley, Woolhope, May Hill, Tortworth, and Usk. With Palaeontological Appendix. Mem. geol. Surv. G.B. 2 (1), 1-386, pis. 1-30. reed, f. r. c. 1900. Woodwardian Museum notes: On the British Species of the Genus Conocoryphe. Geol. Mag. (4), 7, 250-7. rushton, a. w. a. 1966. In Ann. Rep. Inst. geol. Sci. Part 1, Summ. Prog. geol. Surv. G.B. for 1965, 69. 1967. The Upper Cambrian trilobite Irvingella nuneatonensis (Sharman). Palaeontology, 10, 339-48, pi. 52. salter, J. w. 1873. A Catalogue of the Collection of Cambrian and Silurian Fossils contained in the Geo- logical Museum of the University of Cambridge. Cambridge. Stubblefield, c. J. 1930. A new Upper Cambrian section in South Shropshire. Summ. Prog. geol. Surv. G.B. for 1929— Part II, 54-62. wells, a. k. 1925. The Geology of the Robell Fawr District (Merioneth). Quart. J. geol. Soc. Loud. 81, 463-538, pis. 30-32. westergArd, a. h. 1922. Sveriges Olenidskiffer. Sver. geol. Unders. Avh. Ser. Ca, 18, 1-205, pis. 1-16. A. W. A. RUSHTON Institute of Geological Sciences Exhibition Road London, S.W. 7 Typescript received 25 February 1967 PROBABLE ANGIOSPERM POLLEN FROM BRITISH BARREMIAN TO ALB IAN STRATA by ELIZABETH M. KEMP Abstract. Two species of the genus Clavatipollenites Couper are described. One is a redescription of the type species, Clavatipollenites luighesii, from the Upper Barremian part of the Wealden Series. The other, Clavati- pollenites rotundas, is a new species from the Middle Albian Lower Gault. Three species of tricolpate angio- spermous grains are described. Only one of these, Tricolpites albiensis, from the Upper Albian, occurs in sufficient quantity to be formally named as a new species. These forms constitute the earliest record of recogniz- able angiosperm grains from England. During an investigation into the spore and pollen assemblages of the marine Lower Greensand and Gault of southern England, several pollen species were recovered which are of particular botanical and stratigraphical interest. Two of these are monosulcate pollens referable to the genus Clavatipollenites Couper; the others are tricolpate angiospermous forms, the earliest undoubted pollen of this group from England. Of these last forms, only one was recovered in sufficient quantity to permit formal designa- tion as a new species. The monosulcate pollens described here are of interest in that they possess some angiosperm characters and their appearance slightly predates that of grains of undoubted angiospermous affinity. Acknowledgements. Research facilities in the Department of Geology, Sedgwick Museum, Cambridge, were made available by Professor O. M. B. Bulman. Mr. N. F. Hughes took a constant interest in the course of this work and provided the sample from the Kingsclere Borehole. Financial assistance was provided by a British Commonwealth Scholarship from 1963 to 1966. LOCATION AND STRATIGRAPHY OF SAMPLES The samples yielding the type material of the new species described here are part of a more extensive collection of Lower Greensand, Gault and Upper Greensand. The systematic and stratigraphic palynology of these sediments has been investigated in detail for an unpublished thesis. The sediments are of Aptian and Albian age, with a rich fauna which permits their subdivision in terms of ammonite zones. The Lower Greensand samples have been collected chiefly from the Isle of Wight, with minor sampling in the Weald and in the northern sedimentary basin near Leighton Buzzard. The Gault and Upper Greensand have been collected over a wider geographic area, extending from near Leighton Buzzard, through sections in the northern Weald to the Isle of Wight and the Dorset coast in the south and south-west. In addition, the sample from 475 ft. in the Kingsclere Borehole, previously studied by Couper (1958), has been re-examined as it provided the type material of the species Clavatipollenites hughesii. This species persists into younger strata. The type sample comes from the Wealden, and is probably of Upper Barremian age. [Palaeontology, Vol. 11, Part 3, 1968, pp. 421-34, pis. 79 81.] 422 PALAEONTOLOGY, VOLUME 11 More detailed references to the samples used in this study and to the sections from which they were collected are given below: Sample FI 54 — Calcareous pale yellowish-grey clay with shell fragments. Gault at Ford Place, Wrotham (Nat. Grid, reference 51/636591). Bed 20, near base; niobe Subzone of the dentatus Zone; Middle Albian. The detailed stratigraphy of this clay pit is given by Milbourne (1963), who subdivided the section into 82 beds on the basis of lithology. Sample K475 — Medium-grey siltstone with small dark patches, possibly plant fragments. Wealden Series, 475 ft. in the D'Arcy Exploration Company's Kingsclere No. 1 Borehole; probably Upper Barremian. Further details of the palynology and stratigraphic position of this sample are given by Hughes (1958) and Couper (1958). Sample F312 — Calcareous friable pale yellowish-green silty sandstone. Upper Greensand, west side of Lulworth Cove, Dorset (Nat. Grid reference 30/825799). 18 ft. beneath the base of the Exogyra Rock, which underlies a series of stone bands at the top of the Upper Greensand and serves as a convenient local marker (Wright, in Arkell 1947). Beds above the Exogyra Rock can be referred to the dispar Zone. The zonal position of the beds below is less certain, but they are probably of Upper Albian age. Sample F270 — Grey and cream mottled sandstone with ferruginous pebbles and patchy iron staining. Munday’s Hill pit, Leighton Buzzard (Nat. Grid reference 42/930265). 6 in. beneath base of Carstone Breccia in the upper part of the Woburn Sands. The precise age of the sample is unknown. The Car- stone Breccia is partly of Lower Albian age; the sample in question comes probably from the Lower Albian. Location of type material. Rock samples, strew slides and individual mounts made from preparations are lodged in the Sedgwick Museum, Cambridge. Slides are deposited in the Palynology Collection of that museum. Formally named species are based on examination of 100 specimens in each case, and all of these are available for restudy. The co-ordinates quoted are those of the Leitz Dialux microscope Serial No. 526724 in the Sedgwick Museum. Preparation procedures. Spore assemblages were extracted by solution of silicates in cold hydrofluoric acid following the removal of carbonates, where necessary, with 10-15 per cent. HC1. Residues were oxidized in Schulze solution, two short periods of oxidation usually being used. The total oxidation time did not in any case exceed twenty minutes. Residues were briefly cleared in less than 1 per cent. NH4OH and the microfossils separated from the remaining mineral debris by centrifuging in zinc bromide solution. SYSTEMATIC PALYNOLOGY Anteturma pollenites Potonie 1931 Turma plicates Naumova 1939 emend. Potonie 1960 Subturma monocolpates Iversen and Troels-Smith 1950 Genus clavatipollenites Couper 1958 1958 Clavatipollenites C ouper, p. 159. 1961 Retimonocolpites Pierce, p. 47. Type species. Clavatipollenites liughesii Couper 1958, p. 159, pi. 31, figs. 19-22. Diagnosis. As given by Couper (1958). Remarks. On a morphological basis Clavatipollenites resembles Liliacidites Couper (1953, p. 56) which was proposed to incorporate monosulcate grains of reticulate exine pattern, or, more precisely, ‘for the reception of fossil pollens of liliaceous affinities that E. M. KEMP: PROBABLE ANGIOSPERM POLLEN 423 cannot be more accurately placed’. The species originally referred to Liliacidites formed part of an Upper Cretaceous and Tertiary complex which had a large component of grains of unquestionably angiospermous origin. It would seem that the retention of a separate genus is valuable for Lower and lower Upper Cretaceous forms the affinity of which is in question. Species from several widely separated localities have been referred to Clavatipollenites. Not all of these, however, conform to the diagnosis of the genus. Pocock (1962, p. 74, a b c text-fig. 1. Grain orientation in Clavatipollenites. L = length, W = width, D = depth. Designa- tions P, Ex, and E3 correspond with those of Couper (1958) for describing monosulcate and tricolpate pollens. Orientation: a, polar view; b, equatorial longitudinal view; c, equatorial transverse view. pi. 12, figs. 190-2) described a large form which he named Clavatipollenites couperi , from the Upper Jurassic-Neocomian Vanguard Formation of western Canada. The specimens he figured do not show the presence of a tectate exine convincingly, and the species may perhaps be more accurately referred to Monosulcites or Ginkgoeyeadophytus. Helal (1965, pi. 17, fig. 19; 1966, pi. 34, figs. 54, 55) recorded both C. couperi and C. hughesii from subsurface Jurassic sediments in the Western Desert, Egypt. Both of the grains figured by this author appear to be distinctly three-furrowed, suggesting Eucoin- miidites rather than Clavatipollenites. The type species, C. hughesii , has been recorded by Brenner (1963) together with a smaller form which he designated C. minutus, from Potomac Group sediments in Mary- land. Von der Brelie (1965, p. 151, pi. 13, figs. 6-8) found C. hughesii in sediments of probable Aptian-Albian age in Germany. Archangelsky and Gamerro (1967) reported the presence of the species in assemblages of probable Barremian age in southern Argentina. Pierce (1961, p. 47, pi. 3, fig. 87) described a monosulcate pollen from Cenomanian clays and lignites in Minnesota. This species, which he named Retimonocolpites dividuus, 424 PALAEONTOLOGY, VOLUME 11 Pierce designated as type species of the genus Retimonocolpites. This genus is regarded as synonymous with Clavatipollenites on the basis of Pierce’s definition. Both the type species, C. hughesii, and a new species, C. rotundus, were recorded from samples investigated during the present study. The expanded description of C. hughesii given below is based on re-examination of specimens from Couper’s type sample. The terminology of the grain dimensions and orientations used in the descriptions is shown in text-fig. 1 . Comparisons of the size and shape of grains assigned to the two species are shown in the scatter diagram text-fig. 4, histograms of size frequency dis- tributions in text-fig. 5. text-fig. 2. Diagrammatic representation of Clavatipollenites rotundus sp. nov. Large figures, X 2,000. A, proximal face, LO pattern, position of sulcus dotted, b, distal face, left-hand side showing reticulum, right-hand side showing exine stratification in section. Dark zone adjacent to the sulcus is stippled, c, equatorial longitudinal view, d, equatorial transverse view. Clavatipollenites rotundus sp. nov. Plate 79, figs. 1-19; Plate 80, figs. 1-8; text-fig. 2 Diagnosis. Monosulcate pollen grains of subcircular to elliptical outline, with a dis- tinctly tectate exine. The sulcus is conspicuous, round-ended, and bordered by thicken- All figures at magnification X 1,000 unless otherwise stated. Figs. 1-19. Clavatipollenites rotundus sp. nov. All specimens from sample F154, niobe Subzone of the dentatus Zone, Lower Gault at Ford Place, Wrotham. 1-5, FI 54/2; 35-0, 92-7. 1, Median focus showing tectate exine in section at equator (x 2,000). 2, Fligh focus on sulcus. 3, 4, Deeper foci. 5, Proximal focus on reticulum with sexine partly detached. 6, 7, F154/1 1 ; 27 0, 108-2; grain obliquely compressed. 6, Deep focus. 7, High focus showing end of sulcus and darkened zone. 8, FI 54/Separate 15, 40 0, 97-6; grain showing pronounced infolding associated with sulcus. 9-11, FI 54/Separate 1 ; 38-3, 99 2. 9, Proximal focus showing reticulum. 10, Median focus on sulcus and associated struc- ture. 11, Distal focus. 12, 13, FI 54/10 ; 34-2, 90-2. 12, High focus on proximal face, sexine partly removed. 13, Equatorial focus. 14, 15, FI 54/12; 35-8, 89-8. 14, Median focus of specimen in end-on view. 15, High focus on reticulum. 16, F154/Separate 8; 39 0, 98-1; median focus showing folding adjacent to sulcus. 17-19, FI 54/Separate 9; 39-2, 98-5; Holotype. 17, High focus on distal face. 18, Median focus on sulcus with exine stratification distinct at equator. 19, Proximal focus. EXPLANATION OF PLATE 79 Palaeontology , Vol. 11 PLATE 79 KEMP, Mid-Albian pollen E. M. KEMP: PROBABLE ANGIOSPERM POLLEN 425 ings or infoldings of the nexine. The reticulate pattern of the sexine is formed from the fused tips of bacula; the mesh diameter is from 0-7 to 2-0 p. The mean grain width lies in the range 22-27 p, mean grain length 25-30 p. Sexine shows a tendency to separate from nexine. Holotype. Sample F154, separate mount 9; 39-2, 98-5. Plate 79, figs. 17-19. Lower Gault, Ford Place, Wrotham; Middle Albian. Grain orientated with sulcus uppermost, extending full length of grain, bordered by darkened zone 2-4-2-5 p wide on either side. Sulcus slightly expanded and rounded at extremities. Nexine 08 p thick, sexine 1-2 /a. Length x width 26x25 p. Dimensions. (Measured on 100 specimens from Sample F154; see text-fig. 5.) Grain length 20(27)32 yu., grain width 18(25)31 /a, grain depth 17(22)28 /a. Orientation. (For grain positions see text-fig. 1.) Grains in position a — 68 per cent., position b — 7 per cent., position c — 25 per cent. Description. The original grain shape, as deduced from measurements and from grain orientations appears to have been subspheroidal to oblate spheroidal. The ratio grain length to grain width is shown in the scatter diagram text-fig. 4. The sulcus extends for almost the full length of the grain and is parallel-sided through- out most of its length, but frequently shows slight expansion at the extremities. The ends are usually rounded. Most specimens show a darkened zone 1 -5-2-0 p wide on either side of the sulcus throughout its length. This either represents a thickening of the nexine adjacent to the sulcus or a doubling of this layer caused by infolding consequent on rupture of the grain along the sulcus. Optical sections of grains (text-fig. 3) suggest that the last interpretation is more likely to be correct. Some specimens, for example, that illustrated in text-fig. 3b , show a small gap in the thickened nexine layer adjacent to the sulcus, suggestive of breakage during infolding and compression. These sections also demonstrate that the sexine is not involved in the infolding. The stratification of the exine is distinct. The nexine is 0-8-T8 p thick. The bacula arising from this layer expand and coalesce at their tips to form the reticulum of the ectosexine. The meshes of the reticulum are irregularly polygonal, separated by muri 0-3-0-5 p wide, which are formed from single rows of fused bacula. The meshes are gene- rally of uniform size over the entire grain surface. The sexine characteristically separates from the nexine; separation is usually more extreme at the grain ends. Some grains were observed in which the sexine had been entirely removed. Remarks. Clavatipollenites rotundas appears similar to Pierce's species Retimonocolpites dividuus , although the single illustration given by Pierce (1961, pi. 3, fig. 87) makes comparison difficult. Pierce does, however, mention the tendency for the exine layers to separate. Brenner (1963) recombined Pierce’s species with Liliacidites, with little descriptive comment. The specimens from the Patapsco Formation show some evidence of infolding adjacent to the sulcus but not to the degree seen in the English forms. The American species appears less spherical and thinner and consequently tends to fold more readily than its English counterpart. Distribution. In England C. rotundas is confined to Albian sediments. Its earliest recorded appearance is in Sample F270, which is probably of Lower Albian age. This occurrence is the only one observed beneath the base of the Gault. The species was noted in greatest abundance in the vicinity of the niobe C 5586 F f 426 PALAEONTOLOGY, VOLUME 11 Subzone at Wrotham and at Folkestone, although it persists throughout the Albian. These concentra- tions, which do not in any case exceed 5 per cent, of the total spore and pollen content, may be due to a facies factor as they are associated with a rich benthonic fauna and ferruginous bands. The distribution in time of C. dividuus is similarly restricted in America. Brenner (1963) records the species as being confined to the Patapsco Formation (Albian). a b c d text-fig. 3. Optical sections of specimens of Clavatipollenites rotundus sp. nov., drawn from photographs to show the structure associated with the sulcus. All x 1,000. Nexine shown solid black. a, FI 54/Separate 8; 39-0, 98T; nexine appears as thickening adjacent to sulcus, b , FI 54/Separate 1 ; 38 3, 99-2; optical section of grain in equatorial view. Slight obliquity causes sulcus to appear in section at top of drawing. Sexine is detached, not in- volved in infolding; the nexine adjacent to the sulcus appears split. Darkened zone next to the remainder of sulcus shown stippled, c, F 1 54/10 ; 36-6, 106 0; split grain showing a similar structure to b. d, F154/2; 32 0, 100 1 ; grain in end-on view. Clavatipollenites hughesii Couper emend. Plate 80, figs. 9-19 Remarks. Couper (1958) gave no separate diagnosis for C. hughesii, so his description has been emended slightly and is here restated as a diagnosis. Emended diagnosis. Monosulcate pollen grain, sulcus extending full length of grain, gaping in its central region and tapering slightly towards extremities. Equatorial outline elliptical to subcircular. Exine consisting of an inner unsculptured nexine 0-5-1 -0 /x thick E. M. KEMP: PROBABLE ANGIOSPERM POLLEN 427 and a sexine formed of baculate projections approximately 1 -0 p long, which either remain discrete or fuse at their tips to form a microreticulum. Lumina of reticulum irregularly polygonal, 1 -0-2-0 p in diameter. • Clavatipollenites hughesii + C. rotundus c/1 a o i_ u E 30- c JZ 20- + + + + + + 4+ + +++4+ + +-H-++ 44 + + 4 + + + + 4+4+444 +4+4-4 4 44 + + + + + +++ + + • + + + + + • • •• Rtt • + + • V •• * + + •••• •• • • cn c O c -5 •— « £ (Z X 8 - 3 "SL x> p 2 a a 2 — o a % y o. £ ™ si _ 60 $0 03 'C 3 e a a Si 3 C ' 7* *■<— ' O o O c' & 5j “ < « . 73 C - 03 ■&£ o v o 3 & £ “ c ^ £ 'o js co oo .S c £ ^ 03 D i_ J-l ^ o cl CO 13 -T £ § 73 •— ( . t ° £<2 c g 2 £ o.g- 8 “ O * . — q= a> a> £ % . C/3 CL >. 03 to £ o 13 c ^ 73 C3 C > o3 X . a X> o ? o 73 CL • -h on £ a> o3 ctf ■ .£ to ) £ 73 r£ ^ CO y .£ c 73 3 <■+-< ^ O O ^ .s & CL co CO <1 si •3 O £ p 3 °- i 8 1 c i w ; x) > y - c/5 I U 452 PALAEONTOLOGY, VOLUME 11 For sporophylls which fell in an inverted position, the dehisced sporangium wall and upper pedicel surface will form the lower compression surface. The pedicel and its hairy alation, together with the sporangium wall, will be preserved. But the position of the sporangium wall in the matrix will be influenced by the pedicel collapsing on top of it. The wall will thus be compressed close to the pedicel surface (PI. 83, figs. 7, 8). The original lower surface of the pedicel and the keel will collapse on to the underlying pedicel surface and will not be preserved. The heel collapses to a more or less vertical position, but remains separated from the pedicel surface (PI. 83, fig. 6). The originally adaxial lamina mid-rib collapses into its morphologically upper surface; that is, down- wards in the matrix. It is suggested that the step between the lamina and pedicel in these inverted specimens is due to the original morphology at the juncture of lamina and pedicel. Petrified specimens (Arber 1914, pi. 22, fig. 9; pi. 24, fig. 20; pi. 26, fig. 43) show that the distal face of the heel is a straight continuation of the lower surface of the lamina, and that there is a slight depression on the upper surface at the distal end of the lamina, directly above the heel. This is likely to be preserved after compression, as a difference in level between lamina and pedicel (text-fig. 3). This interpretation is again consistent with the specimens studied. Specimens in which the lamina mid-rib is seen as a groove depressed into the matrix (i.e. in the lower counterpart) have no heel in section, though the sporangium wall can be seen in trans- verse section (PI. 83, figs. 7, 8). This is never as well defined as in the upright specimens because, as is suggested above, the pedicel collapses on top of the sporangium wall during the process of compression. And when the mid-rib appears as a ridge projecting upwards from the cleaved fossil surface (i.e. in the upper counterpart) the heel is present in the matrix, as seen in section, and the sporangium wall is absent. The keel is never seen in either upper or lower counterpart fossils and the pedicel surface is always more or less flat. Sporophylls which fell in the sediment in an upright position tend to have a smaller lamina length pedicel length ratio than those that fell inverted (text-fig. 4). This is pre- sumably because the relatively greater lamina length favoured settling in the inverted position. The size of the sporophyll would probably be a function of its position on the strobilus; the larger sporophylls towards the base, the smaller ones towards the apex. This conclusion is evident from studies on petrified material (Arber 1914, pi. 23, fig. 13). Specimens of sporophylls which have been compressed laterally show the original vertical aspect of the keel, heel, pedicel, sporangium wall, and lamina. The pedicel alation is preserved in most specimens (PI. 84, figs. 8, 10) as a rugose margin above the pedicel, a very doubtful consequence of compression of the hairs which are seen in the upright and inverted states. From the evidence available in these compression fossils, it is not possible to ascertain the original morphology of the pedicel alation; whether it was horizontal, or upturned at the edges to protect (even partially) the sporangium. Indeed it must be accepted that the attribution of the laterally compressed specimens to this species (which is based primarily on upright and inverted specimens) is somewhat arbitrary. But at least all three states can be reconciled with a single reconstruction such as that offered in text-fig. 1. Until spores are found enclosed on the sporangium, or other indirect evidence is obtained, the correlation of the sideways compressed fossils with the other two states remains hypothetical. BOULTER: COMPRESSED LYCOPOD SPOROPHYLL FROM SOMERSET 453 COMPARISON WITH PREVIOUSLY DESCRIBED SPOROPHYLLS The majority of specific, and indeed generic, diagnoses of detached lycopod sporophylls are concerned principally with the size and shape of the lamina and pedicel, and in some cases with the type of spore extracted. No spores have been found in the present study, and so the only comparable characters are size and shape. Both these have proved to be 10 CD <9. o-8 1 06 O' CD o ° \ x X o o O t X O O o 1 . O ° o o o o X o x o x X O 3 CO 0-4 02 10 20 30 40 50 length of lamina (cms.) text-fig. 4. Length of sporophyll lamina plotted against length of pedicel, for specimens of Lepidostrobophyllum alatum sp. nov. compressed in upright (o) and inverted (x) states. The distribution shows that the longer lamina give a preferred orientation towards the inverted state (clustering of symbol .v at right of diagram). extremely variable in Lepidostrobophyllum alatum and so the comparison with other sporophylls is limited by our ignorance of the details of sporangium and pedicel for most of the earlier described species. Lepidostrobus triangularis Zeiller. The original description by Zeiller (1886) and sub- sequent descriptions (Arber 1922, Crookall 1966) show that this species is very similar in shape and size to the smaller specimens described in this paper as having been fos- silized in the upright state. The pedicel resembles that in specimens of L. alatum which have cleaved across the layer of the pedicel and not along the sporangium wall (PI. 84, fig. 5). But specimens labelled as L. triangularis in the Kidston Collection, Geological Survey, London (nos. 819a, 6249, 8196) show a clearly defined edge to the pedicel, which is too narrow to have alation. Also these specimens have a completely flat con- tour, and so their compression history cannot be compared to that of L. alatum. 454 PALAEONTOLOGY, VOLUME 11 Lepidostrobus goodei Jongmans (Jongmans 1931). In his recent review Crookall (1966) describes this species as having a wedge-shaped pedicel 5-7 mm. long and 2-3 mm. wide. This again could represent upright specimens of L. alatum in which the pedicel alation and sporangium wall are compressed in the matrix, if they are present at all; sporophylls at the extreme tip of the strobilus may not have had a well-developed alation or spor- angium. The lamina of L. goodei is of a comparable size to upright specimens of L. alatum. Lepidophyllum graeile Lesquereux. Lesquereux’s type figure (Lesquereux 1884) has similar shape and size to upright specimens of L. alatum which have cleaved along the sporangium wall. The species has not been described from Britain. Lepidostrobus brevifolius Lesquereux (Lesquereux 1857). This has a wedge-shaped, keeled pedicel, 0-8-1 -0 cm. long, and a triangular lamina, 6-8 mm. long (Crookall 1966). This species can again be compared to small specimens of L. alatum. Lepidostrobus lancifolius Lesquereux. The type figure of this species (Lesquereux 1870) shows identical features to the inverted state of L. alatum : acuminate apex to lamina, hastate base to lamina, prominent sporangium wall to pedicel, and similar dimensions. No details are available in the literature of the step between lamina and pedicel or whether the mid-rib formed a ridge or a groove. Lepidophylloides acuminatus (Lesquereux) Crookall (Crookall, 1966). Although no specimens of L. alatum as large as those described for this species have been collected, the shape and proportions of the two species are identical. Lepidocarpon mazonensis Schopf. Schopf (1938) described a side-compressed sporophyll containing a seed megaspore which has a rugose texture on the margin of the integu- ments. The comparable structure in laterally compressed specimens of L. alatum is what is here regarded as the pedicel alation. Except for the megaspore in Schopf’s specimens, the two types seem identical. If the lateral states of L. alatum had a single functional seed megaspore and an integu- ment, rather than the alated pedicel which is suggested in this paper, then they should be reassigned to Lepidocarpon sp. Alternatively the interpretation of the lateral state of L. alatum given above could be correct. And if the specimens described by Schopf are equivalent (i.e. they have an alated pedicel rather than an integument), both Schopf’s Lepidocarpon mazonensis and the specimens described here as Lepidostrobophyllum alatum must be reassigned to a different genus, describing sporophylls with a single functional megaspore and no integument, namely Acid amy do carp on Schumacker-Lambry (see below). Thirdly, L. alatum may not have contained a single functional megaspore, in which case the present separation of Lepidostrobophyllum alatum from Lepidocarpon sp. holds good. Cantheliophorus spp. Bassler. This genus (Bassler 1919), described as a ‘sporangiophoric lepidophyte’, has a similar appearance to the lateral states of L. alatum. Schopf (in Janssen 1940) refers Bassler’s genus to Lepidocarpon by reinterpreting the structure. This was supported later (Schopf 1941, p. 559) by the identification of single megaspores BOULTER: COMPRESSED LYCOPOD SPOROPHYLL FROM SOMERSET 455 from the illustrations in Bassler’s paper. Photographs of LepidophyJlum waldenburgensis in Nathorst (1914), refigured by Bassler (1919), have a similar character to the specimens of L. alatum which were laterally compressed. All the above-mentioned species are defined principally in terms of size and shape, factors which are variable in L. alatum. It is therefore suggested that the specimens described in this paper should not be assigned to any previously described species, unless characters of the pedicel structure and the compression process are investigated for these established species. Unfortunately, it is not possible to transfer or develop type specimens to obtain this information, and so full descriptions of species described over forty years ago are unlikely ever to be obtained. Inevitably, new species must be made which may overlap the ill-defined limits of other species. GENERIC ATTRIBUTION OF L. ALATUM The assignation of the species here described to Lepidostrobophyllum is uncertain, due to the absence of spores in all specimens. If the sporangium is ever found to have contained one or several tetrads of functional megaspores, or numerous microspores, then the assignation is correct. But if the sporangium contained only one functional megaspore, with three aborted spores (i.e. a Cystosporites type megaspore), then the sporophylls belong to a genus in the Lepidocarpaceae (Schopf 1941). In all specimens of L. alatum the sporangium wall extends beyond the pedicel alation (PI. 83, fig. 4; PI. 84, fig. 4); the alation could never have completely enclosed the sporangium. Therefore the genus Lepidocarpon is not applicable to any of the specimens dealt with here. Schumacker-Lambry (1966) describes a new genus of petrified lepidocarps as Achla- mydocarpon. The pedicel alation is not developed into an integument, rather, it is as in the species described here, or absent. The single functional megaspore is completely protected by a sporangium wall four layers thick. If a single functional megaspore is ever found in any L. alatum specimens, the generic assignation might be changed to Acldamydocarpon , though the identification of a thick sporangium wall in compression fossils presents a severe difficulty. It is evidence that Schumacker-Lambry’s genus Acldamydocarpon may be partially synonymous with Abbott’s genus Lepidocarpopsis ; they both represent ‘unintegumented lepidocarps’, although the evidence for lack of integument is clearer in Schumacker-Lambry’s petrified material than in Abbott’s compressions. In his reinterpretation of Bassler’s Cantheliophorus, Schopf refers it to the genus Lepidocarpon. He did not differentiate between the structures of an alated pedicel and a thick sporangium wall, so it is possible that some, if not all of the specimens described by Bassler are members of Acldamydocarpon. In this case, ‘the surface rugosity of the sclerotic cells on the integument surface’ (Schopf 1941, p. 560) would be cells of the thick sporangium wall, or of the pedicel alation. EVOLUTIONARY SIGNIFICANCE The structure of the pedicel in L. alatum represents an intermediate stage in the evolutionary change from the unprotected sporangium of many species of Lepidostrobus 456 PALAEONTOLOGY, VOLUME 11 sporophylls to the integumented Lepdidocarpon. Abbott (1963) has instituted a new genus for such an intermediate, Lepidocarpopsis, whilst Balbach (1966) suggests that the morphological changes involved are only of specific importance. The evolutionary tendencies towards specialization of the pedicel and sporangium structure in arborescent lycopod sporophylls seems to be more diverse than was pre- viously thought. The simplest Carboniferous Lepidostrobus had completely unprotected sporangia containing microspores and megaspores. Lepidostrobus takhtajanii , L. masleni, and Ach! dmy do carp on belgicum seem to represent one extreme in the evolutionary development with a single fertile megaspore protected by a thick sporangium wall and no integument, whilst Lepidocarpon , with a single fertile megaspore protected by an integument around the sporangium, represents another. The broad extent of the hairy alation in Lepidostrobophyllum alatum is in some degree morphologically intermediate between Lepidostrobus and Lepidocarpon. The lack of information concerning the thickness of the sporangium wall and the type of spores in these specimens make it impossible for us to appreciate the significance of the species, at the present. REFERENCES abbott, m. l. 1963. Lycopod fructifications from the upper Freeport (no. 7) coal in southeastern Ohio. Palaeontographica, B 112, 93-118. allen, k. c. 1961. Lepidostrobophyllum fimbriatum (Kidston 1883) from the Drybrook Sandstone (Lower Carboniferous). Geol. Mag. 98, 225-9. arber, a. 1914. An anatomical study of the Palaeozoic cone-genus Lepidostrobus. Trans. Linn. Soc. ser. 2, B 8, 205-38. arber, e. a. n. 1922. Critical studies of Coal-measure Plant-impressions. J. Linn. Soc. B 46, 171-217. balbach, m. k. 1 966. Paleozoic Lycopsid fructifications. II. Lepidostrobus takhtajanii in North America and Great Britain. Amer. J. Bot. 53, 275-83. bassler, h. 1919. A Sporangiophoric Lepidophyte from the Carboniferous. Bot. Gaz. 68, 73-108. brongniart, a. 1828. Prodrome d'un Histoire des Vegetaux Fossiles, p. 87. Paris. cridland, a. c., and williams, j. l. 1966. Plastic and epoxy transfers of fossil plant compressions. Butt. Torrey Bot. Club , 93, 311-22. crookall, r. 1966. Fossil plants of the Carboniferous of Great Britain. Mem. Geol. Surv., Palaeon- tology, IV, part 4. dix, e. 1934. The sequence of floras in the Upper Carboniferous, with special reference to South Wales. Phil. Trans. Roy. Soc. Edin. 57, 789-838. hirmer, m. 1927. Handbuch der Paldobotanik. R. Oldenbourg, Munich. janssen, r. e. 1940. Some fossil plant types of Illinois. Illinois State Museum, Sci. Pap. 1. Springfield, Illinois. jongmans, w. 1931. ‘ Namenandering bei Lepidostrobus' Jaarverslag over 1930, p. 91. Geol. Bureau, Heerlen. lesquereux, L. 1857. Fossil Plant Coalfields Pennsylvania. Boston J. Nat. Hist. 6, 4, 430. 1870. Fossil Plants of Illinois. Geol. Surv. of Illinois , 4, 2, 442. 1884. Description of the coal flora of the Carboniferous formation in Pennsylvania, vol. III. 2nd Geol. Surv. Pennsylvania, Report of Progress. nathorst, a. g. 1914. Zur fossilien Flora der Polarlander, i. 4. Nachtrage zur Paldozoischen Flora Spitzbergen, 110 pp. Stockholm. schopf, j. m. 1938. Two new Lycopod seeds from the Illinois Pennsylvanian. Trans. III. State Acad. Sci. 30, 139-46. 1941. Notes on the Lepidocarpaceae. Am. Midi. Nat. 25, 548-63. schumacker-lambry, j. 1966. Etude d'un cone de Lepidocarpaceae du houiller beige : Achlamydocarpon belgicum g. et sp. nov. Acad. Roy. Belgique Memoires, Coll. 4°, Deuxieme series, 17, 1. BOULTER: COMPRESSED LYCOPOD SPOROPHYLL FROM SOMERSET 457 scott, d. h. 1920. Studies in Fossil Botany. Part 1. Black, London. selling, o. h. 1944. Studies on Calamitean Cone Compressions by means of serial sections. Svensk Botanisk Tidskrift, 38, 295-330. snigirevskaya, N. s. 1964. An anatomical study of the plant remains from the Donets coalballs. I. Lepidodendraceae. Paleobotanika Akad. Nauk., U.S.S.R. Trudy, Ser. 8, Pt. 5, 1-36. [In Russian.] walton, j. 1923. On a new method of investigating fossil plant impressions or incrustations. Ann. Bot. 37, 379-91. 1936. On the factors which influence the external form of fossil plants; with descriptions of the foliage of some species of the Palaeozoic Equisitalian genus Annularia, Sternberg. Phil. Trans. Roy. Soc. Lond. B 226, 218-37. zeiller, r. 1886. ‘Bassin houil. de Valenciennes.’ Etudes Gites Min. France, 508 pp. (1888); Atlas (1886). M. C. BOULTER Department of Biology West Ham College of Technology Romford Road, London E. 15 Manuscript received 16 June 1967 FUNCTIONAL STUDIES ON THE CRETACEOUS OYSTER ARCTOSTREA by R. M. CARTER Abstract. The three Cretaceous species-groups Arctoslrea colubrina (Lamarck), A. ungulata (Schlotheim), and A. diluviana (Linnaeus) are described. Particular attention is paid to the detailed morphology of A. colubrina , from which its life history is reconstructed. It is inferred that the unusual characters of the genus (especially the arcuate shape, the zigzag commissure and the funnel spines) relate to the size and importance of the gills as primary food-collecting organs. The subspecies A. colubrina ricordeana (d'Orbigny) possesses, in addition to these special adaptations, a set of long spines secreted sub-parallel to the plane of the commissure on the lower valve; these are interpreted as a specific response to the hazards of inhabiting a substrate of soft ooze. Behavioural and structural adaptations used by Recent oysters for combating conditions of high turbidity are discussed, and it is suggested that similar methods were utilized by many extinct species. Funnel spines occur in other bivalves, including Pinna and Eiheria ; it is likely that they also are connected with inhalent current streams. Some taxonomic implications of the functional interpretation of Arctostrea are presented in an appendix. Many malacologists consider Recent Ostrea to represent the peak of bivalvian evolu- tion (e.g. Atkins 1938), and few would dispute that its integrated gill/palp complex represents a feeding organ of outstanding power and efficiency. It is one of the very few Recent bivalve genera for which a large amount of published material is available, owing this distinction mainly to its economic importance. It appears that such spec- tacular success as oysters have achieved has at least partly been made possible by their adoption of a mode of life involving cementation to a hard substrate. Yet because of their irregular shell form consequent upon such cementation, oysters have understand- ably always been the bete noire of most systematists, palaeontologists included. It is the purpose of this paper to apply some of the available information on Recent Ostrea to a study of a distinctive group of extinct oysters in an attempt to better understand their morphology, and hence to clarify their taxonomy. It has become apparent to me during this study that however well known the biology of common European and American oysters may be, there is still a large amount of primary research needed on most of the more exotic tropical oysters before our knowledge of the family can be claimed to be reasonably complete. The genus to be studied, Arctostrea, has its morphology exemplified by the species A. colubrina Lamarck, a common oyster in the Cenomanian rocks of Western Europe (PI. 85, fig. 5). Particularly noteworthy are the superbly developed zigzag commissure (PL 86, fig. 1), and the highly characteristic arcuate shape in plan view (PL 85, fig. 5). Other groups of oysters have zigzag commissures, and virtually all lineages of oysters from the Jurassic onwards have at some time exhibited a tendency to arcuate shape, but only a relatively small group of species ranging through the Jurassic and Cretaceous was able to combine these two trends into a successfully adapted working complex. To detail the biological and stratigraphical details of presently known populations of Arctostrea is far outside the scope of this paper. For the purposes of morphological description it will suffice to recognize three major infrageneric categories, conveniently termed species groups. They are the Arctostrea colubrina group (PL 85, fig. 5), the [Palaeontology, Vol. 11, Part 3, 1968, pp. 458-85, pis. 85-90.] CARTER: FUNCTIONAL STUDIES ON CRETACEOUS OYSTER ARCTOSTREA 459 A. ungulata group (PI. 87, fig. 8), and the A. diluviana group (PI. 85, fig. 1), and will be discussed in turn. Terminology. In order to facilitate unambiguous discussion of detailed ontogenetic changes, the terms proximal and distal are preferred to the more normal dorsal and ventral (text-fig. 4); the shell margins are termed the anterior and posterior arcs. The term ‘vertical zone’ is useful in description of shells with steep (though not necessarily rigorously vertical) slopes around the valve edge. A vertical zone may occur virtually all round the commissure (as in Arctostrea colubrina ), or may be limited to specific areas on the commissure (as in Arctostrea diluviana). It is the inevitable result of shell secretion along sectors of mantle edge where no mantle cell generation is taking place (see Carter 1967). Note on captions and material studied. The following abbreviations are used consistently in all plate and text figure captions: UMZC , University Museum of Zoology, Cambridge University; SM, Sedg- wick Museum, Cambridge University; BM, British Museum (Natural Elistory), London. It should be emphasized that this study is based upon the examination of specimens in these museum collections; statements as to the type of substrate inhabited by different populations of Arctostrea are, in the main, inferences made from the matrix that still adheres to many specimens. THE ARCTOSTREA COLUBRINA GROUP This species group may be recognized by its regularly arcuate shape (PI. 85, fig. 5), well-developed vertical zone, the numerous sharp zigzags around the commissure (PI. 88, fig. 5), and the presence of small tubular spines situated at the crest of each commissural zigzag (PI. 86, fig. 2); the growth track of these spines results in a characteristic divaricat- ing ornament pattern on the fiat shell top. The group was widely distributed, and com- mon, in Lower Cretaceous times, and is known from the Middle East, North Africa, Austria, Germany, France, Portugal, and England. A. colubrina ricordeana ( d'Orbigny ) The availability of specimens of ricordeana preserving every detail of their life history in the form of well defined growth-lines enables an accurate ontogeny to be recon- structed. Though this reconstruction is based on populations of colubrina ricordeana inhabiting the Lower Chalk Sea of England, the greater part of it applies with only minor modifications to other populations of colubrina s.s., and also with slightly greater modifications to members of the ungulata and diluviana species groups. The Chalk specimens of ricordeana are large, often very inflated, oysters with strong zigzag valve edges (PI. 86, fig. 1). Adult specimens may possess a dimension greater than 110 mm. in the plane of the commissure, and have a maximum transverse dimension (‘inflation’) of up to 70 mm. During growth, mantle expansion is at a maximum distally on the commissure, in a narrow arc here named the generative arc of the mantle edge (text-fig. 1); elsewhere on the commissure, outside of the generative arc, the introduction of new mantle cells is minimal, though not altogether absent. Thus shell secretion during the life of the animal results in the building of a vertical zone around the greater part of the commissure; this vertical zone is crossed by steep plicae that represent the tracks of earlier growth stages of the zigzag commissure (PI. 86, fig. 2). The generative arc is not only the site of localized generation of new epithelial cells, it is also the site of introduction of all new plicae (PI. 85, fig. 5). Any zigzag on the current 460 PALAEONTOLOGY, VOLUME 11 commissure can always be traced back up its own plica to the point of its inception as a gentle undulation in the generative arc of an earlier commissure (e.g. plica a, PL 86, fig. 1). h.a. * *1 — text-fig. 1 . Comparison of the shape of the gill of ( a ) Pecten and ( b ) Ostrea with the adult shell outline of four specimens (c-f in- clusive) of Arctostrea colubrina (Lamarck). All X i. All upper (right) valves viewed from above; stippled area represents the unornamented, pre-zigzag dissoconch, correspond- ing to the area of attachment on the left valve, h.a., hinge axis; /, ligament pit; g, generative arc of the mantle edge, arrow indi- cating direction of mantle expansion. Ontogeny. The significant stages in the life history of ricordeana are conveniently sum- marized in discrete steps: 1 . Spatfall on to available hard attachment sites ; commonly pieces of shell and echino- derm test. This was accompanied by immediate cementation by the surface of the left valve; by analogy with Recent oysters, it was followed in the next few days by metal morphosis from the veliger type organization to the adult anatomical form. 2. Continuing shell secretion eventually resulted in the size of the animal exceeding the size of the attachment object. At some stage after this the mantle edges changed CARTER: FUNCTIONAL STUDIES ON CRETACEOUS OYSTER ARCTOSTREA 461 from being roughly planar to being sharply zigzag, passing through stages from gently flexed to more sharply folded (PI. 86, fig. 4). This change in the morphology of the mantle edge may have been partly under muscular control, but it was probably also connected with localized epithelial generation. As shell secretion continued unabated, a clear record of these changes is shown by the successive growth-lines on the valve surface. In particular, there is always preserved an unornamented pre-zigzag dissoconch that is sharply demarcated from the following growth stages and gives a clear indication of the size and shape of the attachment object (PI. 88, fig. 3). 3. When the shell outgrew the area of cementation laterally, the mantle edges of the left (lower) valve on either side of the mantle isthmus secreted tubular spines which were attached in a root-like bundle either to the initial attachment object, or to closely adjacent objects (text-fig. 2). These root-spines were first secreted about the time of the inception of the zigzag commissure, and obviously served to attach the animal tightly to the substrate. They continued to be secreted so long as the animal was still attached to the substrate proximally. Some specimens do not progress beyond this ontogenetic stage. 4. At this stage in ontogeny, which occurred at very different shell sizes in different individuals, the shell either broke free from its proximal attachment or, if it was attached to a relatively small object, overbalanced, and hence came to lie free on the sea floor. Presumably stimulated by contact with the chalky ooze, the mantle edges of the left 462 PALAEONTOLOGY, VOLUME 11 valve secreted a set of long tubular spines at the crests of the zigzag of that valve (i.e. in the troughs of the commissure if the shell is viewed in the life position). These spines (PI. 85, fig. 4; PI. 87, figs. 6, 7; text-fig. 6) extended at right angles to the shell outline, were sub-parallel to the plane of the commissure, and sometimes half as long as the shell itself. They were only secreted along the anterior arc of the commissure. Occasional specimens (e.g. SM B6461) that did not progress beyond the attached phase of ontogeny (stage 3 above) were able to utilize their anterior arc spines as additional ‘roots’ or ‘props’ to further stabilize their attachment. In these cases, of course, the spines were not sub-parallel to the plane of the commissure, but became irregular, and were gener- ally directed downwards towards the surface of attachment. 5. In large, and presumably old specimens of ricordeana there was a slowing of the rate of introduction of new epithelial material in the generative arc, whilst over-all shell secretion continued. As a result extremely high vertical zones were built up. When this was the case, there was often secretion of further sets of spines around the commis- sure; on any one plica there may have been up to four successive spines, each vertically above (with respect to life orientation) its immediate predecessor (PI. 86, fig. 3). Other populations of colubrina The long spines sub-parallel to the plane of the commissure, diagnostic of A. colubrina ricordeana , appear to be restricted to very local populations from particular horizons in the Cenomanian Chalk Marl of England and Normandy. However, populations of the colubrina group agreeing with ricordeana in all other aspects are common from rocks of Cretaceous age. 1. The Haslingfield population. A rich fauna of Arctostrea is known from the locality of Haslingfield, a few miles south of Cambridge. Specimens from this population (PI. 85, fig. 5) are virtually indistinguishable from the Folkestone ricordeana , apart from their lack of long spines along the anterior arc. Occasionally broken spine stumps suggest that some specimens did secrete spines in this position, but the majority of the popula- tion possess very strong clumps of root spines proximally, and were clearly attached throughout their ontogeny. EXPLANATION TO PLATE 85 Fig. 1. Arctostrea diluviana (Linnaeus), Senonian, Essen, Germany, xl. Internal view of a typical upper (right) valve of this species; note the well-developed gill-gutter anterior to the adductor scar, with pustulose shell texture along its bottom. BM L61234. Fig. 2. Arctostrea angulata (Schlotheim), Maastrichtian, Maastricht. X 2. Internal view of an upper (right) valve to show the flat floor; note that the marginal zigzag is of amplitude equal to the depth of the valve. BM LL8600. Fig. 3. Arctostrea imgulata (Schlotheim), Upper Cretaceous, Buzi Valley, Portuguese East Africa. x2. This specimen shows well the characteristic flat-topped shell, and open crested plicae, that are so typical of the species group. BM L56928. Fig. 4. Arctostrea colubrina ricordeana (d'Orbigny), grey Chalk Marl, Folkestone, Kent. Approximately X 4. View of double valved specimen from below, looking up at the divaricating ornament and snowshoe spines of the under (left) valve. SM B.6451 ; specimen figured by Woods, text-fig. 122. Fig. 5. Arctostrea colubrina (Lamarck), Lower Chalk Marl, Haslingfield, Cambridgeshire. X 1. A specimen showing the characteristic divaricating plicae and branchiform shape of the upper (right) valve when viewed from above, g , present position of the generative arc (growing edge) of the mantle margin; g', a previous position of the generative arc. SM B.6559. Palaeontology, Vol. II PLATE 85 CARTER, Functional morphology of Arctostrea CARTER: FUNCTIONAL STUDIES ON CRETACEOUS OYSTER ARCTOSTREA 463 2. The Le Mans population. It was noted earlier that on the shell of an adult Arcto- strea any given plica starts its history in the generative arc of the mantle edge as a small undulation in the growth-lines. In many populations of the colubrina group, for example, that from Cenomanian calcareous sandstones of Le Mans, this may develop next into a marked, short, tubular ‘funnel spine’ (PI. 88, fig. 5), and only later grow into a sharp zigzag. The plica, of course, retains this funnel spine at its crest, and several more may be secreted in homologous positions as growth continues (PI. 88, fig. 4). It is striking that these funnel spines are extremely well developed on the anterior margin, but are only rudimentary on the posterior (PI. 86, fig. 2); they also tend to be rather better developed on the right (upper) valve. THE ARCTOSTREA UNGULATA GROUP Though the over-all shell morphology of A. ungulata is similar to A. colubrina, there are significant differences in detail. New plicae are again introduced distally in the generative arc, but the commissural zigzag is of much greater wavelength and amplitude than in colubrina. There is thus an apparent tendency for new plicae to be introduced in pairs on the anterior and posterior sides of the distal growing edge (PI. 85, fig. 3). The zigzag also has broad, rounded extremities in contrast to the sharp, pointed extremities of colubrina. The initial mantle edge reflection at the start of a new plicae (PI. 85, fig. 3; PI. 86, fig. 6) is broad and strong, and it retains a similar aspect throughout ontogeny. This, linked with the angle at which the mantle edge is held, results in the characteristic ‘open- crested’ plicae that typify the species. A similar open-crested morphology may be seen in the A. pusilla group (PI. 90, fig. 9). The growth lines on the shell of ungulata beauti- fully display the early stage of differential mantle expansion, during which an initially gently undulose commissure is transformed into the broad, strong mantle reflection at the head of each plica (PI. 86, fig. 6). However, after this initial stage of differential mantle expansion, further shell secretion results in the building up of a vertical zone in the usual fashion (PI. 85, fig. 3). It is likely that the differences in morphology between colubrina and ungulata are at least partly related to a difference in habitat, ungulata being usually collected from medium-grained or coarser calcareous sandstones, whilst colubrina is often common in finer grained sediments. The attachment scar is generally small and not well marked, and whilst some specimens may have been attached proximally throughout life, the majority probably lay free on the sea floor. The A. ungulata group is restricted to Upper Cretaceous rocks and has a wide dis- tribution, ranging from Crimea, Bulgaria, and East Africa, as far north as Holland, Belgium, and Scandinavia. It seems to replace the colubrina group at this stratigraphic level. There is little doubt that ungulata developed from an attached species similar to A. pusilla (Nilsson) (see appendix). THE ARCTOSTREA DILU VIA N A GROUP Members of this group have a larger and more solid shell than members of the previous two, this being a direct result of their being permanently cemented to the substrate throughout post-larval ontogeny. 464 PALAEONTOLOGY, VOLUME 11 It is characterized by its somewhat irregular, oval shape (PI. 85, fig. 1) and its large size. Although there may be a zigzag all around the commissure, it is always best developed over the anterior margins, and it is only here that there is any vertical zone built up. The diluviana group is in an Upper Cretaceous development, and seems to have a more limited distribution than ungulata or colubrina', specimens that I have examined all derive from northern Europe (Belgium, Holland, Germany, and Sweden). text-fig. 3. To illustrate the style of shell secretion in an equilateral bivalve; the growth increments (ruled area) on each valve are of wedge shape, tapering towards the hinge axis, h.a., hinge axis; v, ven- tral margin; 5, rate of secretion of shell material. CONSIDERATIONS ON SHELL FORM Fundamental tenets. Analysis of shell form in the Bivalvia becomes hopelessly confused unless a clear distinction is maintained between the basic phenomena of shell secretion per se on the one hand and generation of mantle cells (i.e. expansion of the area of mantle epithelium) on the other (Carter 1967). However, provided this distinction is maintained, it is possible to understand in detail the growth of even complex Bivalvia such as Arctostrea. EXPLANATION OF PLATE 86 Fig. 1. Arctostrea colubrina ricordeana (d’Orbigny), grey Chalk Marl, Folkestone, Kent. x2. Arrow marks the inception of plica ‘a’ in the generative arc; the shell is of a juvenile animal which had not secreted snowshoe spines at the time of its death. BM LL14751. Fig. 2. Arctostrea cf. colubrina (Lamarck), Cenomanian, Trouville, Normandy. X 2. Note the develop- ment of funnel spines (more conspicuous on the anterior than on the posterior margins) in this view of the posterior vertical flank. BM 65748. Fig. 3. Arctostrea colubrina ricordeana (d'Orbigny), Chalk Marl, Norman Cement Works, Cambridge. X 11. View of part of the vertical zone developed on the anterior arc of the lower (left) valve, showing up to four broken snowshoe spine bases (one above the other) on individual plicae. SM B.6557. Fig. 4. Arctostrea colubrina ricordeana (d'Orbigny), same specimen as 1. x4. Note the suppression of the more dorsal plicae. Fig. 5. Pinna rugosa J. de C. Sowerby, Recent, unlocated. X 11. View of the ventral valve edge to show the funnel spines. Note how the earlier spines are sealed off by the secretion of a sheet of shell across their base. UMZC 2008 (Saul Collection). Fig. 6. Arctostrea ungulata (Schlotheim), same specimen as Plate 85, fig. 2. X 4. View of the posterior arc, showing the characteristic open crested plicae and the high amplitude of the zigzag. Fig. 7. Arctostrea colubrina ricordeana (d’Orbigny), grey Chalk Marl, Folkestone, Kent. X 4. Showing a plica whose growth lines display a zigzag that is starting to diminish in amplitude (3) after an initial increase from its inception in the generative arc (1) up to a point of maximal amplitude (2). Plica is situated on the posterior arc of the upper (right) valve. BM L.80734. Palaeontology, Vol. 11 PLATE 86 CARTER, Functional morphology of Arctostrea WM. CARTER: FUNCTIONAL STUDIES ON CRETACEOUS OYSTER ARCTOSTREA 465 Practical implications. In a bivalve of simple equilateral form (e.g. Glycymeris), secretion of shell is at a maximum somewhere along the ventral margins, and decreases dorsally on either side of this around the lateral shell margins (text-fig. 3). By virtue of its most unusual pattern of mantle cell generation, Arctostrea (text-fig. 4) has its point of maximal shell secretion (Z) situated on the anterior arc of the shell, secretion diminishing at points around the commissure both distally and proximally to this; then there is a further proximal text-fig. 4. An idealized adult specimen of Arctostrea marking in selected growth-lines ( a through to e ) and plicae (1 through to 5). The growth stages corresponding to each growth-line are drawn separately as growth series b' to e' \ adductor muscle, solid black; h.a., hinge axis; v, visceral mass; gill, ruled lines; arrow marks functional centre of the inhalent current stream bathing the gill; feathered arrow centre of exhalent stream. Note how this point migrates relatively in a distal direction during growth. Z: point of maximal amplitude of zigzag on the anterior arc of the commissure, diminishing antero- proximally and antero-distally. Z’ \ point of maximal amplitude on the posterior arc of the commissure, diminishing postero-proximally and postero-distally. (It should be emphasized that growth-lines a to e and plicae 1 to 5 are only selected examples from the continuous series of growth-lines and plicae that occur on the shell surface; see also PL 85, fig. 5.) maximum of secretion (Z\ but note that secretion here is itself absolutely less than at Z) on the posterior arc, secretion diminishing both distally and proximally from this point as well. A further point of interest may be mentioned here: the hinge axis in Arctostrea is situated at the proximal end of the shell, and hinging takes place by rotation about this axis (text-fig. 4). However, the extension of the axis itself has a limiting effect upon the adult shell form of the animal for, if efficient hinging is to be maintained, it is impossible for the distal growing edge of the shell to transgress over the extension of the hinge axis. Individual specimens of Arctostrea show many modifications in late adult life to cope with this restriction, two of the commonest of which can be seen on text-fig. 1. Fig. 1 D. Description. The smooth widely umbilicate shells have whorls which vary in cross-section between sub-circular, oval and sub-rectangular; impressed area is always shallow (text-fig. 2). Rarely preserved growth-lines are almost rectiradiate faint undulations showing a weak lateral sinus (PI. 104, fig. 9); they have not been observed on the venter. Suture lines consist of an almost flat ventral saddle, broad, evenly rounded lateral lobes, moderately sharp umbilical saddles, and a widely open dorsal lobe. The umbilical saddle is usually crested forward of the ventral saddle; in several specimens this forward projection of the umbilical saddles is accentuated in late growth stages (text-fig. 2b). One sectioned specimen shows septa near the dorsum to be strongly deflected forward to meet the next succeeding septum, and apparently failing to form complete transverse partitions; this curious configuration may possibly have resulted from distortion. In some specimens (e.g. USGD 6846) the curvature of the lateral lobe is localized on the mid-flank, forming a very broad V. 542 PALAEONTOLOGY, VOLUME 11 Septal necks are short. Late growth stages usually show internal constrictions numbering up to 5 per whorl. text-fig. 2. Comparison of P. (P.) teicherti sp. nov. with P. (P.) pattisoni (M‘Coy) and Clymenia laevigata (Munster). a-j, P. (P.) teicherti sp. nov. x4. a-d, based on USGD 6890; a, whorl section at D = 24-3 mm.; b , the last suture, at D = 24 mm.; last but one suture, at D = 23-5 mm.; last but four suture at D = c. 20-5 mm. ; e,f based on USGD 6891 ; e, whorl section,/, external suture, both at D = c. 19 mm. ; g, h, based on USGD 6843 ; g, whorl section, h, suture, both at D = c. 15-5 mm. ; i,j, based on USGD 6862; /, whorl section,/, suture, both at D = 1 6-9 mm.; k-m, sutures of C. laevigata, k, copied from Sandberger, 1853, pi. 7, fig. 1 e-f; I, copied from Gumbel, 1863, pi. 16, fig. 5c, ‘Munster’s Original’; m, copied from Bogoslovsky et al. in Orlov 1962, fig. 187u; n, suture of C. placida Perna, copied from Perna 1914, fig. 79. o, whorl height plotted against diameter for P. (Pi) teicherti and P. (P.) pattisoni. p, umbilical width plotted against diameter for the same species. Specimens of P. (P.) teicherti represen- ted in these diagrams are indicated by USGD numbers at the left margin. Represented specimens ofP. (P.) pattisoni are indexed by letters a-g; a, holotype of pattisoni , dimensions from Selwood 1960, p. 165; b, c, P. placida from Perna 1914, p. 77; c, P. subnautilina from Petter 1960, p. 27; c1 from Petter 1960, pi. 5, fig. 20; d, e, P. quenstedti, from Wedekind 1914, p. 45; f, P. subnautilina schleizi from Muller, 1956, p. 74; G, C. subnautilina from Sandberger 1855, pi. 1, fig. If. JENKINS: FAMENNIAN AM MONOIDS FROM NEW SOUTH WALES 543 They are rectiradiate to rursiradiate, straight or with a slight forward concavity, and are formed by wave-like thickenings which may be symmetrical or have crests located forward of their mid-line. Constrictions are not detectable externally. Remarks. Smooth species of Platyclymenia are not always readily distinguished from species of Clymenia. Wedekind’s criterion of relative straightness of growth-lines in Clymenia is unsatisfactory since they are rarely preserved in either group and may be nearly straight in late growth stages of Platyclymenia (Schindewolf 1923, p. 462; Muller 1956, p. 74). More practical are the criteria proposed by Schmidt (1924, p. 124) based on sutural and septal form: the wider dorsal lobe and shorter septal necks of P. subnautilina in comparison with C. laevigata. Suture diagrams showing dorsal lobes which are about as deep as they are wide have been figured also for several other species of P. ( Platyclymenia ), (e.g. Perna 1914, figs. 71, 72, 77, 79 and Schindewolf 1934, hg. 12); in this characteristic they contrast sharply with the narrow parallel-sided dorsal lobe of Clymenia laevigata (e.g. Sandberger 1853, pi. 6, hg. 6; pi. 7, hg. 1 ; Gtimbel 1863, pi. 16, hg. 5c; Freeh 1902, text-hg. 4b). The relatively wide dorsal lobes and the short septal necks of the new species therefore assign it to Platyclymenia. Nominal species of Platyclymenia having oval or round cross-section and lacking ornamentation have been justihably synonymized by Selwood (1960) under the name P. ( P .) pattisoni M’Coy 1851 who also rehgured the holotype (pi. 26, hg. 10, not hg. 11 as stated in text). Perna’s species Clymenia placida also belongs, at least in part, to P. ( P .) pattisoni (cf. Schindewolf 1922, p. 124). But the new species described above differs from pattisoni in being more widely umbilicate and in its slower rate of increase of whorl height (text-hg. 2 o and p). In these characters of general form, and in the presence of constrictions, P. (P.) teicherti resembles C. laevigata, but in the present material it does not reach the size of the latter species. It could be regarded as a possible ancestor of Clymenia. Occurrence. Specimens of P. (P.) teicherti outnumber the total of all the other species in the fauna here described. Platyclymenia ( Platyclymenia ) alterna sp. nov. Plate 105, figs. 8-1 1 Derivation of name. Latin alternus, alternating, referring to the ribs. Holotype. USGD 6838, the only specimen found of this species. Preservation. Internal mould of segment of one whorl and the more complete external mould of two whorls, preserved in clayey sandstone. Dimensions (in mm.) D WW WH IJW Holotype, USGD 6838 23-5 4-5 c. 7 11 Diagnosis. A species of Platyclymenia with intercalated ribs and a subrectangular whorl section. Description. The holotype is widely umbilicate with a high subrectangular whorl section. It is slightly distorted and WW as measured may be somewhat reduced thereby. Whorls 544 PALAEONTOLOGY, VOLUME 11 overlap only slightly, the impressed area being about 1 mm. deep at WH = 7 mm. Whorl sides are parallel and the venter is weakly arched. Ribs are concave, projected and in two alternating sets, both ending at the ventro- lateral shoulder to leave the venter smooth. The ribs of the more prominent set are usually continuous across the flanks; they increase in height and width towards the venter and, to a lesser extent, towards the umbilical seam, so that the ribs are relatively weak in the middle portion of the whorl. Intercalated ribs are usually confined to the outer half of the whorl, but may occasionally appear weakly near the umbilical seam. Crests of adjacent ribs are about 1-5 mm. apart at WH = 7, the accentuated rib termina- tions being there about 0-5 mm. wide. Remarks. Two other specimens from the type locality having high subrectangular whorl sections and umbonally accentuated ribs may belong to this species (USGD 6835 and 6881); both represent a growth stage which is later than is preserved in the holotype and so cannot be confidently identified until more complete material becomes available. In whorl shape the species resembles P. ( P .) sandbergeri and P. (P.) walcotti, but the intercalated ribs of a/terna are distinctive. Family rectoclymeniidae Schindewolf 1923 Genus rectoclymenia Wedekind 1908 Type species. Rectoclymenia roemeri Wedekind 1908 by subsequent designation of Schindewolf 1957. Rectoclymenia ? sp. Plate 105, figs. 15-18 Description and remarks. Together with the other described species occur numerous fragments of a large coiled cephalopod. Its siphuncle has not been observed; the simple partial suture preserved on one fragment, the straightness of the low radial undulations on the whorl flanks and the general form suggest Rectoclymenia. A quite exceptionally large size is indicated. One short portion of a whorl which is incomplete ventrally measures 74 mm. in the radial direction, has a whorl thickness EXPLANATION OF PLATE 105 Figs. 1-4. P. {Platyclymenia) teicherti sp. nov. 1, 2, internal mould, USGD 6862, x2; 3, 4, internal mould, USGD 6843, X 2. 2, 4 are dorsal views of the detached incomplete body chambers showing septal face and internal suture. Figs. 5-7. Sporadoceras cf. rotundum Wedekind. 5, internal mould, USGD 6831, x3; 6, 7, internal mould, USGD 6832, x3. Figs. 8-1 1. P. ( Platyclymenia ) a/terna sp. nov. 8, 9, external and internal moulds of holotype USGD 6832, x2; 10, 11, the same, X 1. Figs. 12-14. Sporadoceras inflexion Wedekind. 12, part of impressed area of body whorl showing spiral striae, x3. 13, 14, lateral and ventral views, the latter showing the deformed venter, due to crushing; USGD 6866, X 1. Figs. 15-18 . Rectoclymenia ? sp. 15, small internal mould showing several partial sutures, USGD 6853, x 2; 16, part of an external mould of an outer whorl photographed to show low rectiradiate undula- tions, USGD 6899, xf; 17, portion of an outer whorl, USGD 6865, xf; 18, USGD 6864, X 1. Palaeontology, Vol. 11 PLATE 105 JENKINS, Famennian ammonoids from New South Wales JENKINS: FAMENNIAN AMMONOIDS FROM NEW SOUTH WALES 545 (internal mould) of 24 mm. and an impressed area 12-5 mm. deep; its umbilical margin forms an arc which approximately fits a radius of 70 mm. A diameter of about 300 mm. is thus indicated. Whorl section in late growth stages is compressed high oval, the very gently arched sides converging ventrally from the inner flanks to an evenly rounded unkeeled venter. Sculpture seemingly changes with growth from rectiradiate ribs, evenly rounded in profile and strongest towards the umbilicus, to low dorso-lateral bullae which pass radially into low rectiradiate swellings at a later growth stage. Faint radial growth lines parallel the ribs and also indicate a probable ventral sinus. Three small unribbed specimens may represent an early growth stage. One shows simple sutures with shallow rounded lateral lobes (PI. 105, fig. 15), is moderately evolute, and has the following dimensions (in mm.). R WH WW UW USGD 6853 12 6-7 c. 2 10 (WW may be reduced by compressive distortion). Closest of the described species of Rectoclymenia in general form to these small specimens is R. acuta (Schmidt), but none is recorded as attaining the exceptionally large size indicated for the material described above. According to Schindewolf (in Moore 1957) the genus is confined to the Platyclymenia Zone. It is previously recorded from Europe, Asia, and North America. Suborder goniatitina Hyatt 1884 Superfamily cheilocerataceae Freeh 1897 Family cheiloceratidae Freeh 1897 Genus sporadoceras Hyatt 1884 Type species. Goniatites bidens G. and F. Sandberger by original designation. Sporadoceras inflexion Wedekind Plate 105, figs. 12-14; text-fig. 3 a, b 1908 Sporadoceras inflexion Wedekind, p. 595, pi. 39, fig. 43; pi. 42, figs. 3, 2>a. [non] 1914 Sporadoceras inflexion Wedekind; Perna, p. 36, 98, pi. 1, fig. 14; text-fig. 21 [= A. muensteri var brachyloba Frecfi] 1918 Sporadoceras inflexion Wedekind; Wedekind, p. 149, text-fig. 47 d. 1959 Sporadoceras inflexion Wedekind; Petter, p. 274. Material. One specimen (USGD 6866), an internal mould, somewhat crushed. Diagnosis. A species of Sporadoceras in which the second lateral lobe is pointed and is two or three times as wide as the first lateral lobe which is shorter and skewed towards the venter. Dimensions {in mm.) D WW WH UW USGD 6866 (internal mould) 45-8 11+ 16 nil 546 PALAEONTOLOGY, VOLUME 11 Description. The mould shows a closed umbilicus, the whorls being completely over- lapping; the body chamber extends through three-quarters of a whorl. Whorl section is interpreted as having been approximately oval with an evenly rounded venter; crushing has resulted in a narrowed venter but the original rounded venter is preserved at some points. A detached fragment of the body chamber preserves a sector of its impressed area which clearly shows fine and regular spiral striae (about 30 per cm.). The suture is characteristic of the genus (text-fig. 3 a, b) with rounded saddles and more or less pointed lobes. The inner lateral saddle is pointed and asym- metrical with a straighter outer limb; the outer lateral saddle is shorter, deep and narrow, parallel sided, ending bluntly or sub-acutely with a slight outward direction. Remarks. The specimen described resembles hetero- lobatum Lange but its first lateral lobe lacks the consistent sharpness shown in Lange’s species, and its first lateral saddle (E/A1L) is relatively higher. The ventral deflection of the first lateral lobe in my speci- men, is less than in Wedekind’s figure of the type of S. inflexion , but this is judged to have been possibly diminished by the partial crushing which has con- verted an originally rounded venter to an acute form. S. muensteri var. brachyloba Freeh and S. pseudo- carinatum Petter are somewhat similar to my specimen in shell-form and suture but have a relatively wider first lateral lobe. Sporadoceras inflexion, is previously recorded from 111/3 in Germany and from an unrecorded locality and horizon in North Africa (Petter 1959). Sporadoceras cf. rotundum Wedekind Plate 105, figs. 5-7; text-fig. 3c 1908 Sporadoceras rotundum Wedekind, p. 594, pi. 39, fig. 21 ; pi. 42, fig. 1. 1918 Sporadoceras rotundum Wedekind; Wedekind, p. 148, text-fig. 47c. Two similar small distorted specimens of Sporadoceras are comparable with S. rotundum in form and suture-line. The widely rounded venter and the dorso-ventral direction of distortion by compaction indicate an originally sub-sphaerical form. A partial external suture (text-fig. 3c) on the smaller specimen shows a deep and pointed second lateral lobe and a shallow rounded first lateral; the umbilical sector is indistinct but probably as shown by the dashed line, whereas the ventral sector is not observable. Both specimens show three constrictions about 90° apart; they are rectiradiate on the flanks and show a constant shallow rounded ventral sinus. Sporadoceras biferum is similar in its suture-line and has a named variety ( sulcifera Lange) distinguished by text-fig. 3. Sutures of Sporadoceras. a, b, S. inflexum Wedekind, based on USGD 6866 at D = c. 47 ; x 2. c, 5. cf. rotundum Wedekind, based on USGD 6832 at D = c. 9; x4-25. JENKINS: FAMENNIAN AMMONOIDS FROM NEW SOUTH WALES 547 constrictions, but its shell-form is more compressed than that indicated for the specimens here recorded. S. rotundum is recorded from Zones Ilia and III/3 in Germany and the lower Prolobites Zone (= Ilia) in the Urals. 5. biferum is recorded from Zones II/3-I1I/3 in Germany and Zones II-IV in the Sahara. Acknowledgements. For discussions and information the writer thanks Professors B. F. Glenister and W. M. Furnish of Iowa State University, Iowa City, U.S.A., Professor M. R. Flouse of the University of Hull and Dr. E. B. Selwood of the University of Exeter, England. Financial assistance from the University of Sydney Research Grant is also acknowledged. REFERENCES Bogoslovsky, b. i. 1962. Rare type of ornamentation in clymenids. Paleont. Zh. 1962 (1), 166-8. [In Russian.] frech, F. 1902. Uber devonische Ammoneen. Beitr. Paldont. Geol. Ost.-Ung. 14, 27-1 1 1, pi. 2-5. glenister, b. f. and klapper, g. 1966. Upper Devonian conodonts from the Canning Basin, Western Australia. J. Paleont. 40, 777-842, pi. 85-96. gumbel, c. w. 1863. Ueber Clymenien in den Uebergangsbilden des Fichtelgebirges. Palaeontographica, 11, 85-165, pi. 15-21. hyatt, a. 1883-4. Genera of fossil cephalopods. Proc. Boston Soc. nat. Hist. 22, 253-338. jenkins, t. b. h. 1966. The Upper Devonian index ammonoid Cheiloceras from New South Wales. Palaeontology, 9, 458-63, pi. 72. lange, W. 1929. Zur Kenntnis des Oberdevons am Enkeberge und bei Balve (Sauerland). Abh. preufi. geol. Landesanst., n.f. 119, 1-132, pi. 1-3. m‘coy, f. 1851. On some new Devonian Fossils. Ann. Mag. Nat. Hist. (2) 8, 481-9. moore, R. c. (ed.) 1957. Treatise on Invertebrate Paleontology, Part L — Mollusca 4, Cephalopoda, Ammonoidea, xxii + 490. Geol. Soc. Amer. and Univ. Kansas Press. muller, k. j. 1956. Cephalopodenfauna und Stratigraphie des Oberdevons von Schleiz und Zeulen- roda in Thiiringen. Beih. Geol. Jb. 20, 1-93, pi. 1 and 2. munster, G. Graf zu. 1832. Uber die Planiditen und Goniatiten ini Ubergangskalk des Fichtelgebirges. 1-38, pi. 1-6. Bayreuth. 1834. Memoire sur les Clymenes et les Goniatites du calcaire de transition du Fichtelgebirge. Ann. Sci. naturelles, 2(2), Zool. 65-99, pi. 1-6. (A French translation of the preceding item.) orlov, y. a. (ed.) 1962. Principles of Palaeontology, Mollusca-Cephalopoda. 1. Moscow, 1-425, pi. 68. [In Russian.] perna, e. 1914. Die Ammoneen des oberenNeodevonsvom Ostabhang des Siidurals. Com. Geol.. Mem. S.N. 99, 1-114, pi. 1-4. [In Russian, German summary.] petter, g. 1959. Goniatites devoniennes du Sahara. Mem. Carte geol. Alger. Paleont. 2, 1-313, pi. 1-26. 1960. Clymenies du Sahara. Ibid. 6, 1-58, pi. 1-8. pickett, J. W. 1960. A clymenid from the Wockhaneria zone of New South Wales. Palaeontology, 3, 237-41, pi. 41. rzehak, a. 1910. Der Brunner Clymenienkalk. Zeit. mdhr. Landesmus. 10, 149-216, pi. 1-3. sandberger, g. and f. 1850-6. Systematische Beschreibung and Abbildung der Versteinerungen des rheinischen Schichtensystems in Nassau. 1-564, pi. 1-39 (Atlas). Wiesbaden. sandberger, g. 1853. Einige Beobachtungen uber Clymenien; mit besonderer Riicksicht auf die west- phalischen Arten. Verh. natarhist. Ver. preufi. Rhein. Westphal. 10, 171-216, pi. 6-8. 1855. Clymenia subnautilina n. sp., die erste und bis jetzt einzige Art aus Nassau. Jb. nassau. Ver. Naturk. 10, 127-36, pi. 1. schindewolf, o. h. 1922. Einige Randbemerkungen zu E. Perna's Abhandlung ‘Die Ammoneen des oberen Neodevon vom Ostabhang des Siidurals’. Senckenbergiana, 4, 185-96. 548 PALAEONTOLOGY, VOLUME 11 schindewolf, o. h. 1923fl. Beitriige zur Kenntnis des Palaozoicums in Oberfranken. Ostthiiringen und dem Sachsischen Vogtlande, 1, Stratigraphie und Ammoneenfauna des Oberdevons von Hof a. S. Neues Jb. Min. Geol. Paldont. 49, 250-357, 393-509, pi. 14-18. 19236. Entwurf einer natiirlichen Systematik der Clymenoidea. Centralbl. Min., Geo!., Paldont. 1923, 23-50, 59-64. 1934. Uber eine oberdevonische Ammoneen-Fauna aus den Rocky Mountains. Neues Jb. Min. Geol. Paldont. B. B. (B) 72, 331-50. schmidt, h. 1924. Zwei Cephalopodenfaunen an der Devon-Carbongrenze im Sauerland. Jb. preufi. geol. Landesanst. 44, 98-171, pi. 6-8. selwood, e. b. 1960. Ainmonoids and Trilobites from the Upper Devonian and Lowest Carboniferous of the Launceston Area of Cornwall. Palaeontology, 3, 153-85, pi. 26-29. teichert, c. 1941. Upper Devonian Goniatite Successions of Western Australia. Am. Journ. Sci. 239, 148-53. 1948. Middle Devonian goniatites from the Buchan district, Victoria. J. Paleont. 22, 60-7, pi. 16. voisey, a. h. and williams, k. l. 1964. The geology of the Carroll-Keepit-Rangari area of New South Wales. J. Proc. Roy. Soc. N.S. W. 97, 65-72. wedekind, r. 1908. Die Cephalopodenfauna des hoheren Oberdevon am Enkeberge. Neues Jb. Min. Geol. Paldont. 26, 565-634, pi. 39-45. 1913. Die Goniatitenkalke des unteren Oberdevon von Martenberg bei Adorf. Ges. Naturf. Freunde Berl. 1913, 23-77, pi. 4-7. 1914. Monographic der Clymenien des rheinischen Gebirges. Abh. Ges. Wiss. Gottingen, math.- physik. Kl. n.f. 10, 1-73, pi. 1-7. 1918. Die Genera der Palaeoammonoidea (Goniatiten), (mit Ausschlu/? der Mimoceratidae, Glyphioceratidae und Prolecanitidae). Palaeontographica, 62, 85-184, pi. 14-22. T. B. H. JENKINS Department of Geology and Geophysics The University of Sydney Sydney, N.S.W. Typescript received 24 October 1967 Australia A Q UILAPOLLENITES IN THE BRITISH ISLES by A. R. H. MARTIN Abstract. Species of Aquilapollenites Rouse emend. Funkhouser are present in the interbasaltic lignites of Shiaba and Bremanoir, Mull, Scotland. The bearing of this on the age of the lignites is discussed; it is believed that this may be pre -Eocene. The genus Aquilapollenites Rouse (1957) emend. Funkhouser (1961) has been regarded as an important indicator pollen fossil of the uppermost Cretaceous, Palaeocene, and Lower Eocene of parts of the Northern Hemisphere, at least Western and Plains states of North America and North-west and Central Siberia. Funkhouser (1961) and Stanley (1961) refer to the lack of records of the genus from western Europe. However, at about the same time, a posthumously published paper by J. B. Simpson (1961) described the genus Taurocephalus as new and ascribed to it the single species T. proteus , which was regarded as an extinct member of the Proteaceae. This very characteristic pollen is clearly Aquilapollenites , but as the description differs somewhat from the published description of Aquilapollenites , a brief explanation is offered. Simpson apparently saw in polar view, the fine furrows which in Aquilapollenites have been described as demicolpoids, but interpreted these as a proximal-face triradiate mark such as he believed to occur in some Proteaceae. The two polar lobes of the grain (PI. 106, fig. 6), superimposed one over the other, appear to have been taken for a single distal pore resembling the three equatorial protrusions which were also described as pores. These were correctly described as differing from all extant Proteaceae in that ectexine covered the supposed pores. Simpson’s own photographs (1961, pi. xii, figs. I, 2, and 5, for example) show that two polar protrusions are present. Taurocephalus is described as sometimes having four equatorial arms; this is the only way of reconcil- ing the appearance of Simpson’s figs. 1 and 5 with his description since these are equa- torial views. However, his plate xii, fig. 1 1 is a specimen with four true equatorial arms seen in polar view. The specimen has been re-located and re-photographed (PI. 106, figs. 1-3). This feature had not been observed before in Aquilapollenites. Four-armed grains are probably of low incidence. In examining the material from Mull, in which Aquilapollenites occurs, a second species and a few grains of a third were observed. Indeed Taurocephalus proteus is likely to be a compound of two of these species (see taxonomic section). To find these in sediments believed to be of Miocene or Pliocene age (the interbasaltic lignites of Shiaba and Bremanoir) suggests that the dating of the lignites should be reconsidered. An age between Maastrichtian and Lower Eocene seems most likely. The reliance on high percentages of coryloid, quercoid, and Hamamelidaceae-type pollens seems misplaced when it is realized that these pollens have a very extended time range. Manum (1962), for example, reports many such grains in the Tertiary of Vestspitsbergen, which on zoo- logical grounds is thought to be late Palaeocene to Eocene. British phytogeographers, reluctant to accept the Mull interbasaltics as contemporary with the London Clay [Palaeontology, Vol. II, Part 4, 1968, pp. 549-53, pis. 106. 550 PALAEONTOLOGY, VOLUME 11 flora, would probably be sympathetic towards the earlier rather than the later end of the range in time indicated by AquilapoIIenites i.e. uppermost Cretaceous rather than Eocene. Several other pollen grain types present in the interbasaltic beds are consistent with the earlier dating here proposed. These include Classopollis sp. (PI. 106, figs. 15-16), which though uncommon, does not show any obvious signs of redeposition and, in lignite, is less likely to be re-deposited. The known range is Triassic to Eocene but it is uncommon after the mid-Cretaceous. Haloragis bremanoirensis Simpson is conspecific with Nudopollis thiergartii Pflug, an indicator of the Palaeocene in Germany, though ranging from the Maastrichtian to Lower Eocene. A search for AquilapoIIenites on slides, prepared by Simpson, of the pre-basaltic lignite of Ardslignish, Ardnamurchan, failed to reveal any. All the slides quoted in the paper of 1961 were re-examined. Most were in fair condition, with pollen grains still stained by safranin, though some had developed small crystals in the glycerine jelly mountant. SYSTEMATIC DESCRIPTION angiospermae Incertae sedis Genus aquilapollenites Rouse emend. Funkhouser (April 1961) Taxonomic synonyms. Parviprojectus Mtchedlishvili (July 1961); Taurocephalus Simpson (December 1961). AquilapoIIenites spimdosus Funkhouser ( Taurocephalus proteus Simpson pro parte) Plate 106, fig. 7 Sice range (10 grains). Polar axis 35-43 p (av. 38 p)\ tips of equatorial projections to polar axis 19-25 p (av. 23 p); (ratio 0-5-0-67). Location. Bremanoir and Shiaba interbasaltic lignites, Mull, Scotland. Known time range. Palaeocene-Eocene (Funkhouser 1961); Maastrichtian (Srivastava 1966). AquilapoIIenites attenuatus Funkhouser ( Taurocephalus proteus Simpson pro parte) Plate 106, figs. 1-6 Size range (5 grains). Polar axis 28-42 p (av. 36 p)\ tips of equatorial projections to polar axis 21-24 p (av. 23 p)\ (ratio 0-5-0-8). Four-armed grains are of rare occurrence. Location. Bremanoir and Shiaba interbasaltic lignites. Mull, Scotland. Known time range. Maastrichtian (Funkhouser 1961, Srivastava 1966). Distinction. The characters held to distinguish A. spimdosus from A. attenuatus are the shorter equatorial protrusions, the absence of puncta and the scattering of spinules over the whole surface instead of being grouped. In the Scottish material, while it is possible to find specimens corresponding to the two forms, there are a good many which combine features of both (text-fig. 1). The term ‘punctate’ could be applied to 70% of the Mull specimens, including both long and short-armed ones and to some with scattered MARTIN: AQUILAPOLLENITES IN THE BRITISH ISLES 551 as well as those with grouped spinules. It is too early to suggest that a single polymorphic form is represented, but clearly there is a need for further study of the variability shown by these forms. Ratio: Length of projection tip to polar axis Length of polar axis text-fig. E Number of Aquilapollenites spinulosus and A. attematus grains falling into different polar length/equatorial radius classes. Crossed squares = punctate grains. R indicates grain with spinules restricted to parts of surface. Aquilapollenites pachypolus nom. nov. Plate 106, figs. 8-14 Nomenclatural synonym. Parviprojectus striatus Mtchedlishvili 1961, pi. 73, fig. 1 a-c. The genus Parviprojectus is a later homonym of Aquilapollenites as emended by Funk- houser and within that genus the name striatus is preoccupied by A. striatus Funkhouser. It is therefore necessary to give a new name to this very distinctive form. A translation of Mtchedlishvili’s diagnosis and description of Parviprojectus striatus follows: Diagnosis. Pollen grains rather large sized, tricolpate. Body broadly ellipsoidal in form, equatorial projections short, colpi resembling cracks. Exine thick, pilate, striate, but having thickenings on the polar areas. 552 PALAEONTOLOGY, VOLUME 11 Description. Length of grain 46-0-46-6 p; diameter of body 27-4-31-0 p, length of equatorial projections 9-7 p, their width 7-7-9'0 p. Pollen grains tricolpate. Outline of body in equatorial view broadly ellipsoidal. Exine thickness about 2-Op. Nexine layer only slightly thinner than sexine layer, tapering from the proximal end of the projection. Sexine consisting of pila which have grown together to form comb-like lirae. In sectional view are seen the thinner parts of the columns and the round heads. The lirae are situated at a slight angle to each other and sometimes branch, forming the characteristic striate structure of the grain. At the poles the sexine layer becomes considerably wider, forming thickened polar areas which have a corrugated structure. On the equatorial projections the sexine layer becomes considerably thinner. Colour of grains brown. Size range of Scottish specimens {5 grains). Polar axis 35-43 p (av. 40 p); equatorial axis 28-33 p ( c . 38 p tip to tip of protrusions); thickness of the polar area 4-5-6 0 p. Location. Shiaba interbasaltic lignites, Mull, Scotland. Known time range. Maastrichtian-(?) Danian (Samoilovich and Mtchedlishvili 1961). Distinction. The grains are distinguishable from A. nntrus Stanley, and A. bertillonites Funkhouser by the thickened polar areas. They are partially distinguishable by their generally larger size and shorter projections. The colpoids are very narrow, rather in- distinct, and in one specimen (PI. 106, fig. 10), pores appear to be present. The size of Scottish specimens is smaller than in the two Siberian specimens. EXPLANATION OF PLATE 106 Magnification of all figures, X 1,000. Figs. 1-3. Aquilapollenites attenuatus Funkhouser ( Taurocephalus proteus Simpson pro parte). Specimen with four equatorial arms, also showing demicolpi; slide Bremanoir 5 a (= Simpson 1961, pi. xii, fig. 11). 1, High focus, polar area. 2, High focus, exine pattern, showing baculi (puncta) and spines. 3, Low focus, exine pattern (LO-pattern). Figs. 4-6. Aquilapollenites attenuatus Funkhouser. 4, Equatorial view, showing well-defined double pattern of spines and baculi (LO-pattern) ; slide Shiaba 29 a. 5, Equatorial view of small grain showing typical attenuatus shape, but lacking distinct LO-pattern; slide Bremanoir 5a. 6, Polar view; slide Bremanoir 5a. Fig. 7. Aquilapollenites spinulosus Funkhouser ( Taurocephalus proteus Simpson pro parte). Equatorial view, showing scattered spines and lack of distinct LO-pattern; slide Bremanoir 5a. Figs. 8-14. Aquilapollenites pachvpolus nom. nov. ( Parviprojectus striatus Mtched.). 8, Equatorial view, showing striate pattern, short projecting arms, and thickened polar area; slide Shiaba 1 a. 9, Surface and median view to show thickening of both ectexine and endexine in polar area; slide Shiaba 22 a. 10, Small grain, with apparent pore-apertures on equatorial arms, showing some convergence towards A. bertillonites Funkhouser; slide Shiaba 22 a. 11, Oblique view, sculpture of polar area insulous, colpoid visible in facing projection; slide Shiaba 1 a. 12-13, Successive views through specimen with upward projecting arm, showing the continuation of the striate pattern over the narrow indistinct colpoid, thus demonstrating that the colpoid is restricted to the endexine. 14, The same, focused on pattern of lower surface of grain; slide Shiaba 36/. Figs. 15-16. Classopollis sp. High and low focus of striate exine surrounding pore, which faces to right; slide Shiaba 1 a. All photographs from original slides prepared by J. B. Simpson, and housed in Scottish Geological Survey Office, Edinburgh. Palaeontology , Vol. 11 PLATE 106 MARTIN, Aquilapollenites from Scotland MARTIN: AQUILAPOLLENITES IN THE BRITISH ISLES 553 Acknowledgements. I gratefully acknowledge the receipt of an honorary associateship at University College, London University, the hospitality of Dr. W. G. Chaloner in offering the facilities of his section in the Botany Department, and the kindness of the Scottish Geological Survey Office, Edinburgh in making freely available the slides and materials left by Dr. J. B. Simpson. REFERENCES funkhouser, j. w. 1961. Pollen of the genus AquilapoIIenites. Micropaleontology, 7, 2, 193-8. manum, s. 1962. Studies in the Tertiary flora of Spitsbergen, with notes on Tertiary floras of Ellesmere Island, Greenland and Iceland. Norsk Polarinstitutt Skrifter, Oslo, 125. samoilovich, s. and mtchedlishvili, n. d. 1961. Pollen and spores of western Siberia, Jurassic to Palaeocene. Trudy V.N.I.G.R.Inst., Leningrad, 177, 657 pp. [In Russian.] simpson, j. b. 1961. The Tertiary pollen-flora of Mull and Ardnamurchan. Trans. Roy. Soc. Edinburgh , 64, 16, 421-68. srivastava, s. k. 1966. Upper Cretaceous microflora (Maestrichtian) from Scollard, Alberta, Canada. Pollen et Spores , 8, 2, 497-552. Stanley, e. a. 1961. The fossil pollen genus AquilapoIIenites. Pollen et Spores 3, 2, 329-52. A. R. H. MARTIN School of Biological Sciences University of Sydney Typescript received 28 October 1967 Australia DISCOHELIX (ARCHAEOGASTROPODA, EUOMPH ALACEA) AS AN INDEX FOSSIL IN THE TETHYAN JURASSIC by J. WENDT Abstract. Morphological features, growth, mode of life, and stratigraphical distribution are examined in more than 300 extremely well-preserved specimens of Discohelix from the Sicilian Lower and Middle Jurassic. In some species parabolic ribs similar to those in ammonites are formed owing to resorbtion during brief dis- continuities in growth. Deviated peristomes recurring at an angle of more or less 72 0 and resulting in a penta- gonal outline of the shell, are developed in two species for which the new subgenus Pentagonodiscits (type species D. (P.) angusta n. sp.) is established. Ten species of Discohelix sensu stricto are described, five of which are new: D. dictyota, D. conica, D. centricosta , D. costata, and D. levis. The short stratigraphical ranges of the represen- tatives of the genus are proved here for the first time and may be useful for the division of cephalopod-free series in the Tethyan Jurassic. Few genera of the widely distributed Palaeozoic Euomphalacea crossed the Permo- Triassic boundary and the superfamily became extinct in the Upper Cretaceous. In contrast to its Palaeozoic representatives, amongst which Knight (1934, p. 145) was even able to recognize some with value as index forms, those in the Mesozoic have neither been investigated in quantity nor in relation to their stratigraphical occurrence. Both these omissions can now be corrected using an individually and specifically rich gastropod fauna from the Jurassic of Sicily, of which members of the genus Discohelix are the chief element. The accompanying ammonites have usually allowed all the specimens to be zonally dated, so far as strong condensation permits such accuracy. A total of more than 300 Discohelix, extremely well preserved and with complete shells, have been collected from several Jurassic sections in western Sicily (Rocca Busambra). The only drawback is the recrystallization of the shell material which has left its struc- ture difficult to recognize. Several specimens from the Liassic of the Northern Cal- careous Alps, the originals of Stoliczka (1861) from the north alpine Hierlatz facies, and those of Quenstedt (1884) and Brosamlen (1909) from the south German epi- continental Jurassic have been available for comparison. Acknowledgements. I am indebted to Professor Dr. A. Seilacher (Tubingen) for his interest in the work and for critical reading of the manuscript, to Professor Dr. O. H. Schindewolf (Tubingen) and Dr. E. L. Yochelson (Washington) for stimulating suggestions and discussions. The loan of specimens from the Geologische Bundesanstalt Vienna and the Bayer. Staatssammlung Pal. Hist. Geol. Munich was kindly arranged by Professor Dr. R. Sieber and Dr. D. Herm. The fieldwork was made possible by the financial support of the Deutsche Forschungsgemeinschaft. Finally I would like to thank Dr. D. A. B. Pearson for his translation of the German text, and W. Wetzel (Tubingen) who prepared the photographs. The specimens labelled Ga 1347/1-56 are deposited in the Tubingen Geological Institute collection. MORPHOLOGY Protoconch Because of the more or less planispiral coiling of the shell, the larval whorls are almost always preserved, so that the sculpture of the earliest ontogenetic stages can be precisely [Palaeontology, Vol. 11, Part 4, 1968, pp. 554-75, pis. 107-10.] WENDT: DISCOHELIX (ARCH AEOG ASTROPOD A, EUOMPH ALACEA) 555 followed. After the nucleus, which is always smooth, two or three fine spiral lines appear on the upper and lower sides in the majority of the investigated individuals. After one or two whorls they disappear even if a renewed spiral ornament occurs later on the teleo- conch (PI. 107, fig. 16). In only three species is the protoconch completely smooth (PI. 107, fig. 15). A boundary between the proto- and teleoconch can neither be traced through a particularly accentuated interruption in growth (varix), nor through a sudden onset of the adult sculpture. The number of larval whorls can therefore not be deter- mined. When the shell is biconcave the protoconch is commonly slightly raised (text-fig. 3o), and when convexo-concave a little depressed (text-fig. 3c). Teleoconch Simple ribs. After one or two whorls, ribs appear more or less contemporaneously with the onset of clear growth-lines. They consist at first of weak swellings in the keel region, but with progressive growth they become enlarged as periodic ridges reaching to the suture on the upper and lower side, and running parallel with the growth-lines. On the interior of the shell this sculpture is either totally unrefiected or only barely detectable, so that steinkerns of strongly ribbed forms may be perfectly smooth. As the growth-lines are not noticeably denser on the ribs themselves than between them, they are more likely to have been produced by periodically increased shell deposition than by a deceleration of growth. The number and strength of the ribs can vary considerably and is therefore of only restricted specific value. Parabolic ribs. Long known in certain ammonites (especially perisphinctids) the signifi- cance of so-called parabolic ribs has often been discussed. In both ammonites and gas- tropods they are morphogenetically distinguishable from simple ribs. In form they are similar to simple ribs, but the course of the growth-lines is different. At regular intervals a growth-line (parabolic line) swings backwards in the keel region forming a parabolic notch and truncating the earlier growth-lines (text-fig. 1a). This is even more clearly seen on shells with a spiral sculpture since the sculpture itself is also cut off in the area of the parabolic notch. To some extent the spiral elements curve in sympathy with the notch and thus hint at the following discontinuity in growth (text-fig. 1b). The para- bolic notch may be developed as a shell ridge (D. a/binatiensis), or as a node ( D . cf. gumbeli ), or may take the form of a longitudinally open spine ( D . ( P .) angusta). In the last case the growth-lines are practically continuous. The truncation of earlier growth-lines and the development of special sculpture features show that the normal growth of the shell was interrupted at these points. The finger-like protrusions of the mantle concomitant with the parabolic nodes or notches possessed the ability, on the one hand for resorbtion, and on the other, for shell secretion As is often the case with the siphon it may have projected beyond the calcareous envelope which it depositated. In normal shell formation the parabolic notch was first closed before new material was added to the shell margin. Only then did a marked cessation of growth occur, since: firstly, in D. (P.) angusta the end of a strong inner varix (former peristome) does not lie immediately beneath the parabolic notch but shortly mouth- wards of it; secondly, the existing peristome was only formed after the infilling of the parabolic notch (however, in this final growth stage the last 2-3 ribs are often simple); and thirdly, a clear concentration of growth-lines owing to decelerated growth is only C 5934 O O 556 PALAEONTOLOGY, VOLUME 11 visible immediately after the parabolic line. In this area shell damage which affected the earlier peristome is often found (text-fig. 1b). text-fig. 1. Parabolic nodes and course of the growth-lines from the suture to the middle of the outer side (raised parts of the shell in black), a. Discohelix (£>.) cf. albinatiensis Dumortier, Ga 1347/14, X 13. b. Discohelix ( Pentagonodiscus ) angusta n. sp., with healed injury in a former peristome, Ga 1347/50, X 12, 5. From this it can be concluded that the parabolic notch may not itself be equated with a former aperture, in other words, a halt in growth (of course, strictly speaking each growth-line, and therefore also the parabolic line, represents a temporary shell margin). It is rather a part of a not wholly continuous process of shell secretion between two protracted interruptions in growth, although locally the contrary, namely resorbtion, occurs. Observation of recent gastropods has shown that shell secretion between two varices occupies only a few days and that a long pause follows thereafter, in which the peristome is only strengthened by internal varices or ribs (Fretter and Graham 1962, p. 64). For this reason gastropods in which the shell margin lies within the rapid growth phase are rarely found. The same growth process explains why a parabolic line has never been observed as the final peristome in ammonites. As the extension of the mantle withdrew the parabolic notch which it had produced was infilled, a process that bearing in mind the speed of growth in gastropods could probably have been completed within a few hours. Only later was growth interrupted, which in perisphinctids is com- monly indicated by deep constrictions equivalent to the inner varices of gastropods. It remains unknown whether the temporary finger-like extensions of the mantle had a WENDT: DISCOHELIX (ARCH AEOG ASTROPODA, EUOMPHALACEA) 557 physiological function in addition to the formation of parabolic notches and hollow spines. Nothing is known of this in gastropods either. The interpretation of the parabolae of perisphinctids given by Schairer (1967, p. 26) is incorrect. He proposed as an explanation that ‘the shell at certain points was more slowly deposited than at others’ (author’s translation). If such were the case the growth- lines would be of different density around the curve of the parabolic notch, but this is not so. Moreover, from Schairer’s fig. 11, one can see that on the parabola the sculpture (in this case ribs as the author had only steinkern material) is indeed truncated. To my knowledge only Schindewolf (1934, p. 342, text-fig. 9) has demonstrated that the parabolae of ammonoids (here Upper Devonian clymenias) originated through local resorbtion. Spiral ornament. Beginning with an intermediate growth-stage, commonly on the last 2-3 whorls, fine spiral lines are formed as result of a folding of the mantle edge. They appear first in the keel region and gradually encroach onto the sides, and together with the growth-lines and ribs produce a retiform sculpture typical of many gastropods. Their number and strength may vary considerably between specimens of equal dimen- sions, and spiral lines are therefore of only restricted value in the differentiation of species of Discohelix. For example in the case of D. albinatiensis they can be either equally spaced over the visible part of the whorl or only developed in the keel region, and thus show a transition to specifically inseparable shells without any spiral ornament (D. cf. albinatiensis). Peristome (text-fig. 2). Owing to the pure calcareous facies from which the whole of the material was collected, in no specimen could the peristome be completely cleared and the structures on the inside of the shell made visible. They could only be investigated in section, and the various types are best illustrated when cut in a longitudinal direction. In the simplest case the shell margin can only be recognized through a marked concentration of the growth-lines, without any alteration of the peristome profile (e.g. D. dictyota , text-figs. 2a, 3e). Probably not to be specifically separated from this type is a peristome which is gradually thickened by the addition of new shell substance on the inner side, and which therefore appears somewhat narrowed in cross-section (e.g. D. albinatiensis, D. cornea, D. cf. crenulata — text-figs. 2b, 3b, d, g, c). It may be preceded on the last whorl by one or two similar deposits formed during a pause in growth, but their position is random (text-figs. 2c, d). The broadened, trumpet-like peristome is strengthened by an inner varix on the outer lip (e.g. D. levis, text-fig. 2e, 3l), and the aperture may also be narrowed by an inner varix on the inner lip (e.g. D. cotswoldiae, text-fig. 2f), The highest degree of specialization is achieved when these last two aperture types are repeated at regular intervals ( D . ( P .) reussii and D. (P.) angusta — text-figs. 2g, h). Cross-section (text-fig. 3). A cross-section passing exactly through the protoconch illus- trates best the considerable variety of form within Discohelix. The mean spiral angle may lie between the wide boundaries of 125° to 220°, and the mean umbilical angle between 65° and 145°. The comparison of these two angles in one specimen, and the weak trochospiral coiling of the protoconch, show that perfect bilateral symmetry is not 558 PALAEONTOLOGY, VOLUME 11 achieved in Discohelix. Transitions exist between a slightly concave, a plane, and a convex spire, so that this character cannot always be given specific significance. D. albinatiensis shows the greatest variability in this way and may occur with a plane (sometimes even slightly concave) to a clearly convex spire (text-figs. 3b, a). In one and the same individual text-fig. 2. Types of peristomes (longitudinal sections), x3, 1. A. Discohelix (D.) dictyota n. sp., Ga 1347/4. b. Discohelix (£>.) albinatiensis Dumortier, Ga 1347/12. c. Discohelix (D.) conica n. sp., Ga 1347/28. d. Discohelix (D.) conica n. sp., Ga 1347/29. e. Discohelix (£>.) lev is, n. sp., Ga 1347/41. f. Discohelix (D.) cotswoldiae (Lycett), Ga 1347/48. g. Discohelix ( Pentagonodiscus ) reussii (Hornes, Ga 1347/56. h. Discohelix ( Pentagonodiscus ) angusta n. sp., Ga 1347/51. the degree of involution may change so that the cross-section loses its axial (mirror- image) symmetry. Septa. Septa closing off the uninhabited parts of the shell were shown by Knight (1934, p. 140) to be typical for the Euomphalidae. However, although many longitudinal sections were prepared for the species investigated here, septa were observed in only a few cases. They were found in the first 2-3 whorls but showed no regularity in either number (3-15) or distance from one another (text-fig. 4b). In one specimen of D. (P.) angusta three septa were found lying against the inner varices towards the end of the penultimate whorl (text-fig. 4a). This very late deposition of septa is surprising and permits no generally applicable conclusion about the length of the visceral mass. text-fig. 3. Cross-sections, x 3, 1. A. Discohelix ( D .) albinatiensis Dumortier, Ga 1347/9. b. Disco- helix (D.) albinatiensis Dumortier, Ga 1347/7. c. Discohelix ( D .) conica n. sp., Ga 1347/24. d. Discohelix (D.) cf. albinatiensis Dumortier, Ga 1347/15. e. Discohelix (D.) dictyota n. sp., Ga 1347/2. f. Discohelix (D.) cf. crenulata (Moore), Ga 1347/19. G. Discohelix ( D .) cf. crenulata (Moore), Ga 1347/23. h. Discohelix ( D .) cf. calculiformis Dunker, Ga 1347/42. i. Discohelix (D.) centricosta n. sp., Ga 1347/31. k. Discohelix (£>.) costata n. sp., Ga 1347/35. l. Discohelix ( D.) levis, n. sp. Ga 1347/37. m. Discohelix (D.) cotswolcliae (Lycett), Ga 1347/45. n. Discohelix ( D.) cf. giimbeli Ammon, Ga 1347/49. o. Discohelix ( Pentagonodiscus ) angusta n. sp., Ga 1347/52. p. Discohelix ( Pentagonodiscus ) reussii (Hornes), Ga 1347/55. 560 PALAEONTOLOGY, VOLUME 11 text-fig. 4. Discohelix ( Pentagonodiscus ) angusta n. sp. a. Ga 1347/54, longitudinal section with three septa towards the end of the penultimate whorl (inner whorls not preserved), X3-2. b. Ga 1347/51, inner whorls with septa, x 25. GROWTH In smooth shells no clearly marked interruptions in growth can be read from the density of the growth-lines. Therefore, the formation of the shell must have been more or less continuous. The same applies to shells with simple ribs, but those with parabolic ribs grew discontinuously, as previously discussed (p. 555). Similarly the regular inner varices following one another at intervals of 69° to 75° (mean 71-5°) in D. ( P .) reussii, indicate periodic phases of shell deposition between long pauses (text- fig. 5). text-fig. 5. Discohelix ( Pentagonodiscus ) reussii (Hornes), Ga 1347/56, longitudinal section, x6-3. WENDT: DISCOHELIX ( ARCHAEOG ASTROPODA, EUOMPHALACEA) 561 These two growth-rhythms, expressed in the one case by parabolic ribs and in the other by inner varices, occur together and interfere with one another in D. ( P .) angusta. As in D. ( P .) reussii the inner varices exhibit pentagonal symmetry (means of 704°, 70-7°, and 76° in three measured specimens), and between them three (on inner whorls sometimes two) parabolic ribs are intercalated. These ribs tend to arrange themselves in groups of three towards the last whorl. In this way the first formed parabolic rib of such a group coincides with an inner varix (text-figs. 6a, 7). (The outer varices, seen externally, cannot be exactly transferred into the longitudinal section; their position may therefore be slightly different.) text-fig. 6. Discohelix ( Pentagonodiscus ) angusta n. sp. A. Ga 1347/53, longitudinal section, x 6 3 (arrows indicate outer varices), b. Ga 1347/51, extrapolation of the normal logarithmic spiral (hatched line) from the regularly coiled inner whorls in comparison to the actual spiral (unbroken line), x 6. Thus if one considers the parabolic features as brief discontinuities and the inner varices as drawn out breaks in shell secretion, then D. ( P .) angusta shows complicated double-phased growth, the control of which is unknown. The subgenus Pentagonodiscus assumes a pentagonal outline in the last 2-2 f whorls. In text-fig. 6b the normal logarithmic spiral of the regularly coiled inner whorls has been extrapolated omitting the various irregularities of the shell (inner varices, ribs). This clearly shows that the aberrant shell form commences with the first inner varix and becomes more accentuated as the varices increase in strength towards the aperture. The most prominent deviations from a logarithmic spiral occur at the points of periodic 562 PALAEONTOLOGY, VOLUME 11 constrictions of the former peristomes. In the intervening growth-phases the spiral is again followed and thus shows itself to maintain control of the pentagonal coiling. Among the gastropods, there appear to be only two cases of angular instead of regular spiral coiling: Triangularia paradoxa Freeh and Tremanotus polygonus Barrande. The first has a triangular, and the latter a pentagonal outline. Nothing is known of the internal structure of either of these two species. Periodically constricted apertures were probably also formed and might be revealed by inner varices. text-fig. 7. Discohelix ( Pentagonodiscus) angusta n. sp. Ga 1347/53. Diagrammatic view of the intervals of the inner and outer varices on the last whorl (n = no. of outer varices, counted from the centre). Deviations from a logarithmic spiral are widely distributed among ammonoids, but commonly affect only the last body chamber. Triangular coiling in early ontogenetic stages, producing an extremely aberrant shell form, is found solely in a few clymenids {Soliclymenia, Wocklumeria, Parawock/umeria ) and goniatites ( Paralegoceras ) (see Schindewolf 1937, p. 110 et seq.). It is conceivable that periodic narrowing of the aperture was also responsible. Deep constrictions, signifying a former aperture at the points of strongest deviation from the normal spiral, occur in two of these genera ( Wocklumeria , Parawocklumeria ) and support this supposition. A narrowed body chamber in early ontogenetic stages, remaining unresorbed, and periodicity in growth lead automatically to an angular instead of a spiral shell form. If one imagines further shell growth after the last known peristome of e.g. Oecoptychius refractus (Reinecke), an angular outline must necessarily be the result. Not to be confused with such genuine WENDT: DISCOHELIX (ARCH AEOGASTROPOD A, EUOMPHALACEA) 563 angular forms are those in which the umbilicus has a polygonal outline as a result of deep constrictions (e.g. Morphoceras, some perisphinctids). In longitudinal section they have a perfectly regular logarithmic spiral. No attempt has been made here to fully explain the phenomenon of triangular am- monoids. The control of such aberrant growth remains as speculative as in polygonally coiled gastropods. MODE OF LIFE The planispiral coiling of many Discohelix species is reminiscent of ammonoids and suggests adaption to a nektonic mode of life. Thus Koken (1897, p. 42) proposed that the Upper Triassic KokeneUa (Pleurotomariacea), which is first coiled in a trochospiral and later in a planispiral, was adapted to a free swimming in deep sea. However, no recent gastropod with an almost symmetrical shell is known to be capable of active swimming. A single exception are the pelagic Heteropoda (e.g. Atlanta) with a weakly calcified, very thin or wholly vestigial shell. They cannot be compared with the thick- shelled, often strongly sculptured Discohelix , for which only a benthonic mode of life can be considered. If the planispiral coiling were an adaption to a nektonic mode of life, then this feature would normally tend to be realized as a result of selection. It turns out, however, to be extremely variable, not only on the specific but also on the generic level. The disc shape, in fact, would appear to be a disadvantage since it is much more susceptible to water movement than a helicocone. This can scarcely have been com- pensated for by the narrow whorls and the small aperture, indicating a correspondingly small and weak foot. For this reason the most favourable environment for Discohelix may have been in quiet sublittoral water or in the shelter of reefs rather than in a dis- turbed littoral. In a vertical position the pentagonal form can be seen to have an advantage over the spiral form, since the angularity provides a kind of resting surface (cf. Murex ) allowing the animal a greater stability on the sea-floor. The efficiency of this is increased through the support of lateral spines in D. ( P .) angusta. DEVELOPMENT AND VALUE AS INDEX FOSSILS There are still great gaps in our knowledge of the development of the Euomphalidae in the Triassic. For this reason the exact origin of the genus Discohelix has not yet been elucidated. Several Triassic forms described as Euomphalus belong to Woehrmatmia, but the majority are insufficiently known for an exact generic determination. Typical biconcave shells with more or less bilaterally symmetrical whorls and upper and lower keel appear for the first time in the lowermost Liassic. A greater variety of forms are known from the Sinemurian, especially from the Northern Calcareous Alps (Hierlatz-facies), amongst which are some with a plano-concave cross-section. The transition from biconcave to convexo-concave shells probably occurred several times independently, as has been observed in D. albinatiensis (p. 558). A special off-shoot of pentagonally coiled forms (subgenus Pentagonodiscus) also commenced in the Sine- murian but with only two stratigraphically isolated species, it has not yet been fully verified. The observed variation of individual features reduces their phylogenetic value and makes it necessary to separate species according to character combinations. Despite 564 PALAEONTOLOGY, VOLUME 11 the approximately forty known (many insufficiently) Discohelix species their lines of evolution can only be traced at a few points. Their stratigraphical distribution has not been proved because of their sparsity, occurrence as fissure-faunas, and the failure of accompanying index fossils. Therefore the well dated material presented here assumes a special stratigraphical value. From text-fig. 8 it can be seen that the species have only a very short range, extending through only one or a few zones. Similar conclusions can be drawn from the stratigraphical information given by Hudleston (1892, pp. 315-21) for the English Middle Jurassic. D. dictyota D. a/binatiens/s D. cf a/binatiens/s ! 1 D. conica D, centr/costa D. costa ta 1 -s? 1 D.cfca/cu/iformis D.(P.)angusta -cs 1 27 50 52 28 > 100 7 13 9 6 n 17 1 Malm Oxford/an u L 1 1 ■ ■ ■ Dogger Calloi/ian U M L Bathonian U M L Bajocian U M 16 T 1 L 28 1 ■ ■ 1 1 ■ 1 1 1 s 1 ■ I B ■ ■ L r ■ ■ ■ i Aa/enian U L Lias Toarcian U ~120 ■ 1 T M 148 T T 1 1 r 1 L text-fig. 8. The stratigraphical distribution of Discohelix in the condensed Jurassic of Rocca Busambra (Western Sicily); numbers indicate number of specimens. The strong condensation in the sections made it necessary to use substages instead of zones for the time-scale in text-fig. 8. In three cases the boundaries of the substages could not be drawn since up to five zones are inseparably concentrated in one strati- graphical unit (Wendt 1965, p. 300). For this reason the exact range of some of the species cannot be given. The representatives of Discohelix which begin the fossil record in the Middle Toarcian (Bifrons Zone) may have originated earlier. The absence of species from the base of the Bathonian to the Upper Callovian is owing to a stratigraphic gap characteristic of the west Sicilian Jurassic. A diminution of the Discohelix fauna is already evident earlier, and the last representative on the Dogger-Malm boundary is almost coincident with the extinction of the genus in the Lower Oxfordian. WENDT: D1SCOHELIX (ARCH AEOGASTROPODA, EUOMPH ALACEA) 565 It should not be overlooked that the stratigraphical range of the species investigated here is based on a series of sections from a single locality. To date it has been verified in two cases by specimens from the Northern Calcareous Alps. Furthermore well-dated material of Discohelix is wanted to complete and, if necessary, correct the ranges of the various species. Only two imperfectly preserved examples were available to Dunker when he estab- lished the genus Discohelix in 1847, so that even he himself was uncertain of its indivi- duality five years later (see footnote in Reuss 1852, p. 116). Rich material allowed Stoliczka (1861, p. 180 et seq.) to distinguish Discohelix from Euomphalus and to define it precisely for the first time. Without altering this concept of the genus Cossmann (1915, p. 133 et seq.) included in it both Triassic and Cretaceous species, although they did not completely pass within his diagnosis. In the same publication he proposed the genus Colpomphalus (1915, p. 136) for forms with a convex spire transitional to Disco- helix. This gradation is in fact so continuous, even in one and the same species (see p. 558), that Colpomphalus can no longer be maintained as an independent unit. Among the Triassic predecessors of Discohelix, Amphitomaria Koken differs through its prosocyrt growth-lines on the outer side, and Anisostoma Koken through its peculiar, deflected peristome. (Specimens placed in the latter genus, from which the aperture has been broken, could, in fact, be early members of Discohelix.) Type species. D. calculiformis Dunker 1847. Diagnosis. Widely to extremely widely umbilicate representatives of the Euomphalidae, with numerous, not overlapping whorls of trapezoidal cross-section, with an upper and lower keel. Mostly dextral, rarely sinistral. Spire concave, plane or convex. Protoconch homeostrophic, not sharply separated from the teleoconch. Growth-lines on the upper and lower sides prosocline, on the outer side opistocyrt. Ornament collabral and/or spiral. Peristome more or less tangential to foregoing whorl, its outline parallel to the axis of coiling, simple or with thickened outer and/or inner lip. Operculum unknown. Distribution. Triassic? Jurassic (Hettangian to Oxfordian). Synonym. Colpomphalus Cossmann 1915. SYSTEMATIC DESCRIPTIONS Genus discohelix Dunker 1847 Discohelix ( Discohelix ) dictyota n. sp. Plate 107, figs. 1-8; text-figs. 2a, 3e Derivation of name, dictyotus = retiform. Holotvpe. Ga 1347/1. Plate 107, figs. 1-4. Ga 1347/1 Ga 1347/2 Ga 1347/3 No. dp d h/d ha/wa nw nt sa ua 200 0-29 1 -49 9 34 0-28 1-33 8 (30) 0-29 1-78 8 33 18-3 16-4 214° 145° 566 PALAEONTOLOGY, VOLUME 11 Diagnosis. Dextral, spire concave; protoconch spirally ornamented, teleoconch with retiform sculpture; peristome unthickened. a text-fig. 9. Shell measurements, a = axis of coiling, dp = maximum diameter measured at peri- stome (in mm.), /; = height of shell, ha — height of aperture, p = cross-section through peristome (hatched), sa = mean spiral angle, ua = mean umbilical angle, wa = width of aperture. Additional abbreviations in the measurement-tables: d = diameter (in mm.), nt = number of tubercles (nodes), numbers in ( ) referring to half whorls, nw = number of whorls. Description. Fine ribs appear after the protoconch contemporaneously with the onset of clear growth-lines. They develop on the keel as nodes, and with good preservation may be recognized on the last whorls as a parabolic feature. On this last whorl the spiral lines, 7-10 on the upper and lower sides and 16-20 on the outer side, are strongest; they fail on the inner whorls of the teleoconch. The final peristome shows no special structure. Before it, on two specimens, a short pause in growth is indicated by a gentle angulation in the shell outline. Discussion. D. reticulata Stoliczka shows a similar sculpture, the spiral lines being restricted to the outer side, however, the whorls are more compressed, and broader than high in cross-section. D. orbis (Reuss) can be distinguished by its much finer ornament and clearly differentiated keel, and D. dunkeri Moore by its fewer nodes and plano-concave cross-section. Distribution. Middle Toarcian (Bifrons Zone). Discohe/ix ( Discohelix ) albinatiensis Dumortier 1874 Plate 107, figs. 9-15; Plate 108, figs. 8-10; text-figs. 2b, 3a-b 1874 Discohelix Albinatiensis Dumortier, p. 284, pi. 59. figs. 3-5. ? 1892 Straparol/us pulchrior Hudleston, p. 318, pi. 25, fig. 9. EXPLANATION OF PLATE 107 Figs. 1-4. Discohelix (£>.) dictyota n. sp., holotype. Ga 1347/1, X2. Figs. 5-8. Discohelix (D.) dictyota n. sp., Ga 1347/3, X 2. Figs. 9-12. Discohelix (D.) albinatiensis Dumortier, Ga 1347/6, X3. Figs. 13-14. Discohelix (D.) albinatiensis Dumortier, Ga 1347/8, x3. Fig. 15. Discohelix (D.) albinatiensis Dumortier, Ga 1347/11. Inner whorls with protoconch, X 20. Fig. 16. Discohelix (D.) cf. albinatiensis Dumortier, Ga 1347/14. Inner whorls with protoconch, X 20. Palaeontology, Vol. 11 PLATE 107 WENDT, Discohelix from the Tethyan Jurassic WENDT: DISCOHELIX (ARCHAEOG ASTROPOD A, EUOMPEIALACEA) 567 Lectotype (here selected). Dumortier 1874, pi. 59, figs. 4-5. No. dp d Ipd ha/wa nw nt sa ua Ga 1347/5 15-3 0-29 1-55 8 (16) 184° 120' Ga 1347/6 14 2 0-32 1 43 8 27 Ga 1347/7 14-2 0-30 1-48 8 (13) 188° 117' Ga 1347/8 13-2 0-42 1 -22 8 25 Ga 1347/9 14-4 0-48 1-47 9 (18) 132° 81 Ga 1347/10 12-4 0-33 1-63 8 12 Diagnosis. Dextral, spire weakly concave, plane, or weakly convex; protoconch smooth, retiform sculpture on teleoconch; peristome thickened. Description. The majority of the available specimens are plano-concave, several speci- mens, however, have a slightly concave or a more or less strongly convex spire. The larval whorls are free of sculpture, with an angulation near the suture, slightly concave upper side, and a smooth keel. The ornament of the teleoconch, beginning after 1 \-2 whorls, is similar to, but somewhat coarser than in D. dietyota. 20-27 parabolic nodes may be counted on the keel. 10-20% fewer nodes are present on the lower keel, thus emphasizing the asymmetry of the shell. Seven examples varying in diameter between 7 and 13 mm. stand apart from the remainder of the material in that the nodes are more widely spaced (12-15 instead of 20-23 on typical individuals, see PI. 108, figs. 8-10). A clear spiral ornament begins only after a diameter of 10 mm. has been reached. It may completely fail on the upper and lower sides and be restricted to a small area adjacent to the keel on the outer side. The peristome is internally thickened and therefore narrowed. Discussion. The larval whorls and peristome were not preserved in the material used by the authors cited in the synonymy, so that the present specimens can only be named with some reservation. Those with the most raised spire cannot be distinguished in cross- section and sculpture of the teleoconch from D. exsertus (Hudleston 1892, p. 320, pi. 26, figs. 3-4). Distribution. Middle Toarcian (Bifrons Zone). The type was thought by Dumortier to have come from the Opalinum Zone, but was found in a loose block of uncertain stratigraphic origin. According to Hudleston (1892, p. 319) D. pulchrior occurs rarely in the Murchisonae Zone. Discohelix ( Discohelix ) cf. albinatiensis Dumortier 1874 Plate 107, fig. 16; Plate 108, figs. 1-7; text-figs. 1a, 3d ? 1935 Discohelix albinatiensis Dumortier; Roman, p. 32, pi. 6, figs. 1, 1 a. No. dp d hid lia/wa nw nt sa ua Ga 1347/13 22-9 0-37 1-40 9 29 Ga 1347/14 210 041 1 96 9 30 Ga 1347/15 19-4 0-39 1-52 9 (16) 171° 99' Ga 1347/16 16-3 0-47 1-73 8 (14) 165° 90 Description. In contrast to typical members of the species spiral sculpture is totally lacking on the last whorls, but 2-3 spiral lines may be traced on the upper side of the 568 PALAEONTOLOGY, VOLUME 11 protoconch. From these two characters a decisive separation from D. albinatiensis is not possible. Distribution. Middle Toarcian (Bifrons Zone). Discohelix ( Discohelix ) cf. crenulata (Moore 1867) Plate 108, figs. 11-16; text-figs. 3f, g cf. 1867 Solarium crenulatum Moore, p. 90, pi. 4, figs. 19-20. No. dp d hid ha/wa nw nt sa ua Ga 1347/19 14 4 0-46 1-41 9 (30) 148° CC O Ga 1347/20 1 3 3 0-38 1-42 8 50 Ga 1347/21 12-8 0-52 1 -50 54 Ga 1347/23 7-7 0-57 1-36 7 147° 76° Description. The shell form of this dextrally coiled species varies from almost plano- concave ( sa = c. 170°) to clearly convexo-concave. The mean umbilical angle varies correspondingly between relatively wide boundaries. The protoconch has distinct spiral lines, and after \b-2 whorls fine ribs appear which rise to small nodes on the keel. Later, spiral lines reappear so that the sculpture is similar to that of D. dictyota and D. dun- driensis (Tawney), as Fludleston (1892, pi. 26, fig. 2) figured in detail. However, the external spiral lines are seldom regularly distributed over the whole outer side of the last whorl, and a smooth band is left free in the centre. In the available material two varieties differing in their whorl number can be distin- guished: one having 7 whorls at a maximum of approximately 8 mm. in diameter, and the other 8-9 whorls at 13-15 mm. in diameter. The aperture of the small examples is thickened (text-fig. 3g) and that one of the large ones is simple and unthickened (text- fig. 3f). Discussion. Moore’s figure and description do not give a clear impression of this species and it has not since been mentioned. Therefore no comparison of the shell cross-sections can be made, but aside from this the present specimens have a much wider umbilicus. A similar umbilicus is possessed by D. crussoliensis Roman (1935, p. 102, pi. 3, figs. 5, 5a), but this species is considerably more conical. D. helenae (Dumortier 1874, p. 141, pi. 36, figs. 1-4) has a comparable shell shape but the nodes on the keels are coarser and more widely spaced (18 in contrast to 50-60 in D. cf. crenulata). Distribution. Middle Toarcian (Bifrons Zone). The type was said to be from the upper Middle Liassic. EXPLANATION OF PLATE 108 Figs. 1-2. Discohelix (D.) cf. albinatiensis D urn order, Ga 1347/13, x2 Figs. 3-4. Discohelix (£>.) cf. albinatiensis Dumortier, Ga 1347/18, x2, Figs. 5-7. Discohelix (D.) cf. albinatiensis Dumortier, Ga 1347/17, X2, Figs. 8-10. Discohelix ( D .) albinatiensis Dumortier, Ga 1347/10, X 3. Figs. 11-13. Discohelix (D.) cf. crenulata (Moore), Ga 1347/21, X 3. Figs. 14-16. Discohelix (D.) cf. crenulata (Moore), Ga 1347/22, x3. Figs. 17-21. Discohelix ( D .) conica n. sp., holotype, Ga 1347/25, x3. Fig. 21. Discohelix (D.) conica n. sp., Ga 1347/27, x3. Palaeontology, Vol. 11 PLATE 108 WENDT, Discolielix from the Tethyan Jurassic WENDT: DISCOHELIX (ARCHAEOGASTROPODA, EUOMPHALACEA) 569 Discohelix ( Discohelix ) conica n. sp. Plate 108, figs. 17-21; text -figs. 2c-d, 3c Holotype. Ga 1347/25. Plate 108, figs. 17-20. No. dp hid hajwa nw so UCl Ga 1347/24 91 0-64 1 -52 1 124° 63° Ga 1347/25 9-3 0-65 1-42 7 Ga 1347/26 10-7 0-53 1 31 7 h 132° 74° Diagnosis. Small, dextral, spire convex; protoconch spirally ornamented, on teleoconch only growth-lines, smooth upper and finely tuberculated lower keel ; peristome thickened. Description. Three spiral lines can be traced on the depressed protoconch and disappear after 1-1| whorls. From approximately the fourth whorl fine nodes appear close to the suture, stretched out in the direction of the growth-lines but not reaching the smooth upper keel. The lower keel is formed by a succession of fine tubercles which are accom- panied on the outer side by 2-3 undulating spiral lines. The externally thickened peri- stome can sometimes be seen to have been preceded on the last whorl by 1-2 growth interruptions (text-figs. 2c-d). Discussion. D. conica contrasts with the above species in having a higher spire. Further- more the failure of spiral ornament and the smooth upper keel are good specific charac- ters. Distribution. Middle to lower Upper Toarcian (Bifrons to c. Thouarsense Zone). Discohelix ( Discohelix ) centricostan n. sp. Plate 109, figs. 1-7; text-fig. 3i Holotype. Ga 1347/30. Plate 109, figs. 1-3. No. dp d hid ha/wa nw sa ua Ga 1347/30 230 0 31 1 51 8 Ga 1347/31 25-6 0-35 1-29 9 217° 136° Ga 1347/32 18-4 0-32 1-75 8 Diagnosis. Dextral, spire concave; protoconch spirally ornamented, nepionic whorls ribbed, neanic whorls smooth; peristome trumpet-like. Description. The shell consists of 7-9 whorls and the spire is almost as depressed as the umbilicus resulting in a nearly symmetrical cross-section. The 2-4 whorls following the spirally ornamented protoconch have weak ribs and finely tuberculate keels. These elements become less accentuated with growth. The last whorls are completely smooth or show only traces of a slight undulation. The aperture, which is only preserved in two cases, is broadened and thickened to an almost circular cross-section in the larger specimen (text-fig. 3i), and in the smaller (Ga 1347/32) is similar in cross-section to the preceding whorls. 570 PALAEONTOLOGY, VOLUME 11 Discussion. Without knowledge of the protoconch this species cannot be separated from D. calculiformis Dunker. Distribution. Middle to lower Upper Toarcian (Bifrons to c. Thouarsense Zone). Discohelix ( Discolielix ) costata n. sp. Plate 109, figs. 12-18; text-fig. 3k Holotype. Ga 1347/34. PI. 3, figs. 15-18. No. dp d hid ha/wa nw nt Ga 1347/34 24-8 0-42 1-83 9 29 Ga 1347/35 250 0-38 1-46 8 (13) Ga 1347/36 19-8 0-32 1-42 8 27 Diagnosis. Dextral, spire concave; protoconch spirally ornamented, teleoconch ribbed; peristome thickened, somewhat broadened. Description. 4-5 spiral lines are present on the upper and lower side of the protoconch, but are quite absent from the following whorls. They are ornamented by regular ribs which may be traceable in the nepionic stage or only on the last whorl. The outer side, between the strongly corrugated keels, is completely smooth except for fine growth-lines. Discussion. The biconcave shell form, the failure of spiral lines, and the strong sculpture make D. costata a very distinctive species. A phylogenetic connection with D. centri- costa seems probable, D. costata having originated from this species through a persis- tence of the ribbing to the last whorl. Distribution. Upper Toarcian (c. Variabilis to Levesquei Zone). Discohelix ( Discohelix ) lev is n. sp. Plate 109, figs. 8-11 ; text-figs. 2e, 3l Derivation of name, levis = smooth. Holotype. Ga 1347/38. Plate 109, figs. 8-9. No. dp d h\d ha/wa nw sa ua Ga 1347/37 22-7 0-38 1-29 8 211° 124° Ga 1347/38 19-7 0-36 1-51 8 Ga 1347/39 15-6 0-33 1-30 Ga 1347/40 19-8 0-37 1-59 8 EXPLANATION OF PLATE 109 Figs. 1-3. Discohelix (D.) centricosta n. sp., holotype, Ga 1347/30, X2. Figs. 4-6. Discohelix (£>.) centricosta n. sp., Ga 1347/32, X 2. Fig. 7. Discohelix (D.) centricosta, n. sp., Ga 1347/33, X 3. Figs. 8-9. Discohelix (D.) levis n. sp., holotype, Ga 1347/38, X 2. Figs. 10-11. Discohelix (D.) levis n. sp., Ga 1347/40, X 2. Figs. 12-14. Discohelix (D.) costata n. sp., Ga 1347/36, x2. Figs. 15-18. Discohelix (D.) costata n. sp., holotype, Ga 1347/34, x2. Palaeontology, Vol. 11 PLATE 109 WENDT, Discohelix from the Tethyan Jurassic WENDT: D1SCOHELIX ( ARCHAEOG ASTROPOD A, EUOMPH ALACEA) 571 Diagnosis. Dextral, spire concave; protoconch and teleoconch smooth; peristome broadened and thickened. Description. The smooth shell is only ornamented with growth-lines, which become stronger towards the keel causing it to be corrugated. The slight asymmetry of the cross-section is seen in the greater depression of the umbilicus relative to the spire, and in the irregular curvature of the outer side. In three specimens the trumpet-like aperture is preserved and is reinforced by an inner varix on the outer lip. Discussion. The shape is very similar to D. excavata (Reuss) without being as high as this species {lt\d = 043-0-48 in Stoliczka’s originals). Apart from this the keel in D. levis remains smooth as far as the peristome, whilst in D. excavata more or less strong nodes appear on the last half-whorl. The sculpture of the protoconch of this last species is unknown. Distribution. Aalenian to Lower Bajocian (condensed Opalinum to Sauzei Zone). Discohelix ( Discohelix ) cf. calculiformis Dunker 1 847 Plate 110, figs. 1-4; text-fig. 3h cf. 1884 Discohelix calculiformis Dunker; Quenstedt, p. 325, pi. 197, figs. 32, 33, 35. No. dp d hid ha/wa nw set net Ga 1347/42 16-1 0-36 1 -80 «n m O m .) cf. giimbeli Ammon, Ga 1347/49, x3. Fig. 17. Discohelix ( Pentagonodiscus ) angusta n. sp., Ga 1347/53, X 3. Figs. 18-21. Discohelix (Pentagonodiscus) angusta n. sp., holotype, Ga 1347/50, x2. Figs. 22-24. Discohelix ( Pentagonodiscus ) reussii (Hornes), lectotype, Geol. Bundesanst. Vienna 2980, x 2. Palaeontology, Vol. 1! PLATE 110 WENDT, Discohelix from the Tethyan Jurassic WENDT: DISCOHELIX (ARCHAEOG ASTROPOD A , EUOMPHALACEA) 573 Discohelix (Discohelix) cf. gumbeli Ammon 1893 Plate 1 10, figs. 13-16; text-fig. 3n cf. 1893 Discohelix Gumbeli Ammon, p. 215, text-fig. 39. No. dp h/d ha/wa nw nt sa ua Ga 1347/49 140 0-38 1-40 6\ 17 222° 142° Description. That the perfectly biconcave shell is dextral, can only be seen in the some- what asymmetrical outline of the peristome. The spiral lines of the protoconch disappear after \ \ whorls with the onset of strong ribs which swell to parabolic nodes on the keel. Spiral ornament reappears only on the last two whorls and remains restricted to the keels leaving a smooth band between them on the external side. The peristome, in con- trast to the growth-lines, lies almost in a shell radius and has thickened margins. Discussion. A comparison with the holotype must be restricted to its figure, since the specimen itself, formerly in the Bayer. Staatssammlung Pal. hist. Geol. in Munich, has been lost. From this figure it appears that D. gumbeli has higher whorls (h/d is said to be 0-5) and a more asymmetrical cross-section. D. sapplio (Orbigny) is similarly ornamented, but is plano-concave. Distribution. Upper Callovian/Lower Oxfordian (together with Distichoceras bicostatum (Stahl)). The holotype is from the Lias-Dogger boundary. Discohelix ( Pentagonodiscus ) subgen. nov. Type species. D. (P.) angusta n. sp. Diagnosis. Subgenus of Discohelix in which the last 2-3 whorls are pentagonal in outline owing to periodical deviation of the peristome at intervals of approximately 72°. Inner whorls coiled in a normal logarithmic spiral. Spire concave, dextral. Growth- lines and ornament as in Discohelix ( Discohelix ). Discohelix ( Pentagonodiscus ) angusta n. sp. Plate 110, figs. 17-21; text-figs. 1b, 2h, 3o, 4a-b, 6a-b, 7 Derivation of name, angustus = narrow (referring to the peristome). Holotype. Ga 1347/50. Plate 110, figs. 18-21; text-fig. 1b. No. dp d h/d ha / wa nw nt sa ua Ga 1347/50 14-6 0-38 1-38 7 12 Ga 1347/51 12 9 0-33 1-28 6 Ga 1347/52 12-6 0-37 1 25 6 210° 129° Diagnosis. Protoconch smooth, retiform sculpture, and parabolic ribs on teleoconch; peristome with inner varices on inner and outer lip. 574 PALAEONTOLOGY, VOLUME 11 Description. The sculpture is similar to that of D. cf. gumbeli , but much more irregular through the arrangement of the ribs into groups of three. Between the nodes the whorl cross-section is almost round. In contrast D. (P.) reussii has broad trapezoidal whorls and is also distinguishable by the failure of ribs. (See also sections ‘Parabolic ribs’ and ‘Growth’). Distribution. Aalenian to Middle Bajocian (condensed Opalinum to c. Subfurcatum Zone). Discohelix ( Pentagonodiscus ) reussii (Hornes 1853) Plate 110, figs. 22-4; text-figs. 2g, 3p, 5 1853 Euomphalus Reussii Hornes; Hauer, p. 760. 1861 Discohelix Reussi Hornes; Stoliczka, p. 184, pi. 3, figs. 13, 14 a-c, non fig. 14 d. 1897 Discohelix Reussi Hornes; Krafft, p. 21 1. 1911 Discohelix Reussi Hornes; Gemmellaro, p. 215, pi. 9, fig. 14. 1935 Discohelix Reussi Hornes, Wahner, p. 22. Material. Eight specimens, of which five are Stoliczka’s originals. Lectotype. (here selected) Stoliczka 1861, pi. 3, figs. 14 a-b (here refigured PI. 110, figs. 22-4). No. d h/d ha/wa nw sa ua 2980 14-5 0-43 1 -97 6 Ga 1347/55 15-8 0-42 L58 6 221° 123° Diagnosis. Fine retiform sculpture on teleoconch; peristome narrowed with inner varix on outer lip. Description. Even without a longitudinal section Stoliczka supposed the occurrence of inner varices. The aperture, recurring periodically at mean intervals of71-5°(Ga 1347/56) is depressed on the outer side and somewhat broadened on the upper and lower sides. This results in regular swellings which are especially strong in the specimen figured by Stoliczka on pi. 3, fig. 13. In the remaining examples the keel has a regularly undulat- ing profile. The peristome is only slightly thickened on its inner lip, but narrowed by a strong inner varix on the outer lip as in D. ( P .) angusta. Distribution. Sinemurian-Pliensbachian. According to their label Stoliczka's specimens came from the Oxynotum Zone (Hierlatz) and the Obtusum Zone (Kratzalpe). The horizon of Wahner’s material is not known; a personal specimen from the same area (Sonnwend Mountains) was found with Arietites sensu lato. The examples of Krafft and Gemmellaro are said to be younger (Middle Lias), as is that from the Tubingen Pal. Inst, collection which is the only known extra-alpine occurrence of the species. CONCLUSION The stratigraphical division of thick sequences of the Jurassic in the Tethys region is often made extremely difficult by the sparsity or failure of zonal ammonites, and the knowledge of other fossil groups is still insufficient to replace them as stratigraphical indicators. The present investigation of the little-known gastropod genus Discohelix may help to fill this gap, since, although this genus is only sporadic in the epicontinental Jurassic, it is occasionally very frequent in the Alpine-Mediterranean area (e.g. WENDT: DISCOHELIX (ARCHAEOG ASTROPODA, EUOMPH ALACEA) 575 Hochfelln Limestone, Hierlatz Limestone, Liassic Reef Limestone of Sicily). The value of its species as index fossils, shown here, may make possible the division of such cephalo- pod free or poor series in the Tethyan Lower and Middle Jurassic. REFERENCES ammon, l. 1893. Die Gastropodenfauna des Hochfelln-Kalkes und iiber Gastropoden-Reste aus Ablagerungen von Adnet, vom Monte Nota und den Raibler Schichten. Geogn. Jh. 5, 161-219. brosamlen, r. 1909. Beitrag zur Kenntnis der Gastropoden des schwabischen Jura. Palaeontographica, 56, 177-321, pi. 17-22. cossmann, M. 1915. Essais de Paleoconchologie comparee, 10, 292 pp., 12 pi. dumortier, E. 1874. Etudes paleontologiques sur les depots jurassiques du bassiit du Rhone. 4me partie. Lias superieur. 339 pp., 62 pi. dunker, w. 1847. Ueber einige neue Versteinerungen aus verschiedenen Gebirgsformationen. Palae- ontographica, 1, 128-33, pi. 18. fretter, v. and graham, a. 1962. British Prosobranch Molluscs, xvi+755 pp., London. gemmellaro, M. 1911. Sui fossili degli strati a Terebratula aspasia della contrada Rocche Rosse presso Galati (prov. di Messina). Cefalopodi (fine) — Gasteropodi. G. Sci. nat. econ. Palermo, 28, 203-46, pi. 8-10. hauer, f. 1853. Ueber die Gliederung der Trias-, Lias- und Juragebilde in den nordostlichen Alpen. Jb. K. K. geol. Reichsanst. 4, 715-84. hudleston, w. h. 1887-96. A monograph of the Inferior Oolite Gastropoda, being part I of the British Jurassic Gastropoda. Palaeontogr. Soc. (1886-96), 1-514, pi. 1-44. knight, j. b. 1934. The gastropods of the St. Louis, Missouri, Pennsylvanian outlier: VII. The Euom- phalidae and Platyceratidae. J. Paleont. 8, 139-66, pi. 20-6. koken, e. 1897. Die Gastropoden der Trias um Hallstatt. Abh. K. K. geol. Reichsanst. 17, H. 4, 1-112, pi. 1-23. krafft, A. 1897. Ueber den Lias des Hagengebirges. Jb. K. K. geol. Reichsanst., 47, 199-224, pi. 4. lycett, j. 1850. Tabular view of fossil shells from the middle division of the Inferior Oolite in Gloucestershire. Ann. Mag. nat. Hist., ser. 2, 6, 401-25, pi. 11. moore, ch. 1867. On the Middle and Upper Lias of the South West of England. Proc. Somerset, archaeol. nat. Hist. Soc. 13, 1-128, pi. 1-7. quenstedt, f. a. 1884. Petrefaktenkunde Deutschlands. 1. Abt. 7. Bd: Gasteropoden. 867 pp., pi. 1 85— 218, Leipzig. reuss, a. 1852. Ueber zwei neue Euomphalusarten des alpinen Lias. Palaeontographica, 3, 113-16, pi. 16. riche, a. and roman, f. 1921. La Montagne de Crussol. Etude stratigraphique et paleontologique. Trav. Lab. Geol. Univ. Lyon, 1, 1-196, pi. 1-8. roman, f. 1935. La faune du minerais de fer des environs de Privas. Ibid. 27, 1-52, pi. 1-8. schairer, G. 1967. Biometrische Untersuchungen an Perisphinctes, Ataxioceras, Lithacoceras der Zone der Sutneria platynota (reinecke) (unterstes Unterkimmeridgium) der Frankischen Alb. Inaug.-Diss. naturw. Pak. Univ. Munchen, 131 pp., 18 pi. schindewolf, o. h. 1934. Uber eine oberdevonische Ammoneen-Fauna aus den Rocky Mountains. Neues Jb. Miner. Geol. Paldont. Bei/Bd. 72, Abt. B, 331-50. 1937. Zur Stratigraphie und Palaontologie der Wocklumer Schichten (Oberdevon). Abh. preufi. geol. Landesanst., n.f. 178, 3-132, pi. 1-4. stoliczka, F. 1861. Uber die Gastropoden und Acephalen der Hierlatz-Schichten. Sber. Akad. Wiss. Wien. math, naturw. Cl. 43, 157-204, pi. 1-7. wahner, f. 1935. Das Sonnwendgebirge im Unterinntal. Ein Typus alpinen Gebirgsbaues. Zweiter Teil bearb. u. voll. v. e. spengler). xvi+200 pp., 29 pi., Leipzig and Wien. wendt, J. 1965. Synsedimentare Bruchtektonik im Jura Westsiziliens Neues Jb. geol. Paldont. Mh. (1965), 286-31 1. j. wendt Institut fur Geologie und Palaontologie Typescript received 18 January 1968 Tubingen, Germany PEDICELLARIAE OF TWO SILURIAN ECHINOIDS FROM WESTERN ENGLAND by D. BRYAN BLAKE Abstract. Pedicellariae have been discovered on the Silurian (Ludlovian) echinoids Echinocystites pomum and Palaeodiscits ferox from Leintwardine, England. All pedicellariae of both species are tridentate in form and simple in morphology. Most are interambulacral in position. Pedicellariae of E. pomum are abundant, all are probably three-valved, and of variable size. Those of P. ferox are relatively rare, apparently two-valved, and small in size. These pedicellariae, and an Australian occurrence, are the oldest known echinoid pedicellariae. The unusually large size of one fossil pedicellaria, combined with a general morphology of all pedicellariae inter- mediate between that of spines and modern pedicellariae suggests that pedicellariae were derived from spines. While studying materials in the Echinoderm Collections of the British Museum (Natural History), Dr. J. Wyatt Durham observed pedicellariae on specimens of the Silurian echinoid Echinocystites pomum and made latex casts of four specimens (three with pedicellariae) of this species and of two specimens of Palaeodiscus ferox. In subse- quent studies, the writer observed the presence of pedicellariae on both specimens of P. ferox. All fossils are from the Ludlow Flags of Leintwardine, Herefordshire. These occurrences, with that of Philip (1963), who recorded pedicellariae associated with a lepidocentroid from rocks regarded as Ludlovian in age from New South Wales, Australia, represent the oldest known echinoid pedicellariae and thus merit detailed consideration. The Leintwardine material has been studied by many workers including Gregory (1897), Thomson (1861), and Hawkins (1927) but the presence of pedicellariae has not been previously noted in spite of the fact that one of the specimens is the lectotype of Echinocystites pomum. Mortensen (1928-51) considered pedicellariae important in echinoid systematics in his monograph of the group but noted and regretted the rarity of fossil pedicellariae (op. cit., vol. 1, p. 33). He did not mention pedicellariae in his discussion of either of these two genera (op. cit., vol. 2, pp. 54-6, 60-3). Reports of the few described Palaeozoic pedicellariae are reviewed by Geis (1936). Acknowledgements. Dr. Durham made available latex casts of specimens of E. pomum and P. ferox', the original external moulds of these specimens are in the collections of the British Museum (Natural History) in London. Dr. Porter M. Kier loaned a specimen of E. pomum belonging to the Geological Survey of Great Britain which he had prepared for study and identified. Thanks are due to Dr. R. P. S. Jefferies and the authorities of the British Museum (Natural History) for permitting Dr. Durham to prepare latex casts of their specimens and to Dr. Jefferies for information on the histories of some of the specimens. Echinocystites pomum All pedicellariae appear to be of the tridentate type, and all probably originally had three valves. (‘Tridentate’ is one of the four basic types of pedicellariae; most, but not all, have three valves (Mortensen, op. cit., vol. 2, p. 33).) They are numerous; sixty were counted on one incompletely preserved test. The blades of a pedicellaria are in contact [Palaeontology, Vol. 11, Part 4, 1968, pp. 576-9, pis. 1-2.) BLAKE: PEDICELLARIAE OF TWO SILURIAN ECHINOIDS 577 throughout their length; they are slender and elongate with a circular cross-section, and are attenuated distally. The blades bear longitudinal striae but apparently no marginal dentition; their general morphology is essentially identical to that of the spines. The bases are triangular and grade evenly into the blades. Unlike blades of many tridentate pedicellariae, there is little or no constriction near the base. The interior of the base is massive, apparently without a sharp keel-like apophysis. Many heads are located near to tubercles suggesting that stems were lacking, very short, or had contracted upon death. Small pedicellariae are most common, but a considerable range in size exists; there are no distinct size classes. Bases range in diameter from under 0-25 mm. to about T75 mm. In general, head length in essentially complete examples ranges from about 0-75 mm. in specimens with small bases to about 3-75 mm. on a specimen of base diameter of 0-75 mm. However, one unusually large pedicellaria (text-fig. 1a) has a base diameter of 1-75 mm. and a length slightly over 11 mm. Blade length is not neces- sarily directly proportional to base diameter because blade length on bases of a given size is variable. Only two valves are visible on a number of pedicellariae. However, the great variability of preservation of the three-valved forms and the essential similarity of morphology between those with three valves and those with only two implies that the apparent two-valved pedicellariae are probably incompletely preserved typical three-valved pedicellariae. With two possible exceptions, all pedicellariae are interambulacral in position. They are widely distributed on the interambulacral columns; on one specimen, pedicellariae occur on plates which border the madreporite, while on another, pedicellariae occur on plates near the peristome. They may be densely grouped, occurring on neighbouring plates, or in some cases, more than one on a single plate. There appears to have been no areal restriction of pedicellariae of certain sizes, for large and small pedicellariae occur on neighbouring plates. Material examined. Casts of B.M. (N.H.) BMNH 40158; E34352 (lectotype); 40156; GSM 102622. Palaeodiseus ferox Pedicellariae are rare on specimens of this species; only two well-preserved pedicel- lariae have been observed on each of the two specimens. In addition, a number of doubtful pedicellariae are present. All pedicellariae are tridentate in form but with only two valves; there is no evidence for a third valve. The heads are elongate, triangular, and rounded proximally. The blades, like those of Echinocystites, are morphologically very similar to the spines; they are slender, elongate, circular, or sub-circular in cross-section, with longitudinal striae, and in contact throughout their length. There appears to have been no marginal dentition. Bases appear to have been triangular and only slightly enlarged. The heads are very small with a basal dimension of approximately 0-2 mm. and a length of about 0-8 mm. One of the well-preserved pedicellaria is ambulacral in position located near the groove between the rows of ambulacral plates; the others are interambulacral in position. Although relatively few pedicellariae can be identified, the similarity of morphology between the blades and the spines suggests more pedicellariae may be present but are incompletely preserved or exposed and therefore unrecognized. Material examined. Casts of Palaeodiseus ferex, BMNH E34360 and E34362. 578 PALAEONTOLOGY, VOLUME 11 DISCUSSION The presence of previously unrecorded pedicellariae, in one case abundant, on much- studied echinoid species suggests fossil pedicellariae are probably more common than text-figs. 1a-g. Echinocystltes pomum pedicellariae. 1, GSM 102622, x7-5. 2-5, 7, B.M. 40156, X 20. 6, from B.M. E34352, x 20. h. Palaeodiscus ferox pedicellaria. B.M. E34360. previously has been thought. Further, because two families of Silurian echinoids are involved, evolution of these structures must have been relatively early and fairly wide- spread. In many features, the Silurian pedicellariae are spine-like. The valves are circular in cross-section, the surface ornamentation is similar to that of the spines, marginal dentition is absent, the base of the blade is not strongly constricted, blades of a pedicel- laria are in contact throughout their length, the bases are massive without a sharp apophysis. Further, one unusually long spine-like pedicellaria is present. These features render the Silurian pedicellariae morphologically intermediate between spines and modern types of pedicellariae and suggest the possibility that pedicellariae were derived from spines. BLAKE: PEDICELLARIAE OF TWO SILURIAN ECHINOIDS 579 REFERENCES geis, h. l. 1936. Recent and fossil Pedicellariae. J. Paleont. 10, 427-48, pi. 58-61. Gregory, j. w. 1897. On Echniocystis and Palaeodiscus — two Silurian genera of Echinoidea. Quart. J. geol. Soc. Loud. 53, 123-36, 7 text-fig., pi. 7. hawkins, H. L. and hampton, s. m. 1927. The occurrence, structure, and affinities of Echinocystis and Palaeodiscus. Ibid. 83, 574-603. mortensen, t. 1928-51. A monograph of the Echinoidea. 5 vols. Copenhagen. philip, G. m. 1963. Silurian echinoid pedicellariae from New South Wales. Nature, Lond. 200, 1334. Thomson, w. 1861. On a new Palaeozoic group of Echinodermata. Edinb. New. Phil. J. N.s. 13, 106-18. D. BRYAN BLAKE Department of Geology University of Illinois Urbana, Illinois 61801 U.S.A. Typescript received 10 November 1967 MACROCYSTELLA CALLAWAY, THE EARLIEST GLYPTOCYSTITID CYSTOLD by C. R. C. PAUL Abstract. Macrocystella mariae Callaway 1877, type species of Macrocystella, has a stem which is divisible into proximal and distal portions; a theca composed of 4 basal, 5 infra-lateral, 5 lateral, 6 radial, and some oral plates ; a large periproct surrounded by 5 thecal plates ; biserial unbranched brachioles grouped into 5 ambulacra and arising from the margins of the flattened oral surface. In all these respects it agrees with Mimocystites bohemicus Barrande 1887, type species of Mimocystites which becomes a subjective junior synonym of Macro- cystella. Macrocystella azaisi (Thoral) has 7 orals and thus Macrocystella differs from the rhombiferan Cheiro- crinus Eichwald only in the absence of pectinirhombs. The Macrocystellidae are therefore transferred to the rhombiferan superfamily Glyptocystitida. Macrocystella evolved into Cheirocrinus by the acquisition of pectinirhombs. In Macrocystella respiration probably took place through all the thecal plates which are very thin. In Cheirocrinus respiration was restricted to the pectinirhombs thus allowing much thicker and stronger thecal plates to develop. Macrocystella led a freely vagrant existence and may have had internal buoyancy devices. The stem did not provide permanent fixture and may have been used as a organ of locomotion in conjunction with the brachioles. The cystoidea, as currently defined (Kesling 1963) is probably an artificial group. The main character which is used to unite the cystoids as a class is the possession of pore- structures (rhombs and dipores) developed in the thecal plates. However similar pore- structures are found in at least some representatives of other Palaeozoic echinoderm classes (blastoids, crinoids, paracrinoids, eocrinoids, for example) and one genus of rhombiferan cystoids entirely lacks pore-structures. This paper deals with another genus, Macrocystella Callaway 1877, which lacks true pore-structures but which is thought to be the oldest known representative of the Glyptocystitida, one of the three major rhombiferan superfamilies. Macrocystella has a complex taxonomic history (see below) and has been variously regarded as an eocrinoid, a rhombiferan cystoid or as a link between these classes. Close comparison indicates that Macrocystella is identical to the rhombiferan genus Cheirocrinus Eichwald in all details except the possession of pectinirhombs. Hence Macrocystella is regarded as a rhombiferan. It is believed that the absence of pectinirhombs in Macrocystella is a primitive character. Many Ordovician pelmatozoans independently developed thecal or calycinal pore structures, apparently in response to respiratory needs. To group all such echinoderms together obscures their true relationships, it is essential to consider other characters in addition to the possession of pore-structures, especially when the latter are so variable. Acknowledgements. The author is indebted to Dr. R. P. S. Jefferies and Mr. H. G. Owen, British Museum, Natural History (BMNH), Professor B. Kummel and Mr. J. Sprinkle, Museum of Compara- tive Zoology, Harvard (MCZ), Mr A. G. Brighton, Sedgwick Museum, Cambridge (SM), and Dr. R. A. Robison, United States National Museum, Washington (USNM) for the loan of, or access to, specimens. Dr. I. Strachan, Birmingham University (BU) supplied latex impressions of Callaway’s original specimens of M. mariae and Professor K. E. Caster, University of Cincinnati (UC) impressions of Macrocystella and Cheirocrinus from France and Bohemia. Part of this work was undertaken during the tenure of a National Environmental Research Council research studentship at the Sedgwick [Palaeontology, Vol. 11, Part 4, 1968, pp. 580-600, pis. 111-13.] C. R. C. PAUL: MACROCYSTELLA CALLAWAY 581 Museum, Cambridge and part during the tenure of a Post-doctoral Fellowship at the University of Michigan (NSF grant GB-3366). Both are gratefully acknowledged. PREVIOUS RESEARCH MacrocysteUa (type species M. mariae Callaway 1877) was first described from the Lower Ordovician (Tremadoc) Shineton Shales of Shropshire. Barrande (1887, p. 163) described a closely similar genus, Mimocystites, for a single species, M. bohemicus Barrande. Jaekel (1899, p. 171) suggested that these two genera were synonymous but used the name Mimocystites. Jaekel regarded Mimocystites as the progenitor of the cystoids and most closely related to the rhombiferan Cheirocrinus Eichwald. Bather (1899, p. 920) proposed the family Macrocystellidae which he assigned to the Rhombi- fera. He did not elucidate the composition of the Macrocystellidae but later (1900, p. 56) included MacrocysteUa, Mimocystites and Lichenoides Barrande 1887. Although Bather thought MacrocysteUa and Mimocystites hardly differed he used both names. Bather’s reconstruction of MacrocysteUa mariae (1900, p. 95, fig. 18) was inaccurate in depicting branched arms and considerably influenced subsequent opinions on the affinities of MacrocysteUa. Jaekel (1918, p. 27) retained both MacrocysteUa and Mimocystites (possibly because of Bather’s reconstruction of the former) and placed them, with his new genus Polypty- chella, in the Macrocystellidae. A separate family was proposed for Lichenoides. The Macrocystellidae and Lichenoidae were assigned respectively to the orders Plicata and Reducta of the Eocrinoidea. Thoral (1935, p. 113) considered MacrocysteUa and Mimo- cystites to be distinct but based his opinion of the former on the original description. Bassler and Moodey (1943) reverted to Bather’s classification, included Lichenoides in the Macrocystellidae and that family in the Rhombifera. Cuenot (1948) also assigned MacrocysteUa to the Rhombifera and later (1953) regarded Mimocystites as a junior synonym. Moore (1954, p. 127, fig. 2d) published an inaccurate plate diagram of MacrocysteUa which depicts 5 basals, 5 radials, 5 orals, and a minute periproct. Moore regarded MacrocysteUa as an eocrinoid. Sdzuy (1955) accepted MacrocysteUa and Mimocystites as separate genera only if published reconstructions were accurate. He based his opinion of MacrocysteUa on Bather’s (1900) and Moore’s (1954) work, the accuracy of which he doubted. More recently Prokop (1966) and Ubaghs (1968) have suggested that MacrocysteUa and Mimocystites are synonymous. Ubaghs regards MacrocysteUa as most closely related to Cheirocrinus. Quite apart from the synonymy of MacrocysteUa and Mimocystites another problem arises as to the status of Cystidea Barrande 1868. Barrande (1867, p. 179) published two nomina nuda, Cystidea sedgwicki and C. bohemicus. Later he introduced C. bavarica (Barrande 1868, p. 106) this time accompanied by a description and figures. Barrande made it quite clear in both publications that he intended Cystidea as a collective group name not a formal generic name and he so used it again (1887), erecting several more species. Cystidea bavarica Barrande is a valid binomen and could be construed to be type species of the genus Cystidea Barrande by monotypy. Pompeckj (1896, p. 90) and Sdzuy (1955, p. 170) have attributed Cystidea bavarica to MacrocysteUa. There is no doubt they are correct in this action and therefore Cystidea Barrande 1868, if accepted as an available generic name, should take precedence over both MacrocysteUa and Mimocystites. Such action is not in the interests of nomenclatorial stability. No author 582 PALAEONTOLOGY, VOLUME 11 has accepted Cystidea as a valid generic name whereas MacrocysteUa has been widely used and is figured and described in standard text-books in English, French, and German. Application has therefore been made to the International Commission on Zoological Nomenclature for the suppression of Cystidea Barrande 1868 under the Plenary Powers (Paul 1967fi). In anticipation of a favourable decision, MacrocysteUa is used throughout this work. The systematics and composition of the Macrocystellidae and the suggested synonymy between MacrocysteUa and Mimocystites can only be settled after a revised account of the morphology of MacrocysteUa mariae has been given. Latex impressions of the original specimens of M. mariae have been re-examined and additional material studied. This has been compared with Barrande’s (1887) and Jaekel’s (1899) descriptions and figures of Mimocystites bohemicus, with latex impressions of some of Barrande’s original material and additional material of M. bohemicus in the Schary Collection, Museum of Comparative Zoology, Harvard. Latex impressions of M. azaisi (Thoral) have added further information. MacrocysteUa and Mimocystites are identical and quite distinct from Lichenoides to judge from Ubagh’s (1953) account of L. priscus Barrande. Polyptychella Jaekel was founded on isolated plates and its systematic position cannot be settled without further information. The Macrocystellidae thus contains the single genus MacrocysteUa. As previously stated the Macrocystellidae is assigned to the Rhombifera (Glyptocystitida). SYSTEMATIC PALAEONTOLOGY Superfamily glyptocystitida Bather 1899 Diagnosis. A superfamily of Rhombifera with well-developed stem divided into proximal and distal portions; theca composed of 4 basals, 5 infra-laterals, 5 laterals, 4-6 radials, and 7 orals; with pectinirhombs (when pore structures are developed). All Glyptocystitida are characterized by a theca composed of 25-7 thecal plates arranged in five circlets termed basal, infra-lateral, lateral, radial, and oral. All but two genera — MacrocysteUa and Amecystis Ulrich and Kirk — have pectinirhombs. These characters distinguish glyptocystitids from members of the other two major rhombi- feran superfamilies, the Hemicosmitida and Caryocystitida. The former have thecal plates arranged in three or four circlets and slightly different rhombs. The latter have a large variable number of plates, some of which may be added during growth, and a completely different type of rhomb (Paul 1968). Family macrocystellidae Bather 1899 emend. Jaekel 1918 Diagnosis. A family of Glyptocystitida without pectinirhombs; with cylindrical theca having 6 radials; large periproct surrounded by 5 thecal plates and covered with a flexible plated integument; brachioles confined to oral surface, grouped into 5 ambulacra. Genus macrocystella Callaway 1877 1868 Cystidea Barrande, p. 106. 1877 MacrocysteUa Callaway, p. 669, pi. 24, fig. 13. C. R. C. PAUL: MAC ROCYSTELLA CALLAWAY 583 1880 Mcicrocystella Callaway; Zittel, p. 420. 1887 Mimocystites Barrande, p. 163, pi. 28 (1), figs. 1-20. 1891 Macrocystella Callaway; Carpenter, p. 13. 1891 Mimocystis [,s7c] Barrande; Carpenter, p. 13. 1896 Mimocystis [s/c] Barrande; Haeckel, p. 149. 1899 Macrocystella Callaway; Jaekel, p. 171. 1899 Mimocystites Barrande; Jaekel, p. 172, fig. 33. 1900 Macrocystella Callaway; Bather, p. 56, fig. 18. 1900 Mimocystis [s/c] Barrande; Bather, p. 56. 1913 Macrocystella Callaway; Springer, p. 157. 1918 Macrocystella Callaway; Jaekel, p. 27. 1918 Mimocystites Barrande; Jaekel, p. 27. 1935 Mimocystites Barrande; Thoral, p. 1 10, 1 13. 1943 Macrocystella Callaway; Bassler and Moodey, p. 6. 1943 Mimocystites Barrande; Bassler and Moodey, p. 6. 1948 Macrocystella Callaway; Regnell, p. 11. 1948 Macrocystella Callaway ; Cuenot, p. 18, fig. 17. 1953 Macrocystella Callaway; Cuenot, p. 619. 1953 Mimocystites Barrande; Choubert, Termier, and Termier, p. 137. 1954 Macrocystella Callaway; Moore, p. 127, fig. 2a. 1954 Mimocystites Barrande; Termier and Termier, p. 92, figs. a-e. 1955 Macrocystella Callaway; Sdzuy, p. 269. 1966 Macrocystella Callaway; Prokop, p. 820. non Cvstidea Barrande 1867 (nomen nudum) nee Barrande 1887 nee Haeckel 1896 (inde- terminate echinodernt fragments). Diagnosis. As for family. Regional distribution and stratigraphic range. Macrocystella is recorded from the Tremadoc of England and Wales (M. mariae Callaway), Bavaria (M. bavarica (Barrande) 1868), Bohemia (M. bohemicus Barrande 1887), and France (M. azaisi (Thoral) 1935). Macrocystella is also recorded from Greenland, the South American Cordillera and Korea (for detailed references see Regnell, 1948, pp. 11-12). Choubert, Termier, and Termier record Macrocystella from the Llandeilo of Morocco. Available specimens confirm the genus from the Llandeilo of Pu-piao, Northern Shan States, Burma, (SM), from the Tremadoc of Mexico (USNM) and possibly from the Caradoc of Corwen, Wales, and Girvan, Scotland (BMNH). M. pachecoi Melendez (1944) from the Ashgill of Aragon, Spain is probably a Heliocrinites. Macrocystella thus ranges from the Tremadoc to the Llandeilo and possibly Caradoc (Lower-Middle Ordovician). Macrocystella mariae Callaway 1877 Plate 111, figs. 1, 3-6; Plate 1 12, figs. 1-3, 5-10; Plate 113, fig. 2 1877 Macrocystella mariae Callaway, p. 670, pi. 24, fig. 13. 1896 Macrocystella mariae Callaway; Haeckel, p. 149, pi. 4, fig. 30. 1900 Macrocystella mariae Callaway; Bather, p. 56, fig. 18. 1905 Macrocystella mariae Callaway; Fearnsides, p. 617. 1911 Macrocystella mariae Callaway; Kirk p. 16, pi. 2, fig. 17. 1913 Macrocystella mariae Callaway; Springer, p. 157, fig. 249. 1927 Macrocystella mariae Callaway; Stubblefield and Bulrnan, pp. Ill, 118. 1943 Macrocystella mariae Callaway; Bassler and Moodey, pp. 27, 175. 1952 Macrocystella mariae Callaway; Termier and Termier, p. 363, fig. 9. 1953 Macrocystella mariae Callaway; Cuenot, p. 618, fig. 15. 1955 Macrocystella mariae Callaway; Sdzuy, p. 270, pi. 1, fig. 14. 1964 Macrocystella mariae Callaway; Castell, p. 58, pi. 3, fig. 6. 584 PALAEONTOLOGY, VOLUME 11 Diagnosis. A species of Macrocystella of small size; with 10-15 brachioles, triangular in section; circular outer proximal columnals with thin, blade-like external flanges. Type. BU 409 (PI. 113, fig. 2) is selected as lectotype. It is possibly the original of Callaway 1877, pi. 24, fig. 13 and is from the Shineton Shales of Shineton, Shropshire. Parts of other specimens on this slab are accepted as paralectotypes. Horizon and locality. Stubblefield and Bulman (1927) record M. marine from the Clonograptus tenellus and Slnanardia pusi/la zones (Middle and Upper Tremadoc respectively) of the Wrekin. M. mariae is also recorded from the Slmmardia pusilla zone of Arenig (Fearnsides, 1905, p. 617) and Macrocystella sp. from the same zone near Portmadoc, Caernarvonshire (Fearnsides 1910, pp. 161-2). Material. Crushed remains of four more or less complete thecae, one complete stem and many isolated thecal plates and fragments. Description, a. Stem The stem has a proximal and a distal portion. One complete proximal portion (PI. 112, fig. 8) has 20 outer proximals each with a blade-like unornamented external flange. This crushed portion tapers from 5 mm. adorally to 2 mm. in approximately 15 mm. At the junction with the distal stem small distal columnals appear between the flanged columnals. Throughout the preserved portion of the distal stem, flanged and unflanged columnals alternate but this alternation becomes less obvious distally. The distal portion of the stem (PI. Ill, fig. 1) tapers gradually from 1 mm. proximally to 0-5 mm. at the tip. It is about 35 mm. long. The topmost distal columnal is a thin annulus; the terminal distals are cylindrical and about three times as high as wide. There is no alternation of flanged and unflanged distals in this stem. Preserved proximal stems are straight or quite strongly curved. Curved distal stems are also preserved but there is no evidence to support Bather’s (1900) interpretation of the distal stem with a distinct distal coil. Counterparts of isolated columnals indicate the mode of construction and articula- tion of the stem. Both proximal and distal portions are composed of two types of columnals. Outer proximals (text-figs. 1 a-b) are annular. Each has a smooth, sharp- edged outer flange and a narrow inner flange (text-fig. 2, PI. 1 12, fig. 7). Outer proximals alternate with inner proximals and the latter abut against the inner flanges of the former (text-fig. 2). The inner wall of each outer proximal has two sockets set opposite each other on both the upper and lower surfaces. The inner flanges are thickened adjacent to these sockets to form fulcra (text-fig. 2). If the sockets on one surface are orientated N.-S., those on the opposite surface of the same columnal lie NW.-SE. (text-figs. 1 a-b). Both upper and lower surfaces of the inner proximals are flattened to form facets. EXPLANATION OF PLATE 111 Stereophotos of Macrocystella mariae Callaway and M. azaisi (Thoral) Figs. 1, 3-6. M. mariae Callaway. 1. Complete stem showing proximal and distal portions. BMNH E291 13. 3. Left lateral view of crushed theca. BMNH E29110a. 4. Right lateral view of same theca. BMNH E29109a. 5. Anterior lateral view of another crushed theca. BMNH E29109b. 6. Posterior view of crushed theca to show outline of periproct and small periproctal plates. BMNH E29113. Fig. 2. M. azaisi (Thoral). Proximal and part of distal stem to show ornament of flanges on outer proximals. BMNH E23697. All figures of latex impressions whitened with ammonium chloride sublimate. All X 2. Palaeontology, Vol. 11 PLATE 111 PAUL, Macrocystella C. R. C. PAUL: MACROCYSTELLA CALLAWAY 585 at the same two opposite points (text-figs, lc, 2). The outer margin of an inner proximal protrudes at these points and the protrusions key into the sockets in the inner wall of the outer proximals above and below (text-fig. 2). The fulcra on the outer proximals and the facets on the inner proximals articulate and the axis of articulation changes by approximately 45° with each outer proximal. This results in a right-handed spiral arrangement of articulation facets in M. cizaisi and presumably in other MacrocysteUa. outer proximal columnal (OP) in the same orientation to show different orientations of fulcra (Fu). c. Inner proximal columnal (IP). Ef, external flange; Fa, facet; If, internal flange. text-fig. 2. Diagrammatic reconstruction of part of proximal stem of MacrocysteUa mariae Callaway to show arrangement of outer (OP) and inner (IP) columnals and spiral arrangement of facets (Fa) and fulcra (Fu). Ef, external flange; If, internal flange; IW, inner wall of outer proximal columnal. This has been drawn as a left-handed spiral although M. azaisi is known to show a right-handed spiral. Each inner proximal is keyed into the sockets of the outer proximals above and below. Thus although highly flexible, the stem was quite resistant to rotation about its axis. Both inner and outer proximals are annular and the proximal stem has a wide lumen. Flexing of the proximal stem was probably achieved by muscles housed in this lumen. The mechanical keying of the columnals prevented rotation which would have sheared such muscles. Larger (sometimes flanged) and smaller (unflanged) distals alternate in the distal stem but this becomes less apparent distally. The most proximal distals, which appear to be 586 PALAEONTOLOGY, VOLUME 11 newly formed, are annular but most distals are cylindrical. All distals have narrow lumina and the articulating surfaces are smooth (PI. 112, fig. 3). The nature of the outer proximals may prove to be a useful specific character. In M. mariae the outer proximals are approximately circular with thin, blade-like outer flanges (PI. 1 1 1, fig. 1, PI. 112, fig. 7). In M. bavarica the outer flanges are also thin and sharp-edged but are produced into ten angles or incipient spines (Sdzuy, 1955, pi. 1, figs. 8-10, text-fig. 1//). In M. azaisi the outline is circular but the flanges are thicker and have fine irregular granules or spines encircling them (PI. 11 1, fig. 2, PI. 1 13, figs. 1, 8). The outer flanges of M. sp. nov. from Mexico have four rounded lobes arranged in two pairs. b. Theca All known thecae are crushed and it is impossible to describe all thecal plates from one specimen. BMNH E29 109-10 are counterparts which show two crushed thecae in different orientations. There is another theca which shows details of the periproct. text-fig. 3. Diagrammatic reconstruction of plate arrangement of Macrocystella mariae Callaway. Based on BMNH E29109a-b, E29110a-b, and E29113. B1-B4 basals, IL1-IL5 infra-laterals, L1-L5 laterals, R1-R6 radials. A composite plate arrangement and a reconstruction based on these specimens are depicted in text-figs. 3 and 15. The theca was composed of five circlets of plates, four of which can be seen in M. mariae. The subvective system was confined to the oral surface from the margins of which the brachioles arose in five groups. Some details of text-fig. 3 are restored but all plates shown existed. The four basals (BB) unite aborally to form an invagination around the stem. One basal (presumably B4) was hexagonal (PI. 112, fig. 2). The infra-laterals (ILL) form a closed circlet; IL4 and IL5 contribute to the periproct border. IL1, IL2, and IL3 are roughly hexagonal. The five laterals (LL) apparently form a closed circlet; LI, L4, and L5 contribute to the periproct border. L5 is distinctly smaller than the other laterals and has a radial (R5) C. R. C. PAUL: MAC ROCYSTELLA CALLAWAY 587 directly adoral to it. One pair of counterparts (PI. Ill, figs. 3, 4) seem to have three hexagonal laterals which means either there were six laterals in this specimen or only LI and L5 contributed to the periproct border. Unfortunately the relevant portion of this theca is crushed and this appearance may be misleading. The six radials (RR) form a closed circlet; one is directly adoral to L5. There are seven orals (00) in M. azaisi. text-figs. 4-5. Camera lucida drawings of the periproct (Pe) of Macrocystella mariae Callaway. 4. BMNH E291 10b cf. PI. 2, fig. 9. 5. BMNH E29113 cf. Plate 1, fig. 6. IL4-IL5 infra-laterals, LI, L4, L5 laterals. The mouth, gonopore, and hydropore have not been detected in M. mariae but were all on the oral surface in M. azaisi (text-figs. 1 1 a, b ). Critical details of the periproct show in BMNH E291 10b, and more clearly, in BMNH E29113 (text-figs. 4, 5). There are five plates around the periproct which is large. The periproct was covered by a thin, flexible, plated integument in life. Some small periproctal plates are preserved (PI. Ill, fig. 6) but the position of the anal pyramid in unknown. The outlines of the individual thecal plates vary with their positions in the theca. Periproct border plates can be recognized easily, as can basals and radials. The remain- ing five laterals and infra-laterals are difficult to distinguish from each other. All plates have raised umbones from which ridges radiate to the middles of the sides, connecting centres of adjacent plates (text-fig. 6). Auxiliary ridges are developed parallel to the primary ridges to form ‘rhombs’. These ‘rhombs’ are not true rhombs as they are composed of folds in the thecal plates. There is no development of thin-walled thecal canals. The external ridges of Macrocystella are formed by the folds in the plates and are not solid strengthening struts such as occur in Cheirocrinus. The thecal plates of M. mariae are approximately OT mm. thick. Q9 C 5934 588 PALAEONTOLOGY, VOLUME 11 text-fig. 6. Isolated thecal plate of Macrocystella mariae Callaway to show folds. BMNH E29119. c. Subvective system The exothecal portion of the subvective system consists entirely of brachioles. These are long, slender, biserial, unbranched structures. The most complete brachioles (PI. 112, fig. 10) have approximately 120 brachiolar plates and are slightly longer than the thecal height (up to 16 mm.). The brachioles have a triangular cross-section with sides twice the width of the adoral surface (text-fig. 7). The food groove ran down the centre of the adoral surface and was covered by lappets in life. The lappets are between one and a half times and twice as numerous as the brachiolars and apparently were flexible. Some have been preserved covering the food groove; others in an ‘open’ position (PI. 112, fig. 10). The lappets alternate and each one imbricates over its more distal neighbour (text-fig. 8). EXPLANATION OF PLATE 112 Figs. 1-3, 5-10. M. mariae Callaway. 1. Internal and external views of two isolated thecal plates. The internal view (above) shows the folds in the plates. The other plate was one of five bordering the periproct (probably LI). BMNHE291 12, X 5. 2. External view of B4 and an isolated distal columnal. BMNHE29112, x5. 3. Three isolated thecal plates. All originally bordered the periproct. BMNH E29112, x 5. 5. Internal mould of isolated thecal plate BMNH E29 11 9, X3.6. Inner proximal colum- nal within outer proximal columnal. BMNH E291 12, x 2. 7. Outer proximal columnal showing inner flange and fulcra. BMNH E7574, X 1 -5. 8. Crushed stem and theca. Note the larger distals are flanged. BMNH E7574, X 1-5. 9. Stereophotos of posterior lateral view of crushed theca showing periproct. Counterpart to Plate 1, fig. 5. BMNH E29110b, x2. 10. Detail of brachioles to show lateral, ab- and adoral views. BMNH E29109, X 3. Fig. 4. M. aiaisi (Thoral). Lateral view of theca and stem. BMNH E23697, X 2. All except Fig. 5 latex impressions, all whitened with ammonium chloride sublimate. Palaeontology , Vol. 11 PLATE 112 PAUL, Macrocyste/la C. R. C. PAUL: M AC RO CYSTELLA CALLAWAY 589 They coalesce towards the margins of the brachiole to form a continuous narrow band (text-fig. 9). The free portions were able to curl in on themselves. This curling in is not preservational and the lappets may have been only partially calcified in life. 7 text-figs. 7-9. Brachioles of Macrocystella marine Callaway. 7. Diagrammatic section through brachiole without lappets. 8. Camera lucida drawing of portion of brachiole with lappets (La) closed over food groove. Note that the lappets alternate and imbricate. 9. Camera lucida drawing of aboral view of brachiole with lappets in ‘open’ position. Br brachiolar plate. Text-figs. 8-9 based on BMNH E29113. No more than three brachioles arise in any one ambulacrum and there were presum- ably 10-15 brachioles in all. There is no evidence to support Bather’s (1900, fig. 18) reconstruction of branched brachioles. BMNH E29110a shows three brachioles in one radius (PI. Ill, fig. 3): all three are separate entities from their origin at the margin of the theca. Macrocystella mariae Callaway is characterized by the following: 1 . A stem which is divisible into two portions: a short, rapidly tapering, highly flexible, proximal portion composed of two types of annular columnals with a wide lumen; and a long distal portion composed of cylindrical columnals with a narrow lumen. 2. A theca with plate formula 4BB, 5ILL, 5LL, 6RR, 700. 3. A large periproct surrounded by five thecal plates and covered with a flexible plated integument. 4. Biserial, unbranched brachioles which arise from the margins of the flat oral surface and are grouped into five ambulacra. COMPARISON WITH MIMOCYSTITES BARRANDE Mimocystites boliemicus Barrande has a subvective system which is confined to the flat oral surface and consists of about twenty brachioles grouped into five ambulacra. 590 PALAEONTOLOGY, VOLUME 11 Three brachioles were present in one radius of an available specimen (UC latex impres- sion of original of Barrande 1887, pi. 28 (i), fig. 14). The theca is composed of 4BB, 5ILL, 5LL, 6RR, and some 00. Jaekel’s analysis of the arrangement of thecal plates (1899, p. 201, fig. 36) is reproduced here in a slightly modified form (text-fig. 10). This interpretation differs slightly from that of Macrocystella mariae. Only two laterals text-fig. 10. Jaekel’s (1899) interpretation of the plate arrangement of Mimocystites bohemicus Barrande with modern notation of plates. B1-B4 basals, IL1-IL5 infra-laterals, L1-L5 laterals, R1-R6 radials. Cf. text-fig. 3. (LI and L5) contribute to the periproct border and LI has a straight upper border with R6 directly adoral to it. This is an unexpected arrangement. Two specimens in the Schary collection (MCZ) apparently show the arrangement figured for Macrocystella mariae. However one specimen of M. mariae (Counterparts BMNH E29109a and 291 10a) has apparently three hexagonal laterals as shown in Jaekel’s figure of Mimocystites bohemicus. It seems possible that the plate arrangement varied slightly in different specimens. Mimocystites azaisi Thoral has a plate arrangement identical to that in Macrocystella mariae. Both the present and Jaekel’s interpretations agree in most respects, particularly in the five plates around the periproct. EXPLANATION OF PLATE 113 Stereophotos of M. mariae Callaway, M. azaisi (Thoral) and M. azaisi multicristata (Thoral). Fig. 2. M. mariae Callaway. Lateral view of lectotype. BU 409. Figs. 1, 3, 5, 8. M. azaisi (Thoral). 1. Anterior lateral view of theca to show well-developed folds in B2 and ornament of stem flanges. CU. 3. Lateral view of another theca. CU. 5. Oral view of same. 8. Lateral view of another theca to show well-developed folds in B2 and ornament of stem flanges. CU. Figs. 4, 6, 7. M. azaisi multicristata (Thoral). 4. Oblique oro-lateral view to show ornament of orals and lateral food grooves alternating in ambulacrum IV (left). CU. 6. Interior view of oral surface of same theca. 7. Lateral view of same theca to show more strongly developed folds in radial plates. All figures of latex impressions whitened with ammonium chloride sublimate. All X 2. Palaeontology, Vol. 11 PLATE 113 PAUL, Macrocystella C. R. C. PAUL: MAC ROCYSTELLA CALLAWAY 591 The thecal plates of Mimocystites bohemicus and M. azaisi are identical to those of MacrocysteUa mariae except in the number of folds, which is variable in each species. The proximal stem of Mimocystites bohemicus is identical to that of MacrocysteUa mariae except that the outer proximals have thicker flanges which are less blade-like. Mimocystites azaisi has still thicker flanges, the peripheries of which are granulose or spinose (PI. 1 1 1 , fig. 2, PI. 113, figs. 1 , 8). in A text-fig. 11. Camera lucida drawing of oral surface of MacrocysteUa azaisi (Thoral). a, entire surface to show oral plates and arrangement of food grooves. Cf. Plate 113, fig. 5. CU. b, detail of gonopore and hydropore area of same. FG, food groove; G, gonopore; H, hydropore; LFG, lateral food groove; M, position of mouth which was probably much larger than shown; 01-07, orals. I-V, ambulacra; l1, I2, I3, facets of ambulacrum I; IV1, IV2, facets of ambulacrum IV. The type species and one other species of Mimocystites therefore exhibit all the features which characterize MacrocysteUa mariae. All three are congeneric. Specimens from the Montagne Noire, France (UC latex impressions) have yielded additional information on the morphology of MacrocysteUa. One example of M. azaisi (Thoral) has an almost complete oral surface (text-fig. 11a, PI. 113, fig. 5). Another 592 PALAEONTOLOGY, VOLUME 11 example of M. azaisi multicristata (Thoral) shows both the internal and external surfaces of the oral area (text-fig. 12, PL 113, figs. 4, 6). There were seven orals, arranged as shown in text-fig. 1 \a. A slit-like hydropore and an oval gonopore are developed across the common suture of Ol and 07 (text-fig. Mb). There are five main ambulacral grooves each of which has lateral branches leading to brachiole facets. Apparently the branches are consistently to the left of ambulacra I and IV in the example of M. azaisi (PI. 113, fig. 5) but regularly alternate in ambulacrum IV in the example of M. azaisi multicristata (PI. 113, fig. 4). Details of the other ambulacra are not well preserved in either specimen. In internal view the mouth is large (4-6 mm.xl-7 mm.) and is covered by ambulacral cover plates. The orals and ambulacral flooring plates are visible (PI. 113, fig. 6). The latter are between, not on, the orals and form part of the thecal wall. Only primary ambulacral flooring plates can be detected. Close to the mouth is a deep pit which apparently connected to the hydropore. This pit is separated from the mouth and the supposed gonopore by two internal ridges, one on each side. The gonopore is ap- parently represented by a small circular pit some distance from the mouth (text-fig. 12). Unfortunately the critical area of the external oral surface is not preserved and it is not possible to match up external and internal openings. Ambulacra IV and V are more deeply impressed than I, II, and III. Another pit is developed obliquely under the aboral of the two internal ridges near the hydropore (PI. 113, fig. 6). The significance of this is unknown. COMPARISON WITH OTHER PELMATOZOA Several authors have grouped Macrocystella with Lichenoides while others have placed them in separate families. In addition Macrocystella has been variously assigned within the Pelmatozoa. The most recent and most complete account of the morphology of Lichenoides is that of Ubaghs (1953) who showed that it completely lacks a stem. The thecal plates are arranged in four circlets but the total number is variable, 5-12 basals, 5 infra-laterals, 5-7 laterals, and 5-7 radials. (The homologies with the Glyptocystitida implied by the use of the same terms for the plate circlets are unjustified.) All the infra-laterals, laterals, and radials bear epispires. These are a type of pore-structure with a single sutural pore which leads to a narrow channel in the external surface of both adjacent plates (see III text-fig. 12. Camera lucida drawing of internal oral surface of Macrocystella azaisei multicristata (Thoral). cf. Plate 1 1 3, fig. 6. CU. G, supposed internal opening of gonopore; H, supposed internal opening of hydro- pore; M, mouth; 01-07, orals; I-V ambulacra. C. R. C. PAUL: MACROCYSTELLA CALLAWAY 593 Ubaghs 1953, figs. 3, 11). There is no lateral periproct in Lichenoides. The brachioles arise from both lateral and radial plates and apparently they are not grouped into five radii. Thus while Lichenoides resembles MacrocysteUa in having definite plate circlets and biserial, unbranched brachioles it differs in the absence of a stem and lateral periproct, in the presence of epispires and brachioles on lateral and radial plates, and in the total number and position of the thecal plates. These differences are considered to be impor- tant taxonomically. The Lichenoidae and Macrocystellidae are maintained as separate families as suggested by Jaekel (1918) and Ubaghs (1953). text-fig. 13. Plate arrangement in Cheirocrinus radiatus Jaekel. Based on Jaekel, 1899, p. 213, fig. 36. B1-B4, basals; IL1-IL5, infra-laterals; L1-L5, laterals; R1-R6, radials. Among cystoids MacrocysteUa most closely resembles Cheirocrinus Eichwald. This latter genus is characterized by a stem with proximal and distal portions. The construc- tion and articulation of the stem are identical to that of MacrocysteUa and the helical arrangement of the articulations of the proximal stem was described by Billings (1858) in Cheirocrinus anatiformis (Hall) (= Glyptocystites logani Billings). The theca of Cheirocrinus is composed of 27 plates arranged in 5 circlets: 4BB, 5ILL, 5LL, 6RR, and 700. The periproct is large, surrounded by 5 thecal plates (1L4, IL5, LI, L4, and L5) and was covered by a flexible plated integument in life. L5 is directly adoral to the periproct and has R5 directly adoral to it (text-fig. 13). The subvective system is restricted to the flat oral surface and there are 20-5 brachioles grouped into 5 ambulacra whose flooring plates lie between the orals, not on them. Cheirocrinus differs from Macro- cysteUa in the possession of pectinirhombs. MacrocysteUa and Cheirocrinus have in common many distinctive features of which perhaps the most important is the detailed structure of the stem. This type of stem is characteristic of and confined to the superfamily Glyptocystitida. It is most unlikely that such a complex organ developed independently in two groups which share other common characteristics. MacrocysteUa probably gave rise to Cheirocrinus and through it to the other Glyptocystitida as originally suggested by Jaekel (1899). 594 PALAEONTOLOGY, VOLUME 11 Thoral (1935) thought it possible to recognize in the Montague Noire a lower horizon with Macrocystella azoisi (upper Tremadoc) and a higher horizon with Cheirocrinus languedocianus (basal Arenig). In Britain M. marine occurs below the oldest Cheiro- crinus. Macrocystella seems to be a characteristic fossil of the Tremadoc whereas the oldest known Cheirocrinus are all Arenig. Stratigraphic evidence agrees with the idea that Macrocystella evolved into Cheirocrinus. In the past the main objections to the inclusion of Macrocystella in the Rhombifera were its branched arms and lack of rhombs. The former was an error but the latter is more important. Regnell (1945) has stressed that the main character which unites the cystoids as a class is the presence of pore-structures. Detailed study of cystoid pore- structures (Paul, 1968) suggests the Rhombifera should be regarded as a distinct class. The rhomb-less Macrocystella is included in the Rhombifera on the same grounds that led Kesling (1963) to include Amecystis in the Rhombifera. Amecystis is effectively a Pleurocystites without pectinirhombs, just as Macrocystella is a Cheirocrinus without pectinirhombs. The many similarities outweigh this single distinction. THE EVOLUTION OF PECTINIRHOMBS The rhombs of Cheirocrinus are fully developed pectinirhombs (which Bather regarded as a highly specialized type of rhomb) even in the earliest species known. There is no evidence of a gradual evolution of pectinirhombs. If Macrocystella evolved into Cheiro- crinus, pectinirhombs either appeared ‘ suddenly ’ or pre-existing structures broke through the thecal plates to appear as pectinirhombs. The internal surfaces of isolated plates in both Macrocystella and Cheirocrinus show no features which could be incipient pectini- rhombs. Pectinirhombs were functional throughout their growth; they are present as external features from the earliest stages. They did not develop internally and become external features later in growth. Rather sudden appearance therefore seems more likely. All rhombs have generally been accepted as respiratory organs. In the simplest case respiration would have taken place through the thecal wall. The amount of oxygen required would have been proportional to the volume, and the amount of respiratory exchange to the surface area, of the theca. The oxygen requirements would thus increase with growth faster than the amount of exchange. This difficulty can be overcome, without materially altering the over-all thecal shape, by the production of evaginations or invaginations of the thecal wall. Macrocystella has the former in the folds of the thecal plates. The dichopores of pectinirhombs are invaginations and produce a slightly better volume to surface area ratio. Exchange is facilitated by a large surface area and a thin exchange surface. Either invaginations or evaginations are almost equally effective in increasing the surface area but the latter are exposed and liable to mechanical damage. The ridges in the thecal plates of Macrocystella probably facilitated exchange by increasing the surface area. The thecal plates were extremely thin (OT mm.) and although strengthened by the ridges they were still very fragile. The dichopores of pectinirhombs are within the theca and there- fore protected. Dichopore walls are much thinner than thecal plates (usually 0-01 mm.). In Macrocystella the entire thecal wall probably took part in exchange. In Cheirocrinus a differentiation of function is seen. Without decreasing the amount of exchange it C. R. C. PAUL: MACROCYSTELLA CALLAWAY 595 became possible to have much thicker and stronger thecal plates. Exchange was re- stricted to specialized areas, namely the pectinirhombs. The thecal plates of Cheiro- crinus, especially in early species like C. languedocianus, are still thin compared with later glyptocystitids but they are thicker (usually more than 0-5 mm.) than those of Macrocystella. text-fig. 14. Pectinirhombs of a specimen of Cheirocrinus languedocianus Thoral. Note the incomplete pectinirhombs. Based on camera lucida drawings of the individual plates. CU. Symbols as in text-fig. 3. In Macrocystella all thecal plates are ridged and all probably contributed to respira- tion. However the distribution of ridges in M. azaisi is uneven. More and better- developed ridges occur on radial plates (PI. 113, fig. 7) and associated with B2 (PI. 113, figs. 1, 8). This is quite independent of the variation between M. azaisi and M. azaisi multicristata. Sdzuy (1955, figs. 1 d-g) figured a similar concentration of ridges on B2 in M. bavarica. In Cheirocrinus languedocianus , probably the earliest known species of Cheirocrinus , dichopores are developed across more sutures than in later species (text-fig. 14). Many sutures bear one or two demi-rhombs and some have only a few randomly spaced dichopores. These latter form incomplete pectinirhombs which have not been recorded in later glyptocystitids. Clearly the arrangement is more random than 596 PALAEONTOLOGY, VOLUME 11 in later species which have a reduced number of pectinirhombs. All species of Cheiro- crinus have pectinirhombs on radials and on B2 however. The similarity between the distribution of ridges in Macrocystella and of pectinirhombs in Cheirocrinus indicates a concentration of respiratory activity in the same areas of their thecae, probably reflecting a similar internal organization, and confirming the idea that Macrocystella and Cheirocrinus are related. MODE OF LIFE The stem of Macrocystella lacks a root structure or anchoring device and there is little evidence for Bather’s reconstruction of a distal coil to the stem. Whatever benefits were conferred on Macrocystella by the possession of a well-developed stem, they were not associated with permanent attachment. M. mariae and other species of Macro- cystella are found in fine-grained sediments in which there is little evidence for suitable attachment surfaces. Nevertheless the general morphology of the theca and subvective system suggests Macrocystella habitually held its theca upright. Only free-floating or free-swimming modes of life can satisfy both lines of reasoning. The distal stem in Macrocystella and in other early glyptocystitids tapers gradually to a small diameter and ends abruptly. The terminal portion of the stem is not modified as in Brockocystis Foerste or Lepocrinites Conrad and there is no way to be certain that the last distal columnal preserved was the terminal columnal in life. It is possible that Macrocystella broke free prior to death and was buried some distance from its point of attachment. However M. mariae and other apparently free glyptocystitids are frequently preserved in clay-grade sediments with little associated fauna which could have acted as substrata for attachment. No Macrocystella , Cheirocrinus nor Pleurocystites has ever been found with a root structure or anchoring device, yet many specimens are excellently preserved with brachioles, periproctal membranes, and other delicate structures intact. Some specimens from the Trenton Limestone of Ottawa and the Starfish Bed of Girvan, Scotland (for example) suggest that death was due to burial alive and therefore that the specimens are complete. The terminal diameter of the distal stem (0-5 mm.) compared with the total length (50 mm.), in M. mariae indicates that if attached the stem was extremely weak at this point. Macrocystella may have been attached early in its develop- ment but the evidence strongly suggests that it was free for the major portion of its life. Pentameral symmetry is well developed in Macrocystella'. there are circlets of five plates and five ambulacra. The theca is cylindrical not flattened on one side. Bather argued that pentameral or radial symmetry was developed in fixed animals and he associated departures from pentamery in echinoderms, particularly flattening of the theca, with departures from an upright fixed position. Pleurocystites , another free glyptocystitid, has a markedly flattened theca which apparently lay on the sea floor. If Macrocystella and Cheirocrinus also lay on the sea floor one would expect a similar flattening of their thecae. The persistance of Macrocystella and Cheirocrinus after the appearance of Pleurocystites suggests the former were adapted to a different mode of life. The subvective system in Macrocystella and Cheirocrinus is confined to the oral surface and forms a cone of collection. Paul (1967#) suggested that such an arrangement is better adapted to collect fa ling food particles in a relatively still sea and requires an upright theca. Both Macrocys el/a and Cheirocrinus have respiratory surfaces developed C. R. C. PAUL: M AC ROCYSTELLA CALLAWAY all round the theca and some of these would have been fouled if the theca rested on the sea floor. Plewocystites however has all its pectinirhombs on one side of the theca, presumably the upper side in life. All these lines of evidence indicate that Macrocystella and Cheirocrinus habitually held their thecae upright. However, when upright, Macrocystella is very top heavy (text-fig. 15) and the theca would have rested on the sea floor if the stem were free. Without a root structure or anchoring device the stem could exert no leverage against the substrate and would have been useless to maintain the theca upright. The distal stem may have been partially buried in the substrate but its whiplike form would provide little resistance in soft mud. There is no evidence for partial burial of the stem during life in specimens which apparently died by entombment. Other pelmatozoans which were fixed in soft sediments have well-developed branching root structures (e.g. Eucalyptocrinites and Caryo- crinites). Even recent brachiopods which live in soft sedi- ments have branching root-like pedicles. If Macrocystella and Cheirocrinus were free and habitually held their thecae up- right they could only achieve this if they floated or swam actively. Both modes of life are known in other pelmatozoans. Recent free-swimming crinoids have negative buoyancy (i.e. they are denser than sea water) and this has generally been assumed to obtain in all echinoderms. If the coelomic fluids and organic tissues are of neutral buoyancy, the only element which contributes to negative buoyancy is the skeleton, the effective density of which is 0-8 gm. per cc. when fully im- mersed in sea water. An echinoderm with negative buoyancy will sink when at rest. Recent free crinoids maintain their position by swimming. They are characterized by a well- developed subvective system with multiple branched arms and, in most cases, by the complete absence of a stem and marked reduction of the calyx: certainly none has a calyx comparable in size to a cystoid theca. Swimming is wholly achieved by the arms, buoyancy devices are absent (there is nowhere to house them) and dead weight in the form of stem and thecal plates is minimal. In Macrocystella there is a relatively undeveloped subvective system, a large theca, and a well-developed stem. Dead weight was considerable and the brachioles may have been rela- tively inefficient organs of locomotion. The flattening of the brachioles in M. marine is in the opposite direction to that one would expect if they were used actively in swimming. text-fig. 15. A reconstruction of Macrocystella mariae Callaway. Br, brachiole; DS, distal stem; Pe, periproct; PS, proximal stem. 598 PALAEONTOLOGY, VOLUME 11 However if the lappets were folded in during an upstroke and extended during a down- stroke the brachioles could have been fairly efficient swimming organs. The relatively large theca has very thin plates and could well have housed buoyancy devices. Quite small gas bubbles would have significantly altered the total buoyancy: in M. mariae a bubble 3 mm. in diameter would completely compensate for the weight of a theca 10 mm. in diameter and 15 mm. high. Brockocystis Foerste, another apparently free glypto- cystitid, has hollow, bulbous thecal plates each of which could have housed a gas bubble. The stem bulb in Brockocystis is also hollow. It is not impossible that Macro- cystella had buoyancy chambers but there is no direct evidence for this. Function of the stem. Most authors, with the notable exception of E. Kirk (1911), have taken stems in pelmatozoans to imply permanent fixture. It is therefore surprising to find among early stemmed pelmatozoans an almost total lack of definite root struc- tures or anchoring devices. What benefit did the possession of a stem confer if it was not attachment ? If an ancestral stem-less form were permanently attached, the development of a stem could have raised the theca off the substrate. This would reduce the chances of accidental burial, of fouling by mud-laden currents, and it could have placed the echinoderm above other benthonic organisms competing for food particles as suggested by Kesling and Mintz (1961). What may have been attempts to raise the theca, or at least the oral surface, are seen in edrioasteroids and some diploporites but none of these produces a true stem, only aboral stem-like projections of the theca. Alternatively if an ancestral stem-less form were free, a stem would weight the theca aborally. This would maintain the animal upright if the theca were buoyant or if the subvective system were used in swim- ming. However, a simple weight placed aborally would satisfy this requirement. The form of the stem in Macrocystella could have allowed considerable variation in buoyancy when the stem was in contact with the substrate. If the theca was buoyant but the weight of the stem gave the whole animal negative buoyancy, equilibrium could have been reached if part of the stem rested on the sea floor. The flexible proximal stem would still allow the theca to be held upright; buoyancy would hold the theca upright without any effort on the part of the animal. Macrocystella probably rested with its theca near but not on the substrate and moved about by means of its brachioles and stem. This interpretation seems preferable to one without a buoyant theca even though there is no direct evidence for buoyancy devices. If Macrocystella used its brachioles to maintain an upright posture it could not have rested in an upright position. As recent free crinoids feed when at rest it is difficult to imagine how Macrocystella fed under these circumstances. Macrocystella is imagined to have had a slightly buoyant theca but to have been weighed down by the stem. Under these circumstances it could have rested with the stem on the sea floor, the theca upright and the brachioles extended to feed. In times of short food supply Macrocystella swam away actively by means of its brachioles and stem, moving just above the sea floor. Macrocystella may have drifted under the influence of gentle currents but current action was not strong in the environment of deposition of M. mariae. Occasionally M. mariae was overwhelmed by a sudden influx of mud possibly brought about by turbidity currents. Macrocystella seems to have lived in much deeper water than most later glyptocystitids (see Paul 1967a). C. R. C. PAUL: MACROCYSTELLA CALLAWAY 599 REFERENCES barrande, J. 1867. Systeme silurien du centre de la Boheme. le partie: recherches paleontologiques, 3. Prague and Paris. 1868. Silurische Fauna aus der Umgebung von Hof in Bayern. Neues Jb. Miner. Geo!. Pa/dont. (for 1868), 641-96, 2 pi. (also a separately published edition in French with different pagination). 1887. Systeme silurien du centre de la Boheme. \e partie: recherches paleontologiques, 7. Leipzig and Prague. bassler, r. s. and moodey, m. w. 1943. Bibliographic and faunal index of Paleozoic pelmatozoan echinoderms. Spec. Pap. geol. Soc. Am. 45, vi+733 pp. bather, f. a. 1899. A phylogenetic classification of the Pelmatozoa. Rep. Br. Ass. Advmt Sci. 68, 916-23. 1900. The echinoderms. in lankester, e. r. A treatise on zoology, 3. London. billings, e. 1858. On the Cystideae of the Lower Silurian rocks of Canada. In Figures and descriptions of Canadian organic remains, Decade 3, 9-74, pi. 1-8. Montreal. callaway, c. 1877. On a new area of Upper Cambrian rocks in South Shropshire, with a description of a new fauna. Q. J! geol. Soc. Lond. 33, 652-72, pi. 24. carpenter, p. h. 1891. On certain points in the morphology of the Cystidea. J. Linn. Soc. (Zool.) 24, 1-52, pi. 1. castell, c. p. (compiler) 1964. British Palaeozoic Fossils. B.M. (N.H.) London. choubert, h., termier, g., and termier, h. 1953. Presence du genre Mimocystites Barrande dans l'Ordovicien du Maroc. Notes Mem. Serv. Mines Carte geol. Maroc, 117, 137-43. cuenot, l. 1948. Classe des Cystides. In grasse, p., Traite de Zoologie, 11, 15-25. Paris. 1953. Echinodermes. Classes des Cystides. In piveteau, j., Traite de Paleontologie, 3, 607-28. Paris. fearnsides, w. g. 1905. On the geology of Arenig Fawr and Moel Llyfnant. Q. Jl geol. Soc. Lond. 61, 608-37, pi. 41. 1910. The Tremadoc slates and associated rocks of South-east Caernarvonshire. Ibid. 66, 142-87, pi. 15-17. haeckel, E. 1896. Amphorideen und Cystideen. In Festschrift zum siebenzigsten Geburt stage von Carl Gegenbaur am 21st. August 1896, 1, 1-80, pi. 1-5. Leipzig. jaekel, o. 1899. Stammesgeschichte der Pelmatozoen, 1, Thecoidea und Cystoidea. Berlin. 1918. Phylogenie und System der Pelmatozoen. Paldont. Z. 3, 1-128. kesling, r. v. 1963. Key for the classification of cystoids. Contr. Mus. Paleont. Univ. Mich. 18, 101-16. and mtntz, l. w. 1961. Notes on Lepadocystis moorei (Meek), an Upper Ordovician callocystitid cystoid. Contr. Mus. Paleont. Univ. Mich. 17, 123-48, 7 pi. kirk, e. 1911. The structure and relationships of certain eleutherozoic Pelmatozoa. Proc. U.S. natn. Mus. 41, 1-137, pi. 1-11. melendez, b. m. 1944. Nuevos datos para la estratigrafia del Paleozoico Aragones. Boln R. Soc. esp. Hist. nat. 42, 129-50, pi. 24-9. moore, r. c. 1954. Echinodermata : Pelmatozoa. In kummel, b., Status of invertebrate paleontology. Bull. Mus. comp. Zool. Harv. 112, 125-49. paul, c. r. c. 1967n. The British Silurian cystoids. Bull. Br. Mus. nat. Hist. (Geol.) 13, 297-356, 10 pi. 1961b. Cystidea Barrande 1868 (Cystoidea: Glyptocystitida) : proposed suppression under the Plenary Powers. Bull. zool. Nom. 24, 304-7. 1968. The morphology and function of dichoporite pore-structures in the cystoids. Palae- ontology, 11, 697-730, pi. 134-140. pompeckj, J. f. 1896. Ein neuentdecktes Vorkommen von Tremadoc-Fossilien bei Hof. Ber. nordober- frank. Ver. Naturk. 1, 89-101. prokop, r. j. 1966. Class Eocrinoidea. In spinar, z. et a/., Systematicka paleontologie bezobratlych, pp. 817-20, figs. regnell, g. 1945. Non-crinoid Pelmatozoa from the Paleozoic of Sweden. A taxonomic study. Meddn Lunds geol-miner. Instn, 108, 255 pp., 15 pi. 1948. An outline of the succession and migration of non-crinoid pelmatozoan faunas in the Lower Paleozoic of Scandinavia. Ark. Kemi Miner. Geol. 26, (13), 1-55. 600 PALAEONTOLOGY, VOLUME 11 sdzuy, k. 1955. Cystoideen aus den Leimitz-Scheifern (Tremadoc). Senckenberg. leth. 35, 269-76, 1 pi. springer, f. 1913. Cystoidea. In Eastman, c. r. (ed.), Textbook of Palaeontology, 1, 145-61. London. Stubblefield, c. j. and bulman, o. M. b. 1927. The Shineton Shales of the Wrekin district: with notes on their development in other parts of Shropshire and Herefordshire. Q. Jl geol. Soc. Load. 83, 96-145, pi. 3-5. termier, g. and termier, H. 1952. Histoire geologique de la biosphere. 721 pp. 8 pi. Paris. 1954. A propos de la structure des Eocrino'ides, C. r. Seanc. Soc. geol. Fr. (1954) 4, 92-4. thoral, M. 1935. Contribution a V etude paleontologique de FOrdovicien inferieur de la Montague Noire et revision sommaire de la faune cambrienne. 362 pp. 35 pi. Montpellier. ubaghs, G. 1953. Notes sur Lichenoides priscus Barrande, eocrinolde du Cambrien Moyen de la Tchecoslovaquie. Bull. Inst. r. Sci. nat. Belg. 29 (34), 1-24. 1968. Eocrinoids. In moore, r. c. (editor) Treatise on invertebrate paleontology, S. Echinodermata 1, S455-S495. zittel, k. A. von, 1880. Handbuch der Palaeonto/ogie. 1. Miinchen and Leipzig. C. R. C. PAUL Geology Department Indiana University Northwest Gary, Indiana Typescript received from author 10 January 1968 A REVISION OF SOME UPPER DEVONIAN FORAMINIFERA FROM WESTERN AUSTRALIA by james e. conkin and Barbara m. conkin Abstract. Arenaceous foraminifera described by Crespin (1961) from the Upper Devonian Virgin Hills and Gogo Formations of the Fitzroy Basin of Western Australia are revised. Four genera and eight species are recognized; these include Oxinoxis Gutschick 1962 and Sorosphaeroidea Stewart and Fampe 1947, which were previously recorded only from the Middle Palaeozoic of the United States. The similarity of the fauna to Upper Devonian foraminiferal faunas in the United States and Europe is described. Several new occurrences of foraminifera are reported from the Silurian and Devonian of New South Wales, Victoria, and Tasmania. In 1964-5, we had the opportunity to collect from the Palaeozoic rocks of Victoria, New South Wales, and Tasmania, south-eastern Australia, in order to assess the stratigraphic occurrence and geographic distribution of arenaceous foraminifera. A preliminary report on the Permian foraminifera of Tasmania has been made (Conkin and Conkin 1965a). We were able, through the kindness of Dr. Irene Crespin and Dr. D. J. Belford of the Bureau of Mineral Resources, Geology, and Geophysics at Canberra to study the types of Dr. Crespin’s (1961) Upper Devonian foraminifera from the Virgin Hills and Gogo Formations of the Fitzroy Basin, Western Australia. It was discovered that some of the forms would require revision in the light of work done on arenaceous foraminifera after the completion of Dr. Crespin’s work. With these revisions made, the relationships between this important fauna of Crespin (1961) and similar faunas in the United States and Europe can be more clearly seen. Past work. Crespin (1961, p. 397) summarized the record of Devonian foraminifera in Australia as follows: In 1918 Chapman described and figured three genera and five species of what he considered to be foraminifera from thin sections of a limestone of Devonian age from the Nemingha area, Tamworth, northern New South Wales. However, Wood (1957) stated that these forms were ‘oolite grains more or less affected by dolomitization and by mechanical distortion’. Parr (in Teichert and Talent 1958) reported poorly preserved arenaceous foraminifera from Devonian rocks in the Buchan area, east Gippsland, Victoria. Chapman (1933) described two species of foraminifera from the Silurian of Victoria, ‘ Trochammina bursaria ’ and ‘ Hemigordius lilydalensis ’ ; the former is perhaps conspecific with Thuramminoides sphaeroidalis Plummer 1945, which may not be a foraminiferan, but possibly a radiolarian or spore-like body (Conkin, Conkin, and Canis 1968); the latter is an indeterminate arenaceous foraminiferan and certainly not the calcareous Hemigordius or even a siliceous replacement of Hemigordius. Crespin’s (1961) fauna, the first important foraminiferal fauna from the Devonian of Australia, contains four genera and eight species as revised in the present paper. [Palaeontology, Vol. 11, Part 4, 1968, pp. 601-9, pis. 114-17.] 602 PALAEONTOLOGY, VOLUME 11 Present work. Silurian and Devonian foraminifera were observed by us in the following states and stratigraphic horizons : Tasmania, upper part of the Middle Devonian Point Hibbs Limestone at Point Hibbs: Sorospha- eroidea sp. Victoria, east Gippsland, Devonian Pyramid Member of the Buchan Cave Limestone at Bindi : Toly- pammina spp., Sorosphaeroidea sp., and Hyperammina ? sp. New South Wales, Yass-Taemas area, Devonian Spirifer yassensis beds: Hyperammina sp. and Colon- ammina ? sp. Victoria-New South Wales border area, Silurian Cowombat Formation at Cowombat Flat: Hyper- ammina sp., Tolypammina sp., and Psammosphaeral sp. The presence of these foraminifera in the Palaeozoic of Australia shows the need for further work on this group along the lines of Crespin (1958, 1961). We are in the process of describing these foraminifera from the Silurian and Devonian of Australia as well as making a detailed study of the magnificent arenaceous foraminiferal faunas from the Permian of Tasmania. DISCUSSION OF GENERA Saccammina Sars 1869. This genus consists of a simple agglutinated chamber with an aperture which may or may not have a short neck; as thus understood, Saccammina has been reported from Silurian to Recent. Sorosphaeroidea Stewart and Lampe 1947 has a distinctive aperture which was not originally noticed by Stewart and Lampe or subsequently by Summerson (1958). Oxinoxis Gutschick 1962, emended Conkin and Conkin 1964. The range of Oxinoxis, previously reported only from the United States, is upper Middle Devonian (middle and upper Hamiltonian) to Lower Mississippian (Kinderhookian and lower Osagean). Middle Devonian forms are almost exclusively single chambered and small. Upper Devonian forms are single to multiple chambered, and small to large. Kinderhookian and lower Osagean forms are generally much like the Upper Devonian, although there is in addition a distinct single chambered species of Oxinoxis in the Kinderhookian (Conkin, Conkin, and Canis 1968). Tolypammina Rhumbler 1895. This attached genus ranges from Ordovician to Recent. Pre-Upper Devonian forms appear to wind or twist in a haphazard, disorganized manner, while some of the post-Middle Devonian forms wind, at least partially, in a more organized pattern, it is probable that the even more ‘organized’, but similar genus, Ammovertella Cushman 1928, which apparently arose in the Kinderhookian, was derived from Tolypammina. Hyperammina Brady 1878 is known from Ordovician to Recent. It is not present in Crespin’s Upper Devonian fauna (1961), but occurs in the Silurian Cowombat Formation in the New South Wales-Victoria border area. The restricted generic definition of Hyperammina of Loeblich and Tappan (1964, p. C190) is not used. Hyperammina is employed in the sense of Brady 1878, as emended by Conkin (1954, 1961), and discussed by Conkin and Conkin (19656, p. 211). It is basically a test with a free proloculus and CONKIN AND CONKIN: REVISION OF UPPER DEVONIAN FORAMINIFERA 603 a free second chamber of more or less tubular shape, tapering and/or slightly constricted in some species. Hyperammina has great potential as a tool in stratigraphic palaeontology within the Palaeozoic; a beginning along this line has been made within the Lower Mississippian of the United States (Conkin 1961). Psammosphaera Schulze 1875 consists of a simple agglutinated spherical or globular chamber with no obvious aperture. It has been reported from Ordovician to Recent. AGE OF THE VIRGIN HILLS FORMATION BASED ON FORAMINIFERA Crespin (1961) reported only two specimens from that portion of the Upper Devonian of Western Australia which may be referred to the Gogo Formation. Most of her specimens came from the overlying Upper Devonian Virgin Hills Formation. Considered strictly in terms of the known stratigraphic ranges of foraminiferal genera included, the Virgin Hills Formation could be of late Middle Devonian, Late Devonian, or Early Mississippian (Kinderhookian) age. Sorosphaeroidea, previously thought by us to be restricted to beds older than Mississippian, has now been discovered in rare numbers in the Kinderhookian McCraney Limestone of north-eastern Missouri. Oxinoxis has not been found below the upper Middle Devonian, but it ranges into the lower Osagean. On a specific level, however, Tolypammina helina, with a test consisting of an initially coiled portion and a final extended portion, is an ‘organized’ kind of Tolypammina which is post-Middle Devonian in its stratigraphic occurrence. The foraminiferal fauna of the Louisiana Limestone (Upper Devonian) from Missouri and Illinois is the oldest fauna from which this type of Tolypammina has been described (Conkin and Conkin 1964) in North America; a similar species is found in the Upper Devonian of Germany (Blumenstengel 1961). Thus, T. helina places the Virgin Hills Formation in the Upper Devonian or Lower Mississippian, rather than the Middle Devonian. SYSTEMATIC PALAEONTOLOGY The following observations are based on a study of Dr. Crespin’s type specimens of Upper Devonian foraminifera at the Bureau of Mineral Resources, Geology and Geophysics, Canberra, A.C.T., Australia, and on a discussion of the fauna with Dr. Crespin. Text-fig. 1 is based on drawings made at that time and on photomicrographs taken subsequently and forwarded to us by Dr. D. J. Belford. Type numbers refer to Commonwealth Palaeontological Collection (CPC), and holotype and para- type designations are the original ones of Crespin (1961). Family saccamminidae Genus saccammina M.Sars 1869 Saccammina glenisteri Crespin 1961 Plate 1 16, figs. 1-7. 1961 Saccammina glenisteri Crespin, pp. 401, 402, pi. 65, figs. 3-7. The holotype (CPC 401 ; Pi. 1 16, figs. 1-3) has slit-like? aperture as figured by Crespin, and a few holes of secondary nature; the test, one grain thick, is composed of coarse grains of quartz in a fine- grained matrix; there is no evidence of attachment. R r C 5934 604 PALAEONTOLOGY, VOLUME 11 Paratype A (CPC 402; PI. 116, figs. 6, 7) has a coarse-grained, subglobular test with an elongate, lunate aperture; there is no evidence of attachment. Paratype B (CPC 403) has a rounded aperture; there is no evidence of attachment. Paratype C (CPC 404; PI. 1 16, figs. 4, 5) has a knob-like structure on the inside of the broken test which is like the structure found on the interior of the test and just below the aperture in Soros- phaeroidea, but which in this case may be secondarily formed. Paratype D (CPC 405) consists of two globular tests attached to a Tolypammina fragment and composed of coarse grains of quartz; the apertures are indistinct, and the tests are smaller than the other types of this species; this specimen came from the Gogo Formation which underlies the Virgin Hills Formation from which the other specimens came. The identity of Paratype D is not certain since the tests are attached on a small surface and are thus not completely free. Genus sorosphaeroidea Stewart and Lampe 1947 Sorosphaeroidea adhaerens (Crespin) 1961 Plate 116, figs. 8-11; text -fig. lc 1961 Sorosphaera adhaerens Crespin, pp. 403, 404, pi. 66, figs. 1-5. The holotype (CPC 406; Pi. 1 16, figs. 8, 9) consists of chambers which are attached to a Tolypammina fragment and the nature of the surface of attachment on the test (whether covered with a floor or open), is not known. Its reference to Sorosphaeroidea is uncertain, but probably correct. Paratype A (CPC 407) consists of several chambers which are indistinct on one side, and attachment areas are somewhat obscure. Paratype B (CPC 408; PI. 1 16, figs. 10, 11 ; text-fig. lc) possesses broad, flattened attachment scars (not originally figured), which are somewhat finer grained than the upper side of test; there is a definite aperture on the upper, unattached side of each chamber. A fragment of a Tolypammina is attached to the specimen. Paratype C (CPC 409) is attached to a Tolypammina fragment, but the attached surface is not visible; an aperture is present on the upper side. Paratype D (CPC 410) has an aperture on the upper unattached side; the attached side is not visible as it is fastened to a fragment of Tolypammina. Remarks. Paratype B is definitely referable to Sorosphaeroidea ; the other types have the attachment surface hidden, or it is obscure (Paratype A), but they probably belong to this genus also. We place this species in Sorosphaeroidea because of the large areas of attachment, and because the apertures generally are like those which we have found in the original types of Sorosphaeroidea of Stewart and Lampe; the aperture appears inverted, as if pushed through from the outside. EXPLANATION OF PLATE 114 Figs. 1-4,9. Tolypammina devoniana (Crespin) 1961. 1, Paratype A; basal, attached side showing rather large attachment scars, smoothly polished along the edges. 2, Paratype D, top view. 3, Side view of holotype showing attachment scar on the proloculus and small areas of attachment along the length of the second chamber. 4, Top view of holotype showing slight notch on proloculus. 9, Paratype E, top view showing break in test. Figs. 5, 6. Tolypammina ? sp. 5, Plesiotype (holotype of ‘ Rhabdammina virgata ’ Crespin 1961), showing fragmentary nature and coarse-grained texture of the test. 6, Reverse side of same specimen; proloculus and aperture are not preserved. Figs. 7, 8. Tolypammina sp. 7, Plesiotype of ‘ Marsipella sp.’, Crespin 1961 ; top view of hemitubular fragment showing rough and coarse-grained test. 8, Basal view of same specimen showing attached lower surface. All figures approximately X 105. Palaeontology, Vol. 11 PLATE 114 9 C0NK1N and CONK1N, Upper Devonian arenaceous foraminifera CONKIN AND CONKIN: REVISION OF UPPER DEVONIAN FORAMINIFERA 605 The Middle Devonian forms of Sorosphaeroidea described by Stewart and Lampe (1947) and Summerson (1958) are more regularly shaped and arranged than the present forms. text-fig. 1. Schematic drawings of attached sides of Tolypammina devoniana(A, e), Oxinoxis ampullacea (b, d, f), and Sorosphaeroidea adhaerens (c). All figures approximately X 50. Hatched areas are flattened surfaces of attachment. Black areas indicate openings into test. White areas are unattached portions of test. Family reophacidae Genus oxinoxis Gutschick 1962, emend. Conkin and Conkin 1964 Oxinoxis ampullacea (Crespin) 1961 Plate 117, figs. 1-11; text-fig. 1b, d, f 1961 Lagenammina ampullacea Crespin, pp. 404, 405, pi. 66, figs. 6-8. 1961 Colonammina imparilis Crespin, pp. 405, 406, pi. 65, figs. 10-13. 1961 Proteonina sp., Crespin, p. 404, pi. 65, fig. 8. The holotype of ‘’Lagenammina' ampullacea (CPC 412; PI. 117, figs. 1-3) has a faint flattened attachment scar at the base, but otherwise the chamber is nearly spherical with a small apertural neck. Paratype A of ‘ Lagenammina ' ampullacea (CPC 413; PI. 117, fig. 11) is slightly flattened on the basal side, with a somewhat finer-grained texture there than on the upper side, but with some larger quartz grains studded on it; there is a small apertural neck. Paratype B of ‘ Lagenammina ' ampullacea (CPC 414) has a rather prominent apertural neck with the basal side flattened; there is also a small, slightly roughened attachment area at the base of the chamber. The holotype of ‘ Colonammina imparilis' (CPC 415; PI. 117, figs. 4-6; text-fig. Id) consists of a nearly spherical chamber with a prominent wedge-shaped attachment scar forming a re-entrant into the test below the neck; the prominent apertural neck is hemitubular with the basal surface flattened and attached. Paratype A of ‘ Colonammina imparilis' (CPC 416) has a flattened attachment scar on the basal side which is finer grained than the unattached, upper, convex side; the prominent tubular apertural neck has no attachment scar. Paratype B of ‘ Colonammina imparilis' (CPC 417; PI. 117, figs. 7, 8; text-fig. 1b) has a large basal attachment scar which is fine grained and has a polished lustre; the apertural neck is only partly attached on its basal side; the convex upper unattached side of the test is coarse grained. Paratype C of ‘ Colonammina imparilis' (CPC 418) is an obscure specimen; the apertural neck figured by Crespin (1961, pi. 65, fig. 13) appears to be part of another specimen and thus foreign to paratype C; the base of the test is flattened and has a prominent fine-grained attachment scar with a polished lustre; this specimen is not typical of the species, but does apparently belong to O. ampullacea. 606 PALAEONTOLOGY, VOLUME 11 ‘ Proteonina sp.’ (CPC 411; PI. 117, figs. 9, 10; text-fig. If) is a compressed test with holes and depressions in the wall left after calcareous grains were leached out; the apertural neck is partly hemitubular and has a small attachment scar on the basal side at the apertural end, and after which the neck is tubular adjacent to the chamber, which is itself unattached; this specimen is from the Gogo Formation. Remarks. The presence of an attachment scar on the chamber of the test and/or on the hemitubular to tubular apertural neck serves to place these specimens in Oxinoxis. O. ampuUacea differs from the type species O. ligula (Gutschick, Weiner, and Young 1961), which occurs in the Upper Devonian and Lower Mississippian in the United States, in having a more coarse-grained test and in being nearly twice as large as the single-chambered, more solidly-walled forms of O. ligula. Family ammodiscidae Genus tolypammina Rhumbler 1895 Tolypammina devoniana (Crespin) 1961 Plate 114, figs. 1-4, 9; text-figs. 1a, e 1961 Hyperammina devoniana Crespin, pp. 406, 407, pi. 64, figs. 1-6. The holotype (CPC 419; PI. 114, figs. 3, 4) has a proloculus which is asymmetrical, shaped rather like the head of a golf club, owing to a notch-like attachment scar on its basal surface where it joins the second chamber; this basal side of the proloculus is partially attached and there are attachment scars at irregular intervals and distances along the elongate second chamber. EXPLANATION OF PLATE 115 Figs. 1-4. Tolypammina helina Crespin 1961. 1, Holotype, basal flattened and attached side of hemi- tubular test. 2, Top view of holotype showing proloculus, 2\ planispiral whorls, and extended por- tion. 3, Paratype A, basal view showing the completely attached hemitubular test, and an early unsuccessful attempt to uncoil (after 1J planispiral whorls) with return to planispiral coiling for 1 1 whorls more before finally uncoiling. 4, Top view of paratype A, showing proloculus, rather obscure whorls, and extended portion. Figs. 5-8. Tolypammina nexuosa Crespin 1961. 5, Paratype B of Tolypammina ‘ helina', top view. 6, Basal view of same specimen showing complete attachment of hemitubular test; no proloculus preserved. 7, Holotype of Tolypammina nexuosa, top view. 8, Basal view of holotype, showing attached side of test. All figures approximately x 105. EXPLANATION OF PLATE 116 Figs. 1-7. Saccammina glenisteri Crespin 1961. 1, Holotype, upper apertural view, showing slit-like aperture in centre of test. 2, Basal view of holotype. 3, Side view of holotype. 4, Paratype C, broken test showing knob-like structure underneath the aperture on the interior of the test. 5, Exterior view of paratype C showing aperture in the centre of the test. 6, Paratype A, upper apertural view of test showing slit-like aperture in the centre of the test. 7, Basal view of paratype A. Figs. 8-11. Sorosphaeroidea adhaerens (Crespin) 1961. 8, Holotype, view of aggregate of chambers shown by Crespin (1961, pi. 66, fig. 1) showing attachment to foreign body. 9, Reverse side of specimen. 10, Paratype B, upper unattached surface of test showing aperture piercing middle chamber and an attached fragment of Tolypammina. 1 1 , Basal view of paratype B showing flattened chambers on attached side of test. All figures approximately X 105. Palaeontology, Vol. 11 PLATE 115 CONKIN and (TONKIN, Upper Devonian arenaceous foraminifera Palaeontology , Vol. 11 PLATE 116 CONKIN and CONKIN, Upper Devonian arenaceous foraminifera CONKIN AND CONKIN: REVISION OF UPPER DEVONIAN FORAMINIFERA 607 Paratype A (CPC 420; PI. 1 14, fig. 1) is like the holotype; a prominent attachment scar beginning about midway along the proloculus and smoothly polished along the edges, extends approximately half the length of the test. Paratype B (CPC 421) consists of a nearly symmetrical test which is, however, slightly flattened with an attachment scar on the basal side; the test is somewhat longer than originally shown. Paratype C (CPC 422) is an asymmetrical test with a slightly flattened attachment scar on the basal side. Paratype D (CPC 423; PI. 1 14, fig. 2; text-fig. 1e) is an asymmetrical and crooked test; the proloculus has a large notched attachment scar which extends a short distance along the second chamber; the edges of the scar are polished and the scar itself has a fine-grained texture. Paratype E (CPC 424; PI. 1 14, fig. 9; text-fig. 1a) is a quite crooked test with the second chamber hemitubular over most of its length owing to attachment scars on one side; short intervals of the second chamber are tubular with no attachment along these portions; the attachment scar extends only slightly on the proloculus. Remarks. T. devoniana is a good example of the genus even though the attachment scars are only slightly developed in some specimens. Tolypammina is attached at intervals along or throughout, the length of the second chamber, while the proloculus may be free, attached, or partially attached as in the present species. The hemitubular nature of the second chamber is characteristically developed to some degree on the attached portions, but this is not so in Hyperammina, which is entirely free. T. devoniana differs from a similar species, T. bulbosa (Gutschick and Treckman 1959), which occurs in the Upper Devonian and Lower Mississippian of the United States, in having a more irregularly shaped and partially attached proloculus, while the pro- loculus of T. bulbosa is free and quite spherical. Tolypammina helina Crespin 1961 Plate 115, figs. 1-4 1961 Tolypammina helina Crespin (in part), p. 407, pi. 67, figs. 1-4. The holotype (CPC 425; pi. 115, figs. 1, 2) has an attached proloculus and a hemitubular second chamber which is flattened all along the basal attached side and is only partly floored; whorls are all roughly in the same plane; the upper, unattached side is somewhat obscure and was shown slightly incorrectly in the original drawing; the present figure (PI. 1 15, fig. 2) shows the nature of the whorls on this dorsal side. Paratype A (CPC 484; PL 115, figs. 3, 4) is a completely attached test which is flattened on the basal side, with the second chamber hemitubular throughout and only partly floored; the whorls are roughly in the same plane. Remarks. T. helina is remarkably similar to T. gersterensis Conkin and Conkin 1964, which occurs in the Upper Devonian and Lower Mississippian of the United States. A similar species, in part at least, is T. irregularis Blumenstengel 1961, from the Upper Devonian and Lower Carboniferous of Germany; the exact relationships between these species are not certain. Tolypammina nexuosa Crespin 1961 Plate 115, figs. 5-8 1961 Tolypammina nexuosa Crespin, pp. 407, 408, pi. 67, figs. 6-8. 1961 Tolypammina helina Crespin (in part), p. 407, pi. 67, fig. 5. 608 PALAEONTOLOGY, VOLUME 11 The holotype (CPC 486; PI. 115, figs. 7, 8) displays an irregular pattern of twisting and winding. Paratype A (CPC 487) possesses a flat, attached, basal side; no proloculus is visible; the rounded hump on the specimen is not the proloculus, but merely a bend in the hemitubular second chamber. Paratype B (CPC 488) has a hemitubular test which is flat and attached on the basal side; no pro- loculus is present. Paratype B of Tolypammina ‘ helina ’ (CPC 485; PI. 1 15, figs. 5, 6) probably belongs to T. nexuosa for it does not possess a planispiral portion; the basal side is flat and the second chamber is hemi- tubular. Remarks. This species is of doubtful status, for a proloculus should be seen in order to erect a species and to make specific identification in the genus Tolypammina ; the early relationship of the second chamber to the proloculus is usually distinctive in a given species. Tolypammina sp. Plate 1 14, figs. 7, 8 1961 Marsipella sp., Crespin, p. 401, pi. 65, fig. 9. This specimen (CPC 400; PI. 1 14, figs. 7, 8) consists of a broken test with no proloculus preserved; the test is hemitubular and attached, with the attached side open, without a floor; upper unattached side is of coarse and clear quartz grains. Remarks. This specimen clearly belongs to Tolypammina, even though it is a fragment, for it is hemitubular and attached. Tolypammina ? sp. Plate 1 14, figs. 5, 6 1961 Rhabdammina virgata Crespin, pp. 400, 401, pi. 64, figs. 7-10; pi. 65, figs. 1, 2. The holotype (CPC 394; PI. 1 14, figs. 5, 6) consists of a test which is broken at both ends so that the nature of the proloculus and aperture is not known; this fragment is tubular, with no evidence of attachment noted, but the test is only partly preserved; the test is composed of medium to coarse quartz grains. Although the evidence for positive generic assignment is not present in this specimen, we prefer to regard this specimen as belonging to Tolypammina rather than Rhabdammina. Paratype A (CPC 395) is a limonitic cast of an orange colour; the test is broken at both ends and is curved; no proloculus is present. EXPLANATION OF PLATE 117 Figs. 1-11. Oxinoxis ampullacea (Crespin), 1961. 1, Holotype of ‘ Lagenammina' ampidiacea, side view. 2, Upper surface of same specimen. 3, Basal attached side of same specimen. 4, Holotype of ' Colon- ammina imparilis ’ Crespin; oblique view of basal attachment and upper surface showing hemi- tubular apertural neck and partially obscured wedge-shaped attachment scar. 5, Basal attached surface of same specimen showing attachment scar. 6, Upper surface of coarse-grained test, same specimen, showing a few coarser quartz sand grains incorporated in the test. 7, Paratype B of ‘ Colonammina imparilis upper convex surface of test showing holes in the wall, not apertural, due to breakage. 8, Basal attached surface of same specimen showing large attachment scar. 9, Plesiotype of ‘ Proteonina sp.’, Crespin 1961, basal attached and pitted surface of test. 10, Upper pitted surface of test, same specimen. 11, Paratype A of ‘ Lagenammina ’ ampidiacea Crespin 1961, upper surface. All figures approximately x 105. Palaeontology, Vol. 11 PLATE 117 CONKIN and CONKIN, Upper Devonian arenaceous foraminifera CONKIN AND CONKIN: REVISION OF UPPER DEVONIAN FORAMINIFERA 609 Paratype B (CPC 396) is a bent and curving test, broken at both ends, with no proloculus or aperture; tube is composed of coarse-grained quartz. Paratype C (CPC 397) is broken at both ends and is curved; no proloculus is preserved. Paratype D (CPC 398) consists of an irregularly coiled test, broken at both ends; no proloculus is preserved. Paratype E (CPC 399) is broken at both ends with no proloculus preserved; the test is red-orange in colour and is composed of quartz sand grains. Remarks. These specimens appear to be fragments of free portions of Tolypammina, but specific identification is impossible. Acknowledgements. We are grateful to: Dr. Irene Crespin for her kind permission to revise these foraminifera; Dr. D. J. Belford for photographing the specimens; Dr. John Talent, who led us to the fossiliferous outcrops of the Silurian Cowombat Formation in Gippsland and southern New South Wales ; Mr. Edmund Gill, who facilitated examination of type specimens in the Victoria Museum, Melbourne, and Dr. K. S. W. Campbell for guiding us in a study of the Yass-Taemas area, New South Wales. This work was made possible by a Fulbright Senior Research Fellowship to the University of Tasmania and sabbatical leave from the University of Fouisville in 1964-5. REFERENCES blumenstengel, v. h. 1961. Foraminiferen aus dem Thiiringer Oberdevon. Geologic, 10 (3), 316-35, pi. 1-3. chapman, F. 1933. Some Paleozoic fossils from Victoria. Proc. R. Soc. Viet. 45 (2), 245-8, pi. 11. conkin, j. e. 1954. Hyperammina kentuckyensis, n. sp. from the Mississippian of Kentucky, and discussion of Hyperammina and Hyperamminoides. Contr. Cushman Fdn. foramin. Res. 5 (4), 165-9, pi. 31. 1961. Mississippian smaller foraminifera of Kentucky, southern Indiana, northern Tennessee, and south-central Ohio. Bull. Am. Paleont. 43 (196), 131-368, pi. 18-27. and conkin, b. m. 1964. Devonian foraminifera: Part 1, The Fouisiana Fimestone of Missouri and Illinois. Ibid. 47 (213), 53-105, pi. 12-15. 1965 a. Permian foraminifera in Tasmania. Aust. J. Sci. 28, 312. 19656. Ordovician (Richmondian) foraminifera from Oklahoma, Missouri, Illinois, and Kentucky. Okla. Geol. Notes, 25 (8), 207-21, pi. 1, 2. and canis, w. 1968. Mississippian foraminifera of the United States. Part III, The lime- stones of the Chouteau Group in Missouri and Illinois. Micropaleontology, 14, 133-78, pi. 1-4. crespin, i. 1958. Permian foraminifera of Australia. Bull. Bur. Miner. Resour. Geol. Geophys. Aust. 48, 1-207, 33 pi. 1961. Upper Devonian foraminifera from Western Australia. Palaeontology, 3, 397-409, pi. 64-7. loeblich, a. R. andTAPPAN, H. 1964. Sarcodina, chiefly ‘Thecamoebians’ and Foraminiferida. In moore, r. c., ed. Treatise on Invertebrate Paleontology, Part C, Protista l.Geol. Soc. Am. and Unix. Kans. Press. stewart, G. A. and lampe, l. 1947. Foraminifera from the Middle Devonian bone beds of Ohio. J. Paleont. 21, 529-36, pi. 78, 79. summerson, c. h. 1958. Arenaceous foraminifera from the Middle Devonian limestones of Ohio. Ibid. 32, 544-88, pi. 81, 82. JAMES E. CONKIN Department of Geology University of Fouisville Fouisville, Kentucky 40208 U.S.A. BARBARA M. CONKIN Jefferson Community College University of Kentucky Fouisville, Kentucky 40202 U.S.A. Typescript received from authors 27 February 1968 A NEW MEDUSOID (?) FROM THE SILURIAN OF ENGLAND by ISLES STRACHAN Abstract. Duodecimedusina palmeri sp. nov. is described from the Upper Llandovery of the Malverns and West Midlands. This extends the stratigraphic range of the genus previously described from the Carboniferous and Lower Devonian. It is also the first record from Europe. One of the specimens described here was picked up some years ago on a student field excursion to the Lickey Hills, just south of Birmingham. Like so many interesting specimens it was a loose block so that there is no direct evidence of which way up the specimen was, or more important, from which geological horizon it came, both Cam- brian and Lower Silurian being exposed in the quarry. In June 1967, a similar specimen (for which the stratigraphic horizon is more precise) was collected by Mr. K. F. Palmer on a student trip to the Malverns. The confirmation of geological age and morphological features make recording of the specimens more important since the range of the genus is considerably extended in time and space. Genus duodecimedusina King This genus was proposed by King (1955) for three species, D. typica, D. wycherleyi, and D. ulrichi. The first two of these, known from a single specimen each, come from Upper Carboniferous of Kansas, and the third species, also from a single specimen, is from the Lower Devonian of Bolivia. Avnimelech (1966) has recently described a fourth species, D. aegyptica from the Lower Carboniferous of Egypt, again from a single specimen. The present specimens therefore extend the range of the genus down to the Lower Silurian as well as being the first record from Europe. Harrington and Moore (1956) classified Duodecimedusina amongst the ‘Medusae Incertae Sedis’ but Avnimelech has proposed transferring the genus to the Proto- medusae on the grounds of similarity to Brooksella which forms the whole class. The latter genus, however, frequently shows supplementary lobes on the sub-umbrellar surface which are not found in the Egyptian and British specimens. The American specimens are all attached to a ‘pedestal’ so that the sub-umbrellar surface cannot be seen. The general category of ‘Medusae Incertae Sedis’ therefore seems to me to be a better resting place for Duodecimedusina for the present. Duodecimedusina palmeri sp. nov. Material. Holotype, BU 302, from base of Upper Llandovery, Gullet Quarry, Malverns, in a slightly micaceous siltstone, way-up not known. Paratype, BU 303, probably from Rubery Sandstone (Upper Llandovery), Rubery Quarry, Leach Green Lane, Birmingham, in a medium to coarse sandstone. Description. Body roughly circular, 16-19 mm. in diameter. There is a central raised [Palaeontology, Vol. 11, Part 4, 1968, pp. 610-11.] STRACHAN: NEW MEDUSOID (?) FROM THE SILURIAN OF ENGLAND 611 area, rather poorly defined, occupying about half of the diameter, surrounded by twelve lobes. The lobes are distinct at the periphery, measuring about 3 mm. across, but merge centrally into the more raised area. The undersurface is poorly preserved in the holo- type and not seen on the paratype. text-fig. 1. a, b, c. Hclotype, BU 302, exumbrellar, side, and subumbrellar views, d. Paratype, BU 303, exumbrellar view. All X 1|. Discussion. D. palmeri is similar to D. typica King in the poorly delimited area of the exumbrellar surface but is little more than half its size. D. wycherleyi King is even smaller (10-1 1 mm. in diameter) but has a central depression, which is also found in the Lower Devonian, D. ulrichi King. D. aegyptica Avnimelech is similar in size to D. palmeri but is much more convex in lateral view and has more clearly marked off lobes. Since the above was written, Dr. Goldring has sent me another specimen from the Gullet Quarry. This specimen (Reading Univ. Geol. Dept. no. 14601) was collected as a loose specimen by Mr. D. C. Smith while mapping the area. It is somewhat larger (about 28 mm. in diameter) and the lobes show weak, but fairly constant, concentric bands, a feature not reported on any other of the species. REFERENCES avnimelech, M. A. 1966. A new medusoid fossil from the Lower Carboniferous of Egypt. J. Paleont. 40, 742-5. Harrington, H. J. and moore, r. c. 1956. Protomedusae, etc. In moore, r. c. Treatise on invertebrate Paleontology, Part F, Coelenterata. Kansas. king, r. h. 1955. In harrington, h. j. and moore, R. c. Fossil jellyfishes from Kansan Pennsylvanian rocks and elsewhere. Bull. geol. Surv. Kansas. 114, 153-64, pi. 1, 2. ISLES STRACHAN Geology Department University of Birmingham P.O. Box 363 Birmingham, 15 Typescript received from author 8 March 1968 THE ATRYPIDINE BRACHIOPOD DAYIA NAVICULA (J. BE C. SOWERBY) by E. V. TUCKER Abstract. The atrypidine brachiopod Dayia navicula (J. de C. Sowerby) is described and a neotype is designa- ted. The musculature is reinterpreted and shown to differ from previous interpretations. Comparisons are made between Dayia navicula s. str. and other forms referred to the species. Between the Ludlow and the Lower Devonian the shells increase in size and incurvature of the umbo becomes less marked. Prior to 1881 the brachiopod now designated Dayia navicula rested uneasily in palaeontological works first under the style of Terebratula navicula and later Rhyn- chonellal navicula. JVTCoy in 1846 anticipated in part the probable relationship between this form and the atrypids when he referred it to Atrypa navicula. In 1881 Davidson, studying material skilfully prepared by the Revd. N. Glass, recognized the distinctiveness of the brachidium and erected the new genus Dayia. Since then material referred to the species D. navicula has received attention especially in the hands of Kozlowski (1929, p. 179) and Alexander (1947). The latter undertook a morphological study of the characteristic British form but her conclusions contain certain misconstructions. In stratigraphy, Dayia navicula has been widely used as an index of Ludlovian rocks, particularly for the Leintwardine Beds (formerly the Dayia or Mocktree Shales of Elies and Slater (1906)) which underlie the Upper Ludlow or Whitcliffe Beds. Straw (1937), however, noted its presence in rocks ranging from the nilssoni zone into the base of the Upper Ludlow at Builth. This extended range was re-emphasized by Earp (1938) and Shirley (1939 and 1952) for other parts of Wales and the Welsh Borderland. Davidson (1869, p. 191) in the monograph of Silurian brachiopods records Wenlock occurrences given in Murchison’s ‘Siluria’ but only the general locality of Builth is mentioned. A similar stratigraphical range has been suggested in continental Europe. Barrande (1879) records the species from the Upper Ordovician of Bohemia but this claim needs re-examination. In Scandinavia, Lower Ludlow forms, although smaller than the typical Dayia navicula, closely resemble certain British Elton Bed (Lower Ludlow) forms. Koslowski’s description (1929, p. 179) and illustrations of Podolian specimens from the Marnes de Dzwinogrod, now believed by Jaeger (1965) and others to be post-Ludlovian and possibly Gedinnian in age, indicate close agreement with the characteristic British form. The existence of other sub-species, or at least variants of Dayia navicula, is suggested, however, by the work of Boucek (1940) and Shirley (1962). Shirley emphasizes the differences between varieties from the Schistes de Lievin and the Kobbinghauser Schichten of the Artois basin and Westphalia respectively, which have previously been referred to as Dayia navicula, and the characteristic British forms; a view endorsed by the present writer. This misidentification and the usage of Dayia navicula as a Ludlovian index fossil is partly responsible for the discordant views which have been expressed on the correlation of French-Rhenish and British successions. [Palaeontology, Vol. 11, Part 4, 1968, pp. 612-26, pis. 118-21.] TUCKER: ATRYPIDINE BRACHIOPOD DAYIA NAVICULA (J. DE C. SOWERBY) 613 In her account, Alexander referred (1947, p. 305) to a specimen in the Geological Survey collection as holotype. This specimen does not match Sowerby’s illustration in Murchison’s Silurian System (1839, pi. 5, fig. 17) however, differing in size, detail of ornamentation, nature of ventral umbo, and in its geographical location. Search else- where has failed to reveal the type specimens and they are presumed to have been lost. To assist future comparison a neotype for Dayia navicula is designated here, the mor- phology is re-described, the musculature reinterpreted and some aspects of variation in the genus are outlined. The material on which the study is based has been collected by the writer, chiefly from the Leint- wardine Beds of Downton Gorge (Herefordshire) and various Silurian inliers in the Welsh Borderland, particularly those of Woolhope and Usk, and from the Elton Beds and Whitcliffe Beds. These personal collections were supplemented by collections housed in the Geological Survey Museum (GSM) and Mrs. Alexander’s collections of silicified specimens in the Sedgwick Museum, Cambridge (SM). (Manuscript notes of Alexander suggest that this collection was made near Greenway Cross (Nat. Grid. SO/461 827), although repeated visits to the same locality by the present writer have revealed few specimens of Dayia navicula. The fauna indicates a Lower Leintwardine age for the beds.) The source of foreign material is indicated in the appropriate section. SYSTEMATIC DESCRIPTION Superfamily dayiacea Waagen 1883 Family dayiidae Waagen 1883 Subfamily dayiinae Waagen 1883 Genus dayia Davidson 1881 Type species. Terebratula navicula (J. de C. Sowerby) 1839. Diagnosis. Small, strongly inequivalve, pedicle valve larger; hinge line gently curved; shell generally elongated medially but sometimes spherical. Brachial valve with sulcus expanding to one-third of shell width at anterior margin; weakly inflated laterally. Pedicle valve highly arched transversely, smoothly curved medially. Umbo normally incurved. Lateral commissure dorsally curved, anterior commissure sulcate. Surface ornamentation of very fine costae and weak growth-lines, indistinct. Spires directed towards sides of pedicle valve but contained within the first coil. In the pedicle valve a platform of secondary callus supported the diductor muscles. Muscle fields deeply impressed, of chevron form. Adductor scars ill-defined. Prominent median septum in the brachial valve extends to three quarters valve length and separates well- developed adductor muscle fields. A septalium divides a subdued cardinal plate at the centre of the hinge-line. Range. Silurian (Ludlovian, possibly Wenlock) to Devonian (Gedinnian). Dayia has been recorded from many Western European countries along a belt from the Dingle peninsula in south-west Ireland to Podolia in the U.S.S.R., and from Scandinavia. It is not known from extra-European areas. Discussion. The pre-Ludlovian records of Dayia require re-investigation. Dayia cymbula (Davidson) has been recorded from the Caradoc Beds of Hendre Wen (North Wales) and from the Drummuck Group (Ashgillian) near Girvan. Reed (1917) described a variant ‘ girvanensis’ associated with it near Girvan. Externally there are many 614 PALAEONTOLOGY, VOLUME 11 similarities between these forms and Dayia navicula. The musculature of the pedicle valve differs in detail however and the teeth are more prominent. Dental plates, observed by Davidson, are lacking in Dayia. The nature of the brachidium of Dayia cymbula is not known. Dayia navicula (J. de C. Sowerby) Plates 118-21 1839 Terebratula navicula Sowerby, p. 611, pi. v, fig. 17. 1846 Atrypa navicula (Sowerby); M‘Coy, p. 40. 1848 Hypothyris navicula (Sowerby); Phillips, p. 281. 1852 Hemithyris navicula (Sowerby); M‘Coy, p. 204. 1859 Rhynchonella ? navicula (Sowerby); Lindstrom, p. 366. 1869 Rhynchonella ? navicula (Sowerby); Davidson, p. 191. 1881 Dayia navicula (Sowerby); Davidson, p. 291. Neotype. A specimen collected by the writer and deposited in the Geological Survey Museum (GSM 103291) is selected as neotype (PI. 1 18, fig. 2). Sowerby’s type locality is vaguely defined. The neotype is from the Lower Leintwardine Beds of Downton Gorge, Herefordshire (Nat. Grid SO/431 733) approximately 55 ft. above the base. Occurrence. Dayia navicula is present throughout the Leintwardine Beds of the Welsh Borderland, being most common in western areas. It extends upwards into the Lower Whitcliffe Beds, especially in adjacent parts of Wales. Closely allied forms also occur in the Elton Beds of these areas and at similar levels in Scandinavia whilst other variants range upwards into Lower Devonian rocks of central Europe. Dimensions of neotype. Length 10-4 mm., width 8 0 mm., thickness 7-2 mm. Diagnosis. As for the genus. The umbo characteristically is gibbous and rests upon the brachial valve, effectively sealing the foramen. The diductor muscle platform is pro- minent in the pedicle valve and the chevron scars occupy much of the shell width. Description. Considerable shape variation can be recognized. Ontogenetic studies demonstrate a progressive relative increase in elongation with age. (PI. 118, figs. 1-5). This is generally accompanied by a deepening of the sulcus in the brachial valve and accentuation of umbonal incurvature in the pedicle valve. This incurvature ultimately caused pedicle atrophy during maturity (Tucker 1965). In juvenile growth stages the sulcus remains shallow and smoothly curved transversely. It becomes steeper sided in mature forms, where there is an equal division between forms with a deep, flat-bottomed sulcus and those more sharply angular. Variation in mature shells is expressed also by differences in degree of elongation and degree of incurvature of the ventral umbo. Most shells have a pronounced elongation (the width to length ratio exceeding 1:1-2) but EXPLANATION OF PLATE 118 Figs. 1-5. Dayia navicula (J. de C. Sowerby). All un-numbered specimens in author's collection. Lower Leintwardine Beds, Downton Gorge, Herefordshire. National Grid: SO/431 733. Dorsal, anterior, and lateral views of five growth stages. Fig. 2 is designated neotype (GSM 103291) (all x4). Palaeontology, Vol. 11 PLATE 118 TUCKER, Dayia TUCKER: ATRYPIDINE BRACHIOPOD DAYIA NAVICULA (J. DE C. SOWERBY) 615 some retain more squarish outlines. All mature shells have a strongly incurved umbo but this is accompanied by excessive elevation of the posterior part of the ventral valve in some specimens, producing an acute transverse arch which accentuates elongation of the shell. Dentition. The teeth diverge at an angle of less than 160°. Their length is almost one- quarter the shell width. They are most prominent posteromedially, hence articulation is best defined towards the centre-line where the inward and partly downward facing edge of each tooth locks into a laterally opening notch within the sockets (see text-fig. lg). Thickening of the inner socket-ridge beneath, and to a lesser extent above, the socket is accompanied by a swelling into the visceral cavity, intensifying the socket and assisting articulation. The stages of development of the inner socket ridge are demon- strated by the pattern of growth lines shown in text-fig. 3 d. The crenulations on the dorsal surface of the teeth referred to by Alexander (1947, p. 307), are rarely apparent although similar features, lying on the inner face, have been observed on one silicified mould in her collection. There is no clear indication of these features in the micro-structure of other, calcareous, shells. Lophophore supports. These comprise crura and laterally directed spires united by a comparatively simple jugum (see text-fig. 1). The crura arise from the inner margin of, and fuse with, geniculiform inner shell surfaces which flank a septalium in the cardinal plate. Geniculation commences on a level with the hinge axis and immediately adjacent to the shell mid-line. Its development is reflected in the micro-structure of the shell where a close relationship is seen to exist with the development of the septalium, beneath which concentration of growth-lines indicates a retarded rate of shell secretion (see PI. 121, figs. 1,2; text-fig 3). The crura develop from the septalium walls as narrow rods and follow the line of the walls as they diverge at an angle of about 55°. Complete fusion and continuity of shell structure is rapidly achieved between each crus and the parent shoulder, which provides additional support. The crura extend ventrally a distance equal to about one quarter the depth of the cavity, and curve gently with the convex edge facing anteriorly; they descend anterior to their point of origin. At their distal end they produce the primary lamellae which first extend laterally to a position near the dental sockets before descending anteriorly. Peels taken from the crura reveal lamellae of fibrous calcite parallel to the surfaces. There is complete continuity with the primary coil. In longitudinal sections the micro-structure of the supporting crural base reveals a sharp projection within the secondary layer representing an early stage of crural development (see text-fig. 2). The base is supported on a flat spoon-shaped depression within the brachial shell. The primary lamellae descend almost vertically at a position close to the outer socket ridge before producing the first coil. The first four or five coils are strongly flexed along their dorsal arc (text-fig. 1 and PI. 119, fig. 10). Up to nine coils are contained in each spire, fiexuring being less pronounced on each successive coil. The coils are more widely spaced dorsally than ventrally, hence the axis of rotation of successive coils migrates ventrally. Alexander (1947, p. 308) aptly describes each spire as being ‘. . . shaped like a sombrero with one half of the brim turned up’. text-fig. 1. Davia navicula (J. de C. Sowerby), Lower Leintwardine Beds, Downton Gorge, Hereford- shire. Series of transverse sections drawn from cellulose acetate peels. X 5. TUCKER: ATRYPIDINE BRACHIOPOD DAYIA NAVICULA (J. DE C. SOWERBY) 617 The jugum, arising from a position low on the first coil, is V-form, the apex pointing postero-ventrally. It is sharply recurved at its union with the primary coils (text-figs. \p-r). A jugal stem points posteriorly. There appears to be perfect continuity in the micro-structure of jugum and primary lamellae. text-fig. 2. Features of shell micro-structure in Dayia navicula (J. de C. Sowerby). ( a ) longitudinal section through muscle platform of ventral valve to show the position of an earlier valve floor and muscle attachment region. ( b ) longitudinal section adjacent to crus showing crural base formed of callus and ‘embryonic' crus. These are in turn supported by a spoon-shaped depression in the brachial valve, (c) longitudinal section through crus. All x 16. Fimbria extend from the anterior part of the primary coils and are directed anteriorly. They originate in pustule-like growths within the coils which produce surface corruga- tions when fimbria are not fully developed. The pustules generally point medially. Discrete pustules occur rarely on the lower part of the jugum. Silicified specimens (PI. 119, fig. 9) clearly show the fimbria; they attain a maximum length of over 0-175 mm. Median septum. At its anterior end the median septum has the form of a low rounded ridge and is little more than an internal bulge complementary to the external sulcus in the brachial valve. In a posterior direction the septum is more prominent and ridge-like with broad, low, supporting buttresses on its flanks. Towards the hinge line secondary callus builds up over the median septum producing a more triangular form in 618 PALAEONTOLOGY, VOLUME 11 cross-section before the median septum fades into the lower part of the cardinal plate, the transverse profile becoming increasingly subdued. Ontogenetic changes of the median septum are preserved in the micro-structure of the shell (text-fig. 3a, b). Within the secondary layer calcite was secreted about the mid-line, successive growth layers being inclined from the shell outer surface laterally inwards and gently convex outwards. The following stages can be recognized in the development of the median septum from this layer: (i) Simple fold of shell (see above); (ii) Addition of secondary callus producing buttresses which are to serve as supports to the median septum; (iii) Successive addition of calcite layers to produce the characteristic ridge form, there being little lateral addition of calcite at this stage; (iv) Concentration of growth lines over crest of median septum indicating suppression of development ventrally but the addition of calcite on the flanks producing a triangular form; (v) Extensions of process operating in (iv) to markedly increase the width; continua- tion of this process finally flattening out the floor of the brachial valve at the mid-line. The full sequence is discernible within the brachial shell just below the hinge line. Peels show darker areas of more concentrated lamellae at intervals. They increase in number posteriorly and must represent growth pauses. One shell revealed at least five and possibly six such growth pauses. EXPLANATION OF PLATE 119 Figs. 1-7, 9-10. Dayia navicitla (J. de C. Sowerby). 1, Upper Llanbadoc Beds, Darran, Usk, Monmouth- shire. National Grid ST/327 979 (GSM 103292). Internal mould of ventral valve showing chevron diductor field, with the accessory diductor field in its apex, and the adductor muscle field. Between the adductor muscle impressions faint ridges (? vascula media) exist (x7). 2, Lower Leintwardine Beds, Greenway Cross, near Norton, Shropshire. National Grid SO/461 827. (SM A1 1207). Internal mould of ventral valve showing diductor muscle field in an immature specimen (x5). 3-4, Lower Leintwardine Beds, Greenway Cross, near Norton, Shropshire. National Grid SO/461 827. (SM A1 1202 and All 206). Internal moulds of brachial valve showing median septum, adductor muscle fields and position of dental sockets (both x 5). 5, Lower Leintwardine Beds, Greenway Cross, near Norton, Shropshire. National Grid SO/461 827. (SM A1 1655-64 one of twelve specimens bearing number 2000). Lateral view of internal mould of ventral valve showing part of chevron and vascular markings ( ? gonadal sacs) ( x 5). 6, Lower Forest Beds. Llancayo, Usk, Monmouthshire. National Grid SO/372 036. Internal mould of ventral valve showing chevron diductor impression (x6). 7 and 1 1, nilssoni-scanicus zone, S. side of Mynydd-y-Gaer, Denbigh Moors, Denbighshire. National Grid SH/974 715. Internal moulds of ventral valves showing chevron diductor impression. Fig. 7 is an immature specimen also showing the adductor impressions (x4 and x6 respectively). 9-10, Lower Leintwardine Beds, Greenway Cross, near Norton, Shropshire. National Grid SO/461 827. (SM A1 1655-664 (specimen numbered 2042) and All 201 ). Silicified specimens showing form of spiralium. 9, lateral view exhibiting pustules and fimbria on spire, photographed through the silica mould (this specimen has been cut along the left side). 10, looking from the anterior (x7 and x6 respectively). Fig. 8. Dayia sp. Kobbinghauser Beds (Lower Devonian), Huinghausen, W. Germany. Internal mould of posterior part of ventral valve showing relatively small diductor muscle field ( x 4). Palaeontology , Vol. 1 1 PLATE 119 TUCKER, Day la TUCKER: ATRYPIDINE BRACHIOPOD DAYIA NAVICULA (J. DE C. SOWERBY) 619 Further changes can be recognized in the shell micro-structure at the mid-line but these are related to the changes in the cardinal plate. Cardinal plate. Serial sections and latex casts provide no indication of the kidney-shaped cardinal process ascribed to Dayia navicula by Alexander (1947, p. 307). Instead, it is seen to have the form of a flattish or gently arched area lying within the brachial hinge plate, divided by a pronounced cleft or septalium. Suppression of the median septum by lateral addition of callus, as described above, reaches its ultimate conclusion in the text-fig. 3. Micro-structure of shell of Dayia navicula (J. de C. Sowerby) at mid-line of dorsal valve (m.s., median septum; i.s., inner shell surface; o.s., outer shell surface; s., septalium; c., crura; d.s., dental socket), (a) transverse section through median septum x 45 (b) transverse section through posterior part of median septum immediately below hinge line x 50. (c) transverse section through cardinal plate below crura x 50, and (d) including crura X 40. production of the cardinal plate. From then onwards crowding of growth-lines and growth pause at the centre-line reverses the form of the inner shell surface to produce the septalium (text-fig. 3 and PI. 121, fig. 2). This trough is flanked by the crura and is in part a result of the development of the crura and expansion of the inner socket ridge. Musculature. There is clear evidence for the former position of certain muscle fields but the interpretation of these fields and of the particular function of the muscles is less easy to deduce. The pedicle valve contains a prominent chevron muscle-scar impressed on a platform produced by the addition of secondary callus to the inner shell surface. The muscle field occupies much of the shell width in mature specimens but only one-third of the width in juveniles (PI. 119, figs. 1, 2). The apex of the chevron is directed posteriorly and extends to a point level with or even above the hinge-line; however the major part of the muscle field lies below this line. Two muscles occupied this field, one each side of s s C 5934 620 PALAEONTOLOGY, VOLUME 11 the shell mid-line, and they diverged at angles between 90° and 100°. One or more narrow troughs, or sometimes ridges, cross the field parallel to its long axis. Weaker, but more numerous, corrugations cross this line obliquely subparalleling the shell b. text-fig. 4. Muscle fields and postulated form of muscles in Dciyia navicula (J. de C. Sowerby). mid-line. These are muscle tracks. The first set probably result from the anterior exten- sion and migration of the muscle attachment areas during growth. This migration is demonstrated by the micro-structure of the supporting platform where the secondary addition of calcite is seen to obliterate earlier muscle fields which now lie ghost-like TUCKER: ATRYPIDINE BRACHIOPOD DAYIA NAVICULA (J. DE C. SOWERBY) 621 within the layer of callus (see text-fig. 2). The oblique structures are believed to be due to the lateral extension of the muscle areas. Situated in the apex of the chevron is a small elliptical area often divided by a trough (text-fig. 4 and PI. 1 1 9, fig. 1). The limits of the area are strongly defined and the area must have provided an attachment region for separately functioning tissue — probably muscular. Together, the above muscle fields are here interpreted as the diductor muscle attach- ment areas, the main diductors being attached to the chevron field and accessory diductors to the elliptical field. This interpretation differs from that placed upon the muscle-fields by Alexander (1947, p. 308, text-fig. 3). Anterior of the diductor muscle platform in the ventral valve two slightly divergent troughs are weakly impressed into the shell: they are not always discernible. They extend from a position close to the lower apex of the diductor chevron halfway towards the ante- rior margin of the shell. The troughs exist in most growth stages but are most prominent in mature shells (PI. 119, figs. 1, 7). These features were not recorded by Alexander; the present writer interprets them as the ventral muscle-fields of the adductor muscles. In the umbonal region of the ventral valve other corrugations exist on the inner shell surface. Certain of these seem to have been interpreted as diductor muscle fields by Alexander but it is difficult to understand how the muscles could function mechanically if this interpretation is correct. These corrugations are possibly responses to folding of the mantle following pedicle atrophy. Alternatively they may represent pedicle attach- ment areas during early stages of growth, although their size tends to preclude such a simple interpretation. No muscle scars are preserved on the cardinal plate. However, the adductor muscles left prominent impressions on the floor of the brachial valve. Although exhibiting some variation in form, these are normally elongate twin fields, the posterior field embracing half its companion (PI. 1 19, figs. 3, 4). The muscle fields lie partly on the flanks of, and are separated by, the median septum which acts in part as a myophragm. The writer’s interpretation of the muscle fields is shown in text-fig. 4, together with a reconstruction of the form of the muscles as they crossed the visceral cavity. During youth the disposition of the (small) diductor muscle fields in the ventral valve and the cardinal plate in the dorsal valve permitted shell opening probably by uniform contrac- tion of the muscles. In mature forms however the posterior arching of the ventral valve transposed part of the main diductor field to a position posterior of the hinge axis. It was inevitable therefore, that the main functional part of the muscle was that part lying below or anterior of the hinge axis and the main diductor muscles became extended anteriorly, with accompanying anterior extension of the muscle platform. Vascular markings. Faint vascular markings are discernible between the diductor muscle field and hinge extremity in the ventral valve of mature individuals and also anterior of the muscle fields in both valves (see PI. 119, figs. 1, 5). Saccate or lemniscate gonadal sacs can possibly be inferred from fig. 5 whilst fig. 1 shows divergent ridges between the ventral adductor scars which may indicate the former site of the vascula media. Other sinuses have not yet been recognized in Dayia navicula. General micro-structure of the shell. Certain aspects of the micro-structure of the shell have already been discussed and many points of detail can be interpreted from text-figs. 2 and 3 and Plate 121. 622 PALAEONTOLOGY, VOLUME 11 The inner secondary layer of the shell, composed of fibrous calcite, is more prominent than the primary layer. The lamellae lie parallel or subparallel to adjacent shell surfaces, notably the inner surface. Considerable secondary callus has been added at the inner surface particularly around the median septum and on the ventral muscle platform. In the latter instance, as well as assisting change of muscle position it partly accounts for the highly arched and elevated form of the posterior region of the ventral valve. Within the umbo of the ventral valve growth layers are concentrated together as a result of growth pauses. The very thin primary layer is frequently lost. Lamellae are closely spaced and seem to be sub-parallel to the outer shell surface, though sometimes lying obliquely to it. Alexander (1947, p. 309) suggested that the lamellae are ‘. . . directed forward and inward’. In the umbonal region of the ventral valve the continuity of shell structure is inter- rupted by a number of ‘suture lines’, seen in transverse section as distinct seams (sometimes followed by cracks (text-fig. 1 a-b)). They may be represented by inflections in the shell microstructure in longitudinal section (see text-fig. 2). The most prominent suture crosses the shell transversely and descends antero-ventrally through the secondary layer meeting the valve floor at the diductor attachment area. Two weaker sutures diverge ventrally from near the mid-line of the transverse suture; they appear to fuse anteriorly and rapidly fade away. These structures are difficult to interpret but bearing in mind the modifications affecting the muscle platform during ontogeny it is not unreasonable to suggest that they either had a similar supporting function or at least defined the limits of the muscles during some early stage of the individual’s develop- ment. In this respect it is worth noting that a somewhat similar structure, a transverse plate (shoe-lifter process), is exhibited by the Ordovician brachiopod Aulidospira (see Williams 1962, p. 253) another member of the Dayiacea. ALLIED FORMS Dayia is not uncommon in rocks of Eltonian (Lower Ludlow) age in many parts of Wales and the Welsh Borderland. The major difference between these forms and Dayia EXPLANATION OF PLATE 120 Variants of Dayia navicula (J. de C. Sowerby). Figs. 1-3. Dayia navicula cf. bohemica Boucek, upper part of Pridoli Beds (e/L), Bohemia. Dorsal, anterior, and lateral views of three growth stages (all X 4). Fig. 4. Dayia navicula (J. de C. Sowerby); Hemse Group (Lower Ludlow) Lau, Gotland, Sweden. Dorsal, anterior, and lateral views of a mature specimen. (x4). Fig. 5. Dayia navicula (J. de C. Sowerby); Skala Beds, Ruchotina, Podolia. Dorsal, anterior, and lateral views. Ventral umbo imperfect (x4). EXPLANATION OF PLATE 121 Microstructure of shell in Dayia navicula (J. de C. Sowerby). Photographs of cellulose acetate peels of calcareous shell, fig. 1 and 2 slightly retouched. All specimens from Lower Leintwardine Beds, Down- ton Gorge, Herefordshire. National Grid SO/431 733. Fig. 1. Cardinal plate, showing shallow septalium (t.s. x 100). Fig. 2. Growth line pattern associated with socket, crus (proximal end) and septalium (t.s. X 125). Fig. 3. Pustules and fimbria on the lower part of the first coil of the spiralium (l.s. X 100). Palaeontology, Vol. 11 PLATE 120 TUCKER, Dayia Palaeontology, Vol. 11 PLATE 121 TUCKER, Dayia TUCKER: ATRYPIDINE BRACHIOPOD DAYIA NAVICULA (J. DE C. SOWERBY) 623 navicuJa is their smaller size. Mature shells have an average length of between 7 and 8 mm. The degree of incurvature of the ventral umbo is similar although the posterior part of this valve is normally proportionately less elevated above the hinge line. Most specimens occur as internal moulds hence the external morphology is difficult to assess fully but there appear to be no fundamental differences from the typical form. The internal structures are also the same although it has not been possible to interpret the spiralium fully. The intensity, but not always the length, of the chevron muscle scars and platform is similar, and the presence of additional impressions within the chevron supports the interpretation of the musculature. Specimens collected from the Mono- graptus scanicus zone of the Denbigh Moors (North Wales) show chevron muscle scars occupying at least two-thirds of the shell width (PI. 119, fig. 11), a relationship which occurs in the Leintwardine Bed forms. Other specimens collected from the upper part of the Elton Beds (locally, Lower Forest Beds) near Usk possess proportionately smaller muscle fields which occupy approximately one-third of the shell width (PI. 1 19, fig. 6). There is thus minor variation between the typical Dayia navicu/a and these earlier forms and also variation within the latter. It would however be premature to append varietal names until more comprehensive collections from rocks of Eltonian age are available for examination. I am indebted to Dr. Anders Martinsson and Dr. H. Jaeger for providing me with specimens of Dayia from the Hemse Group (considered by Hede 1919 to belong to the zone of Monograptus nitssoni ) of Scandinavia. No attempt at a full description is made here but the following points are worth noting. The average dimensions are similar to those of British Eltonian age but the posterior elevation of the ventral valve is greater than in the Usk form: the umbo is again strongly incurved (PI. 120, fig. 4). The Scandi- navian forms possibly possess closer relationships with the Denbigh Moor forms (which unfortunately have not yielded good externals although the moulds show a prominent umbonal feature). The chevron muscle field shows similar proportions to the Denbigh variety also. One characteristic feature of the Scandinavian forms is the almost horizontal disposition of the teeth; this is extremely rare in British forms. Almost all mature shells possess a deep sulcus and a strongly flexed lateral commissure. Few Podolian specimens of Dayia are available for comparison. Kozlowski’s descrip- tion and illustration of this form in 1929, supplemented by material loaned to the writer by Dr. O. Nikiforova, show the close similarity between it and Dayia navicula. Minor differences in external morphology include a squarer outline to the brachial valve resulting from the straighter hinge and the fact that maximum width is attained closer to the hinge (PI. 120, fig. 5). The shell is relatively less wide than British examples but with a deeper ventral valve, strongly incurved but less elevated at the umbo. The species occurs in the Marnes de Dzwinogrod of the Skala Formation. Similarities exist between certain Podolian and large Bohemian specimens (see below) but the Podolian specimens are closer to the British forms. In 1940, Boucek discussed variation within Dayia navicula in Bohemia. He recognized two subspecies which he named Dayia navicula minor and Dayia navicula bohemica. They are recorded from zones e^x and e/32 of the Budnany Beds which, according to Jaeger (1962 and 1965) range from the Ludlovian into younger beds. It is clear from Boucek’s illustrations that the shape of both varieties is different from the typical British form. Horny (1955, p. 438) suggested that a transition exists between the two subspecies. 624 PALAEONTOLOGY, VOLUME 11 A collection of Bohemian Dayia from the Pridoli Beds (belonging to e/32) have been studied by the writer. They include many growth stages. Based on external morphology L DEVONIAN (PRIDOLY BEDS) BOHEMIA text-fig. 5. Scatter diagrams to show the relationship between incurvature of the ventral umbo and overall size (height + width -(-thickness) in four assemblages of Dayia. Assemblages b and c occur 22 and 55 ft. respectively above the base of the Lower Leintwardine Beds. N, no. of individuals in sample; r, degree of correlation between related regression lines. alone, both Boucek’s subspecies appear to be present. With the exception of one assemblage which may represent the true minor , mature individuals generally attain greater dimensions than the British form. The ventral umbo is less strongly incurved (see PI. 120, figs. 1-3) and the foramen probably remained unimpaired until later in TUCKER: ATRYPIDINE BRACHIOPOD DAYIA NAVICULA (J. DE C. SOWERBY) 625 life; hence pedicle atrophy was delayed. The diductor muscle platform is weakly developed. There is a suggestion of vestigial deltidial plates in some small juvenile shells, although the observed features could result from slight curvature of the ventral interarea at the delthyrium edge. Deltidial plates are not apparent in mature shells. The hinge-line is straighter and proportionately longer than in British forms so producing distinctly squarish cardinal extremities. The shells are less elongate and the lateral margins are more sharply angular. Surface ornamentation is similar but the costae are more prominent. The Kobbinghauser Beds of Germany and their age equivalent the Schistes de Lievin of Artois yield ‘ Dayia navicula'. This form is being studied by Dr. J. Shirley who in 1962 expressed the view \ . that the “Z>. navicula ” of these areas and Lievin differs specifically and constantly from the Middle Ludlow (i.e. Leintwardine Beds ) form and is now to be referred to a new species D. tenuisepta sp. nov. (nomen nudum)’. A specimen from the Kobbinghauser Schichten is figured on Plate 119, fig. 8. It is larger than British forms but possesses relatively small and less prominent muscle attachment areas. STRATIGRAPHICAL CONSIDERATIONS The true nature of the sequence of faunas in the various Siluro/Devonian successions of Great Britain and Continental Europe is now more fully understood than a few decades ago and closer agreement is being reached on correlation. Following Jaeger (1962 and 1965) it is apparent that Dayia- bearing strata in the Rhineland, Artois, Bohemia, and Podolia are younger than the British Ludlovian. Jaeger shows that Monograptus extends upwards at least to the top of the Middle Siegenian in Germany and Bohemia. Teller (1964, table 5) for Poland, also shows graptolites in beds younger than the British Ludlow Bone Bed. Correlation between graptolitic and shelly facies in these areas and adjacent Podolia demonstrates the high stratigraphic level attained by Dayia and emphasizes the inapplicability of the genus as an index of Silurian rocks. The characteristic British forms and the, in part contemporary, Scandinavian forms are ancestral to these other varieties. A study of assemblages separated in time brings out a fundamental change in external morphology during phylogeny. Text-fig. 5 shows the relationship between incurvature of the ventral umbo (see Tucker 1965, fig. 2) and overall size for four assemblages ranging in age from Lower Ludlow to Lower Devonian. Regression lines have been computed for each assemblage. The lower degree of correlation of samples (#) and (c) is due to the low number of individuals and lack of juvenile forms in both samples. How- ever, it can be seen that during phylogeny individuals tended to attain greater dimen- sions whilst incurvature was progressively delayed and less intense. These changes are most pronounced in the assemblage of Dayia navicula bohemica from the Lower Devonian of Bohemia. The reduced rate of hooking would clearly postpone pedicle atrophy, a condition facilitating survival of the individual under a wider range of environment than the quiet water conditions suggested for Dayia navicula sensu stricto by Tucker (1965). Acknowledgements. The author is most grateful to Mr. J. D. D. Smith and Mr. D. E. White, and Mr. A. G. Brighton for making available specimens in their care at the Geological Survey Museum and 626 PALAEONTOLOGY, VOLUME 11 Sedgwick Museum respectively. Mr. White drew the author’s attention to the uncertain status of the assumed holotype. He is also grateful to those others mentioned in the text who supplied him with specimens from abroad. Dr. F. A. Middlemiss kindly read the manuscript. REFERENCES Alexander, f. e. s. 1947. On Dayia navicula (J. de C. Sowerby) and Whitfieldella canalis (J. de C Sowerby) from the English Silurian. Geol. Mag. 84, 304-16. barrande, j. 1879. Systeme silurien du centre de la Boheme. v. Brachiopodes. Prague. boucek, b. 1940. O variability ramenonozcu Dayia navicula (Sow.) a Cyrtia exporrecta (Wahl) a o pouziti metod variacni statistiky v paleontologii. Rozpr. Ceske Akad. Ved Umeni, II, 50, no. 22, 1-27. davidson, T. 1864-84. A monograph of the British fossil Brachiopoda, vols, 3 and 5. Palaeontogr. Soc. \ Monogr .]. - 1881. On the Genera Merista, Suess, 1851, and Dayia Dav. 1881. Geol. Mag. 18, 290-3. earp, J. R. 1938. The higher Silurian rocks of the Kerry District, Montgomeryshire. Q. Jl. geol. Soc. Load. 94, 125-60. elles, g. l. and slater, i. l. 1906. The highest Silurian rocks of the Ludlow District. Ibid. 62, 195-220. hede, j. e. 1919. Om en forekomst of colonusskiffer vid Skarhult i Skane. Geol. For. Stockh. Forh. 41, 113-60. horny, r. 1955. Studie o vrstvach budnanskych v zapadni casti barrandienskeho siluru. Sb. Ustred. Ust. geol. 21, 315-447. jaeger, h. 1962. Das Silur (Gotlandium) in Thiiringen und am Ostrand des Rheinischen Schiefer- gebirges (Kellerwald, Marburg, Giessen). Symposiums-Band 2. Internat. Arbeit stagung iiber die Silur I Devon-Grenze und die Stratigraphie von Silur und Devon, Bonn- Bruxelles 1960. pp. 108-35. Stuttgart. - 1965. Referate. Symposiums-Band der 2. Internationalen Arbeitstagung iiber die Silur/Devon- Grenze und die Stratigraphie von Silur und Devon, Bonn-Bruxelles 1960. Geologie, 14, 348-64. KOZtowsKi, r. 1929. Les brachiopodes gothlandiens de la Podolie polonaise. Palaeont. pol. 1. UNDSTROM, G. 1859. Bidrag till Kannedomen om Gotlands Brachiopoden. K. svenska Vetenskakad. Hand/. 17, 337-82. m‘coy, f. 1846. A synopsis of the Silurian fossils of Ireland. Dublin. 1852. Description of the British Palaeozoic fossils in the Geological Museum of the University of Cambridge. London. Phillips, j. 1848. The Malvern Hills compared with the Palaeozoic Districts of Abberley, etc. Mem. Geol. Surv. G.B. 2, pt. 1. reed, f. r. c. 1917. The Ordovician and Silurian brachiopods of the Girvan district. Trans. R. Soc. Edinb. 51, 795-8. shirley, j. 1939. Note on the occurrence of Dayia navicula (J. de C. Sowerby) in the Lower Ludlow Rocks of Shropshire. Geol. Mag. 76, 360-1. 1952. The Ludlow rocks north of Craven Arms. Proc. Geol. Ass. Loud. 63, 201-6. — — • 1962. Review of the correlation of the supposed Silurian strata of Artois, Westphalia, the Taunus and Polish Podolia. Symposiums-Band 2. Internat. Arbeitstagung iiber die Silur/ Devon-Grenze und die Stratigraphie von Silur und Devon, Bonn-Bruxelles 1960. pp. 234-42. Stuttgart. sowerby, j. de c. 1839. In Murchison, r. The Silurian System. London. straw, s. h. 1937. The higher Ludlovian rocks of the Builth District. Q. JL geol. Soc. Loud. 93, 406-53. teller, l. 1964. Graptolite fauna and stratigraphy of the Ludlovian deposits of the Chelm Borehole, Eastern Poland. Studia geol. pol. 13, 7-88. tucker, e. v. 1965. The ecology of the brachiopod Dayia navicula (J. de C. Sowerby). Ann. Mag. nat. Hist. (13) 7, 339-45. williams, a. 1962. The Barr and Lower Ardmillan Series (Caradoc) of the Girvan District, South-west Ayrshire, with descriptions of the Brachiopoda. Mem. geol. Soc. Lond. 3. e. v. TUCKER Geology Department Queen Mary College London, E. 1 Typescript received from author 2 November 1967 PLANICARDINIA , A NEW SEPTATE DALMANELLID BRACHIOPOD FROM THE LOWER DEVONIAN OF NEW SOUTH WALES by N. M. SAVAGE Abstract. Planicardinia is described as a new genus of septate dalmanellid occurring in deposits of probable early Siegenian age in New South Wales. The new genus, known only by the type species Planicardinia canoli sp. nov., is placed in the Mystrophoridae together with the genera Mystrophora , Kayserella , and Hypsomyonia. It is the earliest known representative of the family and its presence at this horizon indicates that the origins of the group are at least as old as the early lower Devonian. Septate dalmanellid brachiopods have received considerable attention since 1953 when Havlicek described the new Middle Devonian genus Prokopia. In 1955 Cooper re-examined the Middle Devonian forms Kayserella and Mystrophora and proposed the new Middle and Upper Devonian genera Phragmophora, Hypsomyonia , and Mone- lasmina. More recently Johnson (1966) described the new genus Vallomyonia from the Middle Devonian of Nevada and Johnson and Talent (1967a) described the new genus Muriferella from the late Lower Devonian of south-east Australia, Nevada, and Arctic Canada. The new genus described herein is from the early Siegenian Mandagery Park Forma- tion near Manildra, New South Wales, where it occurs infrequently in a large silicified brachiopod fauna (Savage 1967, 1968). It is the earliest known representative of the Mystrophoridae and is therefore of particular interest. Little is known of the phylogeny or origin of this family and the new genus does not directly throw new light on the problem of a possible precursor. However it does show that a highly specialized mystro- phorid existed in the middle Lower Devonian and that the ancestral forms of the family can be expected in even older strata. Attempts to classify septate dalmanellid genera have resulted in several different schemes which disagree markedly. Nevertheless most authors believe the group to be polyphyletic with septate dorsal valves evolving independently in more than a single lineage. Cooper (1955) referred Mystrophora and Kayserella to the Mystrophoridae, Phragmophora and Hypsomyonia to the Onniellidae, and Moneiasmina to the Schizo- phoriidae. Wright (in Moore 1965) proposed that Kayserella , Prokopia , Moneiasmina , and Phragmophora be grouped into a new family, the Kayserellidae, with Hypsomyonia the sole member of another new family, the Hypsomyoniidae, leaving Mystrophora the only remaining member of the Mystrophoridae. Johnson (1966) and Johnson and Talent (1967a) referred the new genera Vallomyonia and Muriferella to the Schizophoriidae. These latter authors discuss possible phylogenetic relationships in some detail (1967a, pp. 44-6) and it is clear that they envisage three distinct groups. The genera Muriferella , Vallomyonia , Moneiasmina , and Hypsomyonia are all placed in the subfamily Draboviinae together with Salopina which is suggested as a possible ancestor, Prokopia and Phrag- mophora are separated as a second lineage, and a third group consists of Mystrophora , Kayserella , and the new genus described herein (see Johnson and Talent 1967a, p. 46). [Palaeontology, Vol. II, Part 4, 1968, pp. 627-32, pi. 122.] 628 PALAEONTOLOGY, VOLUME 11 It seems very probable that Planicardinia is closely related to Mystrophora and to KaysereUci. The three genera have a number of features in common, of which most notable is the possession of a cruralium in the dorsal valve. Cooper (1955, p. 47) does not appear to have been convinced that the small cruralium in Kayserella was the sole attachment surface for the dorsal adductor muscles. Furthermore, Wright (1965, in Moore, p. h338) refers to the cruralium as a septalium and suggests that the adductor muscles were attached to the valve floor. However, Biernat (1959, p. 38) has demon- strated that some Polish specimens of Kayserella possess a cruralium extending half the length of the dorsal valve (Biernat 1959, pi. iii, fig. 7) and it now seems probable that this structure was a functional muscle platform. Hypsomyonia appears to be more closely related to Mystrophora than to Muriferella and Monelasmina. It has a large elevated cruralium which clearly functioned as a muscle attachment surface and a typical mystrophorid pentagonal outline with an emarginate anterior margin. Hypsomyonia has therefore been assigned herein to the Mystrophoridae. It is probable that additional material will eventually be found which will indicate the true relationships of the various septate dalmanellids. Meanwhile any classification is likely to be only tentative. It appears to the author that the septate genera with a cruralium form a distinct phylogenetic group and the classification below differs from that of earlier authors in this respect. SYSTEMATIC DESCRIPTION Phylum BRACHIOPODA Suborder dalmanelloidea Superfamily dalmanellacea Schuchert 1913 Family mystrophoridae Schuchert and Cooper 1931 Revised diagnosis. Small, ventribiconvex, dorsally sulcate Dalmanellacea with a long apsacline or anacline ventral interarea and a shorter anacline or hypercline dorsal interarea. Both the delthyrium and notothyrium are open. The ornamentation is costel- late to subfascicostellate. The dorsal interior has a high median septum supporting a large cruralium, a bilobed cardinal process, and ventrally or ventro-laterally elongate brachiophores. The ventral interior has a short, wide muscle field with the adductor scars not enclosed by the diductor scars. Genera included : Mystrophora Kayser, 1871 Planicardinia gen. nov. Kayserella Hall and Clarke, 1892 Hypsomyonia Cooper, 1955 Genus planicardinia nov. Type species. Planicardinia carroli sp. nov. Diagnosis. A plano-convex mystrophorid with an anacline ventral interarea and a hypercline dorsal interarea. Heavy triangular teeth are supported by strong dental N. M. SAVAGE: PLANICARDINIA , A NEW SEPTATE DALMANELLID 629 lamellae. The dorsal interior has a small bilobed cardinal process, ventrally elongate brachiophores and a vertical spoon-shaped cruralium supported on a variable median septum. Discussion. Known only by the type species, Planicardinia most closely resembles the Western European Middle Devonian genus Mystrophora, particularly in the features of the dorsal interior. Comparison with the type species of Mystrophora (Cooper 1955, pi. 11, figs. 45-9) shows a marked similarity in the form of the ventrally elongate brachiophores with a crescentic cross-section and the high spoon-shaped cruralium. The ventral interior of Planicardinia also closely resembles that of Mystrophora with a similar short thickened muscle field. However, the shell shape of Planicardinia is distinc- tive with the anacline ventral interarea and hypercline dorsal interarea together forming a prominent part of the dorsal surface of the plano-convex shell. Another difference is in the orientation of the cruralium which in Planicardinia is inclined almost normal to the dorsal valve floor, a feature clearly related to the extreme curvature of the ventral umbo which necessitates adductor muscles positioned almost parallel to the com- missural plane (text-fig. 1). The ventral ‘median septum’ in Mystrophora , described and figured by Cooper (1955, pi. 11, fig. 50), has been observed in only a single specimen where it showed as a narrow slit in the internal filling of a calcined specimen. This slit probably represents the position of the high dorsal septum extending to the floor of the ventral valve. Hypsomyonia is easily distinguished from Planicardinia for it possesses a much less curved ventral umbo and an interarea which is apsacline and not anacline. Internally the brachiophores are more laterally directed so that they almost parallel the hinge line and the cruralium is more gently inclined. Kayserella is also very distinct from Planicardinia externally. Like Hypsomyonia it has posteriorly directed interareas and a less rounded outline. Internally the dorsal median septum is longer and the cruralium is much smaller and less upright. In addition, the brachiophores are much shorter. Planicardinia carroli sp. nov. Plate 122 Material. The species occurs very infrequently and only 16 specimens have been recovered from several thousand shells dissolved from a silicified limestone horizon. Of these 2 are complete shells with the valves conjoined, 8 are ventral valves, and 6 are dorsal valves. The specimen numbers used are those of the Palaeontology Collection, Department of Geology and Geophysics, University of Sydney. Specimen SU 19540 is designated the holotype. Other illustrated specimens are paratypes. Description. Exterior. The shell is subcircular to subpentagonal in outline with the greatest width at about midlength. The cardinal margins are very obtuse and the anterior margin is emarginate to slightly rounded. In lateral profile the shell is plano- convex (PI. 122, fig. 25). The ventral valve is strongly convex with the umbo swollen and prominent. The beak is erect or incurved and appears to be commonly resorbed by the pedicle (PI. 122, fig. 15). A long concave interarea is anacline and has an apical angle of about 130°. 630 PALAEONTOLOGY, VOLUME 11 The delthyrium, which has a width of about one-quarter that of the hinge-line, includes an angle of about 35°. The dorsal valve is transversely oval to pentagonal in outline and planar or weakly convex. In the mature dorsal valves there is a distinctive convexity at the valve margin (PI. 122, fig. 10). The umbo is poorly differentiated and the small distinct beak is postero- dorsally directed (PI. 122, fig. 25). A long, slightly concave, hypercline interarea makes a very obtuse angle with the valve surface. (PI. 122, fig. 25). It has an apical angle of about 150°. The notothyrium is open and includes an angle of about 50°. A low broad sulcus in the dorsal valve extends from the beak to the gently sulcate anterior commissure. No distinct fold is present in the ventral valve. The surface is costellate with about thirty-four costellae 2 mm. from the dorsal beak. The costellae increase chiefly by intercalation. Interior of ventral valve. In the ventral interior a broad, deep delthyrial cavity is bounded by short vertical dental lamellae which diverge anteriorly at about 40° (PI. 122, fig. 26). The anterior edges of the lamellae recede slightly and meet the strongly curved valve floor at about one-sixth the distance from the beak. Deep lateral umbonal cavities are present. The teeth are strong and pointed with a triangular profile (PI. 122, fig. 15). A short sub-triangular muscle field is slightly elevated on the thickened floor of the delthyrial cavity but individual scars have not been observed (PI. 122, figs. 14, 26). Interior of dorsal valve. The dorsal interior has a small bilobed cardinal process set deeply in the apex of the notothyrium. The brachiophores are ventrally elongate and blade-like, diverging from the valve floor at about 50° and continuous with the margins of the notothyrium for part of their length (pi. 122, figs. 7, 10). They are crescentic in cross-section and bluntly pointed. The sockets are deeply excavated into the posterior of the brachiophore bases just ventral to the interarea margin (PI. 122, figs. 10, 21). EXPLANATION OF PLATE 122 Planicardinia carrolli sp. nov. Figs. 1-5. Ventral, dorsal, anterior, posterior, and lateral views of complete specimen SU 22665. Note the dorsally directed interareas and the strongly convex ventral valve (fig. 5) (x 6-5). Figs. 6-11. Ventral, dorsal, postero-dorsal, antero-ventral, lateral, and postero-ventral views of the isolated mature dorsal valve SU 19541 showing the very high brachiophores (fig. 10), large spoon- shaped cruralium, and the small bilobed cardinal process (fig. 9) (x 6-5). Figs. 12-14. Antero-dorsal, ventral, and dorsal views of broken ventral valve SU 22666. Note the deep, lateral, umbonal cavities and the elevated muscle-field ( X 6-5). Figs. 15-17. Dorsal, ventral, and lateral views of ventral valve SU 19542 showing the resorbed beak, strong triangular teeth (fig. 15), and external ornament (x6-5). Figs. 18-21. Ventral, dorsal, antero-ventral, and lateral views of mature dorsal valve SU 22667. Fig. 18 shows the spoon-shaped form of the cruralium and the degree to which the median septum has been almost totally resorbed (x 6-5). Figs. 22-5. Posterior, dorsal, anterior, and lateral views of complete specimen SU 19540 (holotype) showing the large dorsally directed interareas (fig. 23), the strongly convex ventral valve (fig. 25), and the gently sulcate anterior commissure (x 6-5). Fig. 26. Antero-dorsal view of broken ventral valve SU 19543. Note the strong dental lamellae and the thickened muscle area ( x 6-5). Figs. 27-32. Ventral, dorsal, postero-ventral, antero-ventral, lateral, and posterior views of the young dorsal valve SU 19544 showing the relatively larger median septum extending into the cruralium (fig. 27), and sockets excavated into the bases of the brachiophores (fig. 29) (x 8). Palaeontology , Vol. 11 PLATE 122 SAVAGE, Planicardinia N. M. SAVAGE: PLANICARDINIA, A NEW SEPTATE DALMANELL1D 631 A high, almost vertical cruralium arises close to the brachiophore bases and extends anteriorly to about half the valve length. The spoon-shaped cruralium is supported by a high median septum which extends a little into the cruralium cavity and in young forms has a maximum elevation near the anterior valve margin (PI. 122, figs. 1 8, 27, 3 1 ). The septum is much shorter in mature specimens (PI. 122, figs. 8, 18). Adductor muscles were presumably at- tached to the inner surface of the crura- lium but distinct scars have not been observed in the few specimens available. The interior of both valves usually reflects the external costellation. Measurements (in mm.) SU 19540 Complete shell Length 4-2 Width 4-3 Thickness 21 SU 19541 Dorsal valve 3-6 61 — SU 19542 Ventral valve 4-6 51 (est.) — SU 19544 Dorsal valve 1-5 2-6 — text-fig. 1. Planicardinia carroli gen. et sp. nov. Sketch to illustrate the position of the cruralium, brachiophores, and dental lamellae within a com- plete shell and the probable position of the ad- ductor and diductor muscles. Ontogeny. The very few specimens of this species include a young dorsal valve and two mature dorsal valves. In the mature specimens the median septum appears to be partly resorbed both anterior and posterior of the almost vertical cruralium, but in the young specimen it extends anteriorly almost to the valve margin and posteriorly well into the cruralium cavity (PI. 122, figs. 27, 29). Another feature of the mature specimens is the extreme ventral elongation of the brachiophores and cruralium which greatly exceeds that in the small form (PI. 122, figs. 10, 21). The large spoon-shaped cruralium seems to be deeply resorbed in one of the mature specimens by the two principal trunks of the lemniscate vascu/a media (PI. 122, figs. 8, 11). Occurrence. The species occurs in the silicified limestone horizon at the base of the Mandagery Park Formation 3 miles south of the town of Manildra, New South Wales, at Locality 1 of Savage (1968). Acknowledgements. The author would like to express his gratitude for the assistance of Professor C. E. Marshall and Dr. G. H. Packham of the Department of Geology and Geophysics, University of Sydney where this work was commenced during the tenure of a Teaching Fellowship; also to Professor F. H. T. Rhodes for the use of the facilities in the Department of Geology, University College of Swansea. It is also a pleasure to acknowledge the helpful comments of Dr. J. A. Talent of the Geological Survey of Victoria, and Drs. A. J. Boucot and J. G. Johnson of the California Institute of Technology. REFERENCES biernat, g. 1959. Middle Devonian Orthoidea of the Holy Cross Mountains and their ontogeny. Palaeont. polon. 10, 1-78, pi. 1-12. cooper, g. a. 1955. New Genera of Middle Paleozoic brachiopods. J. Paleont. 29, 45-63, pi. 1 1-14. 632 PALAEONTOLOGY, VOLUME 11 havucek, v. 1953. O nekolika novych ramenonozcich ceskeho a moravskeho stredniho devonu. Vest, ustred. Ust. geol. 28, 4-9, pi. 1-2. Johnson, j. g. 1966. Middle Devonian brachiopods from the Roberts Mountains, central Nevada. Palaeontology, 9, 152-81, pi. 23-7. and talent, j. a. 1967«. Muriferella, a new genus of Lower Devonian septate dalmanellid. Proc. R. Soc. Viet. 80, 43-50, pi. 9, 10. 19676. Cortezorthinae, a new subfamily of Siluro-Devonian dalmanellid brachiopods. Palaeontology, 10, 142-70, pi. 19-22. kayser, e. 1871. Die Brachiopoden des Mittel- und Ober-Devon der Eifel. Z. dt. geol. Ges. 23, 491-647 pi. 9-14. savage, n. m. 1967. Studies in the Silurian and Devonian of the Manildra District, New South Wales. Unpublished Ph.D. thesis, Univ. Sydney. 1968. The Geology of the Manildra District, New South Wales. J. Proc. R. Soc. N.S. W. (in press). schuchert, c and cooper, g. a. 1932. Brachiopod genera of the suborders Orthoidea and Penta- meroidea. Mem. Peabody Mus. Yale, 4, pt. 1, 1-270, pi. 1-29. wright, a. d. 1965. Superfamily Enteletacea. In moore, r. c. (ed.). Treatise on invertebrate paleon- tology, Part H, Brachiopoda, EI328-H346, Lawrence, Kansas. n. m. savage Department of Geology University of Natal Durban, Natal Typescript received from author 18 March 1968 S. Africa PROBABLE DISPERSED SPORES OF CRETACEOUS EQUISETITES by D. J. BATTEN Abstract. Dispersed spores that are considered to be those of Cretaceous Equiselites have been recovered from Wadhurst Clay ( ?VaIanginian) sediments of southern England. They are described as Pilasporites allenii sp. nov. In the course of a study of the palynology of British Wealden delta sediments, the plant microfossil content of Equisetites soil-beds (in which rootlets, rhizomes, and stems of Equisetites, in various combinations, are preserved in situ) and partings with aerial debris of Equisetites (Allen 1941, 1947, 1959, 1967) was examined in the hope that the spores of these plants, which have not previously been recognized in Wealden sediments, might be recovered. Samples used in this study came from Wadhurst Clay horizons in the Cuckfield No. 1 Borehole (Stubblefield 1967) and exposures at the Railway Brickyard, Sharpthorne, and in the lower clay pit at Freshfield Lane Brickworks, Danehill (text-figs. 1, 2). The soil-bed at Freshfield Lane Brickworks was recorded by Gallois (Stubblefield 1963, p. 37) as being ‘the High Brooms E. lyelli soil-bed, some 25 ft. below the top of the Wadhurst Clay’. Fourteen soil-bed and fragment parting samples yielded a characteristic palyno- logical assemblage dominated (> 29-5%, based on a count of 200 specimens) by alete spores described herein from one of these assemblages as Pilasporites allenii sp. nov., and referred to elsewhere as cfA. P. allenii if not from the type assemblage, following the procedure recommended by Hughes and Moody-Stuart (19676). Seventeen adjacent samples yielded assemblages in which these spores are frequent (>4-5-<30%). Although recently some authors have been using the names of living genera for Mesozoic plants, I consider that the convention of using a different name for Mesozoic plants is a useful one and should be retained. Thus, the name Equisetites has been used here rather than Equisetum. The apparent absence from the palynological record of spores of Equisetites , the com- mon Wealden macrofossil, has in the past been variously attributed to lack of preservation, recognition, or production. The restriction of assemblages in which the inaperturate spores mentioned above are dominant, to Equisetites soil-beds and fragment partings, is con- sidered to be evidence in favour of an association of these spores with Equisetites. Of the 350 preparations I have made from Wealden samples, and of the many preparations in the Sedgwick Museum palynology collection, only those prepared from samples associated with soil-beds and fragment partings have yielded these spores in abundance. They have rarely been recorded from other palynological facies. The relationship was tested after the association had been recognized from the Cuckfield and Sharpthorne assem- blages. I decided to prepare a further series of samples collected from Danehill where [Palaeontology, Vol. 11, Part 4, 1968, pp. 633-42, pi. 123.] A B Cuckfield Nol Borehoie Sidnye Farm (TQ 296 273) Sussex rootlet beds ironstone around roots shell, ostracod and planty laminae(mainly Equisetites remains) grey(N5) rhizomes, roots, rootlets rootlets ostracods ostracods SCALE IN FEET Silt,medium and fine 3 — (laminated) SiltyClay Clay 2 Plant Fragments Sample number prefix : CUC x Not. Grid Ref. Shells SOIL BED E_q uisetites (lyelli ) , s tern s rhizomes,root(let)s o-J text-fig. 1. a and b, Wadhurst Clay sediment cores from Cuckfield No. 1 Borehole. Details of litho- logies (diagrammatic, obtained in part from the Geological Survey log) and horizons from which samples have been prepared for microscopical examination. The colour classification (see also text-fig. 2) is from the Rock-Color Chart (1963) published by the Geological Society of America. Colours have generally been noted when there is a departure from the predominantly grey (N3-N7) colours of the Wadhurst sediments. For the sedimentary rock nomenclature, a modified Wentworth grade scale has been followed. In beds where rhizomes and stems have not been seen but rootlets are recorded, it is likely that the latter are those of Equisetites. The medium grey clays (sometimes shales) are generally interlaminated with paler, generally very thin ( < 1 mm.) siltstones. The presence of shells and ostracods has only been noted when they are abundant. BATTEN: PROBABLE DISPERSED SPORES OF CRETACEOUS EQUISETITES 635 soil-beds and fragment partings had been observed in the field. The first samples proces- sed were the fragment parting samples (DB 367, 366, 365, and 358) and I predicted that Railway Brickyard SHARPTHORNEfTQ 375330) Sussex Freshfield Lane Brickworks DANEHILL(TQ382 266) rootlets, especially concentrated in medium grey clay partings(in situ?) 369- medium grey clay(N5) abundant plant remains (Equiseti tes) scattered ostracods moderate brown (5 YR 4/4) Equisetites stems, rhizomes,rootlets tbin plant fragments partings ostracods == Equisetites fragments greyish yellow(5Y 8/4) and lightgrey (N 7) ?stems, rhizomes, root(let)s of Equisetites sp. SCALE IN FEET 3— | 2- 1- light brownish grey (5YR 6/l)rootlets o-J rootlets,? rhizomes ostracods,plant debris medium dark grey (t rootlets rhizomesof E slightly calcareous medium lightgrey (N6) Equisetites stems(?), rhizomes rootlets rootlets slight ly calca reous stems (?), rhizomes, rootlets dark grey(N3)slightly calcareous Equisetites lyelli stems, rhizomes, rootlets scattered ostracods medium grey (N5) Sample number prefix : DB Sample number prefix:DB text-fig. 2. a, Part of the (lower) Wadhurst Clay succession exposed at the Railway Brickyard, Sharpthorne, Sussex, b, (upper) The Wadhurst Clay succession exposed at Freshfield Lane Brickworks, Danehill. Details of lithologies (diagrammatic) and of horizons from which samples have been pre- pared for microscopical investigation. For ‘Key’ and other comments, see text-fig. 1. they would yield assemblages in which cfA. P. allenii is dominant (> 29-5%). The prediction proved to be correct. Several of the soil-bed samples were subsequently found to contain an abundance of these spores (see Table 1 and text-fig. 2b). Tt C 5934 A Cuckfield BH SAMPLE No (prefix CUC) 789 790 791 791/11 792 792/1 792/2 792/3 792/6 793 794 794/6 796 797 cf A. P. allenii 0 8 3 31 58 25 22 60 5 X 3 X 11 X Other inaperturates 7 14 10 10 X 10 12 8 11 11 7 12 13 10 bisaccates 54 32 34 13 9 27 20 5 28 26 29 21 30 43 Classopoll is 4 8 X 2 X 2 2 X 3 2 4 X 5 2 total ‘fern1 spores 28 45 46 31 23 31 35 20 45 55 51 60 25 24 Cicatricosisporites X 15 14 4 5 7 9 6 12 18 17 16 X 2 B 'Cuckfield BH SAMPLE No (prefix CUC) 863 864 865/3 866 867 868 869 cf A. P. allenii 2 28 60 67 5 X X Other inaperturates 7 8 7 10 8 7 8 bisaccates 49 15 12 8 29 28 41 Classopol 1 is 10 2 3 3 X 3 11 total 'fern' spores 23 38 14 9 54 54 25 Cicatricosisporites 2 4 2 X 8 11 5 C Sharpthorne SAMPLE No (prefix DB) 332 331 330 347 329 349 328 327 325 324 321 320 319 318 317 316 cfA. P. al 1 enii 56 54 9 6 5 30 X X 2 2 Other inaperturates 2 4 4 9 6 6 X 7 10 12 16 10 14 X 5 12 bisaccates 25 30 16 16 27 21 43 20 41 40 25 49 24 35 40 26 Classopollis 40 35 5 X 32 3 25 18 16 10 18 20 30 10 25 40 total 'fern' spores 26 22 11 16 19 40 18 22 12 25 31 18 16 16 25 14 Cicatricosisporites 10 6 X 4 X 15 2 X X X 2 X X X X X D Danehill SAMPLE No (prefix DB) 369 368 367 366 365 364 363 362 361 360 359 358 357 355 356 cf A . P. al lenii 11 13 35 38 64 6 6 2 6 36 46 60 6 5 1 Other inaperturates 11 20 17 10 13 14 11 12 8 6 n 12 8 7 7 bisaccates 38 34 20 24 6 41 28 24 36 24 24 8 24 30 36 Classopoll is 3 4 3 7 X 4 X 4 X 2 1 2 total 'fern' spores 33 25 19 18 8 26 49 49 39 28 14 16 54 54 48 Cicatricosisporites 5 4 3 X X 8 10 11 8 8 2 X 10 12 10 table 1 . Distribution of cfA. Pilasporites allenii and major spore and pollen groups in some of the samples studied (see text-figs. 1 and 2) recorded as percentages, based on counts of 200 specimens. A and b, Cuckfield No. 1 Borehole, c, Railway Brickyard, Sharpthorne. D, Freshfield Lane Brickworks, Danehill. X indicates presence, but 1% or less. The absolute bisaccate and fern spore counts are fairly constant (see Hughes and Moody-Stuart 1967a) and are only apparently reduced in numbers when cfA. P. allenii or Classopollis are abundant due to their non-dispersal. The sample numbers are as they appear in text-figs. 1 and 2, from the highest point to the lowest point in the successions and not in numerical order. BATTEN: PROBABLE DISPERSED SPORES OF CRETACEOUS EQUISETITES 637 DISCUSSION OF THE OBSERVATIONS Although spores cfA. P. allenii are usually frequent in the soil-beds, some horizons lack an abundance of these spores (Table 1). This may be for several reasons. The spore species in most of the soil-bed assemblages are poorly preserved. The poor preservation is partially related to sediment type which probably in turn partly reflects the environment of deposition. Allen (1947) stated that the Wealden Equisetites plants probably grew in shallow waters, colonizing a specific locality for a short period of time, perhaps for as little as three or four years. If, like living Equisetum, they grew predominantly in aerated waters, many of their spores may thus have been deposited in an environment, at times unfavourable for preservation, in which they were largely destroyed. When and where conditions were more favourable for preservation, indicated by the generally better state of preservation of most of the spore species, cfA. P. allenii is found preserved in abundance. Winnowing and subsequent dispersal of these spores in the waters of the Wealden basin could account for a lack of abundance in some horizons, and so could a low rate of spore production, or seasonal production and/or a rapid sediment accumula- tion rate in the soil-bed localities. Since the Equisetites communities were probably relatively local and only existed for short periods of time, even if the plants dispersed large quantities of spores frequently, the total number produced would be small when compared with the number produced by the plants living on the near-by land. Thus if Equisetites spores were dispersed over a large part of the Wealden basin, the chances of obtaining an assemblage in which they constitute more than 5% of the total' number of spores counted is slight. If the plants did not produce large numbers of spores, as is possible, the likelihood is even less. They have not been observed to be more susceptible to corrosion than other dispersed spores, and in fact, when abundant, they are generally better preserved than the majority of bisaccates and some of the ‘fern’ spores in the assemblage. The poor preservation of the latter forms may however be due to reworking. Some movement in the water (Allen 1941, p. 371), perhaps the result of storm wave action, at times caused the destruction of the aerial stems of the Equisetites plants and subsequent deposition of the fragment beds and partings. The fragments are preserved as flattened carbonaceous remains as opposed to stems preserved ‘in situ’ in which the pith cavities have generally been filled with sediment forming pith casts. The cones (strobili) of living species of Equisetum rapidly fragment after spore dispersal. If Equisetites cones similarly decomposed, it is unlikely that a mature cone would have been preserved in Wealden sediments. Only when a cone was prevented from reaching maturity, for example if the supporting stem was broken or the entire plant became suddenly inundated with sediments, would preservation be possible. Although a strobi- lus has yet to be recognized, cfA. P. allenii spore masses are frequently found, especially in assemblages extracted from fragment partings, and are probably microscopic remains of strobili (sporangia) broken up prior to or at the time of deposition. Elaters which characterize the spores of living Equisetum have not been seen surround- ing cfA. P. allenii or separated from them in any of the assemblages. They may have been destroyed during the preparation procedure. However, unless their chemical make-up is such that they are readily destroyed by the chemicals used during preparation, or were removed from the spores at the time of deposition and rarely fossilized, they should 638 PALAEONTOLOGY, VOLUME 11 be present in the palynological preparations. It is possible that the Wealden Equisetites species concerned produced spores which lacked elaters. Apart from some Eocene Equisetum spores (Chandler 1964), and Elciterites triferens Wilson 1943, a Pennsylvanian tri-radiate spore, no other fossil spore species recorded to date possess elaters. They were not found surrounding the spores described from Equisetites compressions by Halle (1908), Hartung (1933), or Gould (1968). Specimens recorded were all those present along selected traverses of strew slides. Stage co-ordinates refer to Leitz Laborlux (LI) microscope, number 557187, Depart- ment of Geology, Cambridge University. The slides have been deposited in the Sedgwick Museum palynology collection. RECORD Anteturma sporites H. Potonie 1893 Turma aletes Ibrahim 1933 Subturma azonaletes (Luber 1935) Potonie and Kremp 1954 Genus pilasporites Balme and Hennelly 1956 Type species. Pilasporites calculus Balme and Hennelly 1956, p. 64, pi. 3, figs. 60-4. Remarks. Several genera have been erected for inaperturate miospores but their diag- noses are not distinct. I consider that it is not practical or helpful to attempt to distin- guish this new spore from all published inaperturates of all ages; it is however, desirable and necessary to group it with and distinguish it from other Mesozoic inaperturates. Pilasporites , the late Palaeozoic (Permian) genus has therefore been selected instead of a Tertiary-based name. The alternative solution of erecting a new genus has been rejected because the primary purpose of this paper is to present an accurate account of the new species, rather than to classify it. Pilasporites allenii sp. nov. Plate 123, figs. 1-4, 6, 10-14 Type sample. CUC 792, Cuckfield No. 1 Borehole, Sidnye Farm, Surrey, England (TQ 296 273) from depth 792 ft. Wadhurst Clay, ?Valanginian. Medium dark grey (N4) clay, size of coarse fraction 25 /a; explanation of plate 123 Figs. 1-4, 6, 10-14. Pilasporites allenii sp. nov.; spores from rock sample CUC 792, preparation T 210, X 1,000. 1, Loose ?ektexine, dark area; T 210/10, LI 33-3 124-8. 2, Smooth form; T 210/1, LI 32-0 118 0. 3, Scabrate exine, development of major fold; T 210/5, LI 49-7 120-1. 4, Scabrate exine; T 210/1, LI 34-1 116-1. 6, More or less smooth exine, scabrate perine; T 210/1, LI 42-0 1 16-2. 10, Smooth exine, smooth crinkled perine;T 210/1, LI 34-7 117-1. 1 1 , Smooth exine, closely adhering perine with granules and small verrucate elements attached to the perine; T 210/10, LI 29-0 123-4. 12, Dark area, scabrate, secondary folds developed around dark area; T 210/14, LI 55-0 126-8. 13, Split form; T 210/2, LI 30-8 123-3. 14, Split form; T 210/1 43-5 116-0. Figs. 5, 7, 8, 9, 15, cf A. Pilasporites allenii sp. nov.; spores from rock sample CUC 792/3, preparation T 244, x 1,000. 5, Showing dark area, only faintly visible; T 244/4, LI 27-7 118-1. 7, Major fold, loosened ?ektexine; T 244/4, LI 51-6 125-2. 8, Scabrate exine, major fold; T 244/4, LI 58-0 116-0. 9, Smooth exine, dark area, closely adhering smooth, crinkled perine; T 244/4, LI 28-1 116-0. 15, Scabrate exine, loose ?ektexine; T 244/4, LI 31-9 121-0. Palaeontology, Vol. 11 PLATE 123 BATTEN, Lower Cretaceous dispersed spores BATTEN: PROBABLE DISPERSED SPORES OF CRETACEOUS EQU ISETITES 639 thin fine silt laminations; Equisetites ‘in situ’ and fragments of the same. Preparation: oxidation 10 minutes Schulze solution, 3 minutes 5% ammonium hydroxide, mineral separation using zinc bromide; short centrifuging; strew slides, Clearcol and Euparal. Palynological facies: Pilasporites allenii well preserved, the bisaccates and some of the ‘fern’ spores less well preserved; 58% P. allenii, 23% total ‘fern’ spores of which 7% are Ischyosporites and 5% Cicatricosisporites, 9% bisaccates (based on a count of 200 specimens). Diagnosis. Miospores, alete, mean diameter 36-7 p, standard deviation 5-36 p (200 specimens), amb circular. Exine 1-25-1 -75 p thick, smooth, or scabrate. Frequently only one (major) fold present, usually developed close to and sub-parallel to the margin of the spore. Spores sometimes split (7% of the specimens), may be surrounded by a crinkled and folded perine c. 0-25 p thick. Perine smooth or scabrate. An indistinct darker coloured area is seen on 4-5% of the specimens, averaging 14/x in diameter. Holotype. Slide preparation T 210/1, LI 42-0 1 16-2; Plate 123, fig. 6. Description. Maximum diameter, observed limits 28-48 p\ coefficient of variation 14-6%. 94-5% of specimens fall between 31 and 44 p (text-fig. 3). Shape may be distorted by folding (PI. 123, fig. 7), 2% of spores unfolded. Exine thickness not always readily measured owing to difficulty in obtaining an optical section. The outer layer of the exine (?ektexine) which carries the sculpture, may be loosened (brought about by oxidation) and simulate a closely adhering perine. No elaters seen. The darker area may result from a thickening of the exine. Usually these spores show folds developed only around and not across the area (PI. 123, fig. 12). Preservation and compression. Preservation good. Spores occasionally unfolded but folds usually developed as mentioned above, probably resulting from the dehydration of the spore leading to the collapse of the spore wall in the subsequent compression. Spores very compressed. Probable affinity. Equisetites lyelli (and ? other species of Cretaceous Equisetites). Similarity of recent Equisetum spores. Observed acetolysed spores of 11 of some 25 living species of Equisetum (E. hiemale , E.fluviatile , E. pratense, E. moorei, E. sylvaticum , E. arvense, E. te/mateia, E. variegatum, E. scirpoides, E. palustre , E. ramosissimum) lack a tri-radiate mark, and frequently lack elaters which separate from the spores during acetolysis. They are apparently smooth or scabrate spherical grains mainly 30-50 p in diameter, surrounded by a sometimes scabrate folded perine. Specimens of some species show a tendency to split. Histograms of the diameters of spores of an individual species are unimodal, with a small standard deviation. Hauke (1963) states that in any one strobilus the size range of the spores is small, but strobili of different specimens of the same species may vary considerably perhaps representing differences in the age of the cones from which they come. Abortive spores also increase the size range obtained from an individual strobilus. Acetolysed specimens are frequently folded. Often only one major fold present (prep. TR002, Equisetum telmateia). Distinction. Local: Couper (1958) erected the genus Perinopo/lenites for monoporate perinaceous grains, the presence of a pore being a diagnostic character although not 640 PALAEONTOLOGY, VOLUME 11 necessarily seen on all grains. The presence of the pore and a consistently firmly attached perine distinguishes this genus and its contained species from Pilasporites allenii. Colamospora mesozoica Couper 1958, a spore that is rarely definitely recognized in Mesozoic sediments, is distinct in possessing, by definition, a small tri-radiate mark similar to the dispersed spores of Palaeozoic Colamospora. SpheripoJlenites scabratus Couper 1958, differs from P. allenii in having a poorly defined pore, an internally scabrate exine, and in being smaller. ExesipoIIenites tumulus Balme 1957 possesses a circular depression surrounded by an area of exinal thickening. 2 1 MEAN I 1 2SD text-fig. 3. Frequency distribution histogram of maximum diameters in microns of Pilasporites allenii sp. nov. Preparation T 210, 200 specimens, mean 36-7 p, standard deviation 5-36 p, coefficient of variation 14 6%. Literature: Satisfactory comparisons are frequently difficult or impossible due to the inadequate descriptions and illustrations of the types. Laevigatasporites maximus Delcourt and Sprumont 1955, and the Tertiary species L. intrapunctatus Kedves 1961, differ from P. allenii in being much larger. L. reissingeri , a Tertiary species which Kedves (1961) refers to the ‘Equisetaceae cf. Equisetum' , is larger and has a thicker ‘exospore’. I. dubius Pflug and Thomson 1953 has a thinner exine than P. allenii ; /. limbatus Balme 1957, is larger, has a thinner exine, and possesses a dark peripheral band; Arctu- cariacites australis Cookson 1947 is larger and has a different sculpture; Pilasporites calculus (Balme and Hennelly 1956, p. 64) has a differentially thickened exine; Pila- sporites phtrigens Balme and Hennelly 1956 is smaller; Aulisporites (Leschik) Klaus I960, possesses an occasionally observable, faintly developed tri-radiate mark. BATTEN: PROBABLE DISPERSED SPORES OF CRETACEOUS EQUISETITES 641 Comparison. Local: Similar spores from other samples or localities are attributed as suggested by Hughes and Moody-Stuart (19676, p. 353). cfA. P. allenii sp. nov. 1. Cuckfield No 1 Borehole, (TQ 296 273), depth 792 ft. 3 in., Wadhurst Clay, general size of coarse fraction of sediment 50 p, Equisetites sp. soil-bed, Equisetites fragments; sample CUC 792/3, prep. T 244/4, preservation good, Pilasporites dominant. Maximum diameter 28 (35-5) 45 p (200 specimens). 2. Cuckfield No 1 Borehole, depth 866 ft., Wadhurst Clay, general size of coarse fraction 8 p, Equisetites sp. fragment parting; sample CUC 866, prep. T 269/2, preserva- tion good, Pilasporites dominant. Maximum diameter 28 (36-8) 47 p (50 specimens). 3. Sharpthorne, Sussex (TQ 375 330), (lower) Wadhurst Clay, general size of coarse fraction 50 p, Equisetites sp. fragment parting; sample DB 330, prep. T 189/1, preserva- tion moderately good, some pyrite and other corrosion, Pilasporites dominant. Maximum diameter 27 (35-8) 46 p (75 specimens). Literature: Reissinger (1950), Couper (1953), and Rogalska (1954) described and/or illustrated spores from Jurassic sediments which resemble P. allenii , and which they referred to the Equisetaceae ( lEquisetum ). P. allenii is also similar to Pilasporites marcidus Balme 1957, an Australian Cretaceous species, and to Equisetosporites hallei , the species that Nilsson (1958) erected for dispersed spores which he thought were similar if not identical to those produced by Equisetites ( Equisetostacliys ) suecicus (Nath.) Halle 1908. Some of the specimens of supposed Perinopollenites elatoides grains which Danze-Corsin and Laveine illustrate (in Briche et al. 1963, pi. 8, figs. 4-6) lack a pore and resemble P. allenii. Gould (1968) has described some remarkably similar spores from a cone of Equisetum laterale Phillips 1829 recovered from the Marburg Sandstone (Jurassic) of Australia. Acknowledgements. This paper was prepared in the Department of Geology, Cambridge University with the help of a N.E.R.C. studentship. Material from the Cuckfield No. 1 Borehole of the Geological Survey and access to the log was made available to Mr. N. F. Hughes by Dr. R. G. Thurrell. I thank Mr. J. G. Duckett of the Cambridge University Botany School for information. I am indebted to Mr. N. F. Hughes for discussion and critical reading of the manuscript. REFERENCES allen, p. 1941. A Wealden soil bed with Equisetites lyelli (Mantell). Proc. Geol. Ass., London, 52, 362-74. • 1947. Notes on Wealden fossil soil-beds. Ibid. 57, 303-14. 1959. The Wealden environment; Anglo-Paris basin, Phil. Trans. R. Soc. London, 242B, 283-346. 1962. The Hastings beds deltas: recent progress and Easter field meeting report. Proc. Geol. Ass., London, 73, 219-43. 1967a. Origin of the Hastings Facies in North-western Europe. Ibid. 78, 27-105. 19676. Strand-line pebbles in the mid-Hastings beds and the geology of the London Uplands. Old Red Sandstone, New Red Sandstone and other pebbles. Conclusion. Ibid. 241-76, 3 pi. Anderson, f. w., bazley, r. A. b., and shephard-thorn, e. r. 1967. The sedimentary and faunal sequence of the Wadhurst Clay (Wealden) in boreholes at Wadhurst Park, Sussex. Bull. geol. Surv. Gt. Br. London, 27, 171-235. balme, b. e. 1957. Spores and pollen grains from the Mesozoic of Western Australia. Rep. Fuel Res. Sect. C.S.I.R.O. Aust., Chatswood, N.S.W., 1-48, 11 pi. 642 PALAEONTOLOGY, VOLUME 11 balme, b. e. and hennelly, j. p. f. 1956. Monolete, monocolpate and alete sporomorphs from Australian Permian sediments. Aust. J. Bot. Melbourne, 4, 54-67, 3 pi. Baxter, r. w. and leisman, g. a. 1967. A Pennsylvanian Calamitean cone with Elat elites triferens spores. Am. J. Bot. Lancaster, Pa. 54, 748-54, 2 pi. briche, p., danze-corsin, p., and laveine, j.-p. 1963. Flore infraliasique du Boulonnais. Mem. Soc. Geol. N. Lille, 13, 7-143, 11 pi. chandler, m. e. j. 1964. The Lower Tertiary floras of Southern England. 4. A summary and survey of findings in the light of recent botanical observations. Brit. Mus. (Nat. hist.) London, xii+ 151 pp. 4 pi. couper, r. a. 1953. Upper Mesozoic and Cainozoic spores and pollen grains from New Zealand. Palaeont. Bull. Wellington, 22, 3-77, 9 pi. 1958. British Mesozoic microspores and pollen grains. Palaeontographica, Stuttgart, B103, 75-129, 17 pi. dettmann, M. e. 1963. Upper Mesozoic microspores from south eastern Australia. Proc. R. Soc. Viet. Melbourne, 77, 1-148, 27 pi. could, r. e. 1968. Morphology of Equisetum laterale Phillips 1829, and E. bryanii sp. nov. from the Mesozoic of south-eastern Queensland. Aust. J. Bot. Melbourne, 16, 153-76, 3 pi. halle, t. g. 1908. Zur Kenntnis der mesozoischen Equisetales Schwedens. K. svenska Vetensk. Akad. Hand!. Uppsala, 43, (1), 3-56, 9 pi. Harris, t. M. 1961. The Yorkshire Jurassic Flora. I. Thallophyta-Pteridophyta. Brit. Mus. (Nat. hist.) London, 212 pp. hartung, w. 1933. Die Sporenverhaltnisse der Calamariaceen. Arb. Inst. Palaobot. u. Petr. Brenn- steine, Preuss. geol. Landesanstalt, 3, 96-149. hauke, r. 1963. A taxonomic monograph of the genus Equisetum subgenus Hippochaete. Nova Hed- wigia, Weinheim, 8, 1-123, 22 pi. hughes, n. f. and moody-stuart, j. c. 1967a. Palynological facies and correlation in the English Wealden. Rev. Paiaeobotan. Palynol. Amsterdam, 1, 259-68. - 19676. Proposed method of recording pre-Quarternary palynological data. Ibid. 3, 347-58. kedves, m. 1961. Etudes palynologiques dans le bassin de Dorog. — II. Pollen Spores, Paris, 101-53, 10 pi. nilsson, t. 1958. Uber das Vorkommen eines mesozoischen Sapropelgesteins in Schonen. Acta Univ. land. Lund. N.F. Avd. 2, (54-10), 3-111, 8 pi. potonie, R. 1966. Synopsis der Gattungen der Sporae dispersae, Teil IV. Beih. geol. Jb. Hannover, 72, 1-244, 15 pi. reissinger, a. 1950. Die ‘Pollenanalyse’ ausgedehnt auf alle Sedimentgesteine der geologischen Ver- gangenheit II. Palaeontographica, Stuttgart, B90, 99-126, 9 pi. rogalska, m. 1954. Spore and pollen analysis of the Liassic coal of Blanowice in Upper Silesia. Biul. Inst. Geol. Warszawa, 89, 5-46, 12 pi. (in Polish, English summary). Stubblefield, c. J. 1963. Mem. geol. Surv. Summ. Prog. 1962, London, 91 pp. 1967. Rep Inst. geol. Sci. 1966, London, 197 pp. wilson, l. r. 1943. Elater-bearing spores from the Pennsylvanian strata of Iowa. Am. Midi. Nat. Notre Dame, 30, 518-23. 1963. Elaterites triferens from a Kansas coal ball. Micropaleontology, New York, 9, 101-2. DAVID J. BATTEN Department of Geology Sedgwick Museum Typescript received 9 November 1967 Cambridge THE PALAEONTOLOGICAL ASSOCIATION COUNCIL 1968-9 President Professor Alwyn Williams, The Queen’s University, Belfast Vice-President Dr. W. S. McKerrow, University Museum, Oxford Treasurer Dr. C. Downie, Department of Geology, The University, Mappin Street, Sheffield, 1 Membership Treasurer Dr. A. J. Lloyd, Department of Geology, University College, Gower Street, London, W.C. 1 Secretary Dr. J. M. Hancock, Department of Geology, King’s College, London, W.C. 2 Assistant Secretary Dr. W. D. I. Rolfe, Hunterian Museum, The University, Glasgow, W. 2 Editors Mr. N. F. Hughes, Sedgwick Museum, Cambridge Dr. Gwyn Thomas, Department of Geology, Imperial College, London, S.W.7 Dr. I. Strachan, Department of Geology, The University, Birmingham, 15 Professor M. R. House, The University, Kingston upon Hull, Yorkshire Dr. R. Goldring, Department of Geology, The University, Reading Other members of Council Dr. F. M. Broadhurst, The University, Manchester Mr. M. A. Calver, Geological Institute of Sciences, Leeds Dr. C. B. Cox, King’s College, London Mr. D. Curry, Eastbury Grange, Northwood, Middlesex Dr. Grace Dunlop, Bedford College, London Dr. G. F. Elliott, 60 Fitzjohn Avenue, Barnet, Herts. Dr. A. Hallam, University Museum, Oxford Dr. Julia Hubbard, King’s College, London Dr. J. D. Hudson, The University, Leicester Dr. R. P. S. Jefferies, British Museum (Natural History), London Dr. J. D. Lawson, The University, Glasgow Dr. A. H. Smout, British Petroleum Company, Sunbury-on-Thames Professor II. B. Whittington, Sedgwick Museum, Cambridge Overseas Representatives Australia : Professor Dorothy Hill, Department of Geology, University of Queensland, Brisbane Canada: Dr. D. J. McLaren, Institute of Sedimentary and Petroleum Geology, 3303-33rd Street NW., Calgary, Alberta India: Professor M. R. Sahni, 98 The Mall, Lucknow (U.P.), India New Zealand: Dr. C. A. Fleming, New Zealand Geological Survey, P.O. Box 368, Lower Hutt West Indies and Central America: Mr. John B. Saunders, Geological Laboratory, Texaco Trinidad, Inc., Pointe-^-Pierre, Trinidad, West Indies Western U.S.A. : Professor J. Wyatt Durham, Department of Paleontology, University of California, Berkeley 4, Calif. Eastern U.S.A.: Professor J. W. Wells, Department of Geology, Cornell University, Ithaca, New York PALAEONTOLOGY VOLUME 11 • PART 4 «r CONTENTS Three new Tethyan Dasycladaceae (Calcareous Algae). By g. f. elliott 491 Colour markings in Phacops and Greenops from the Devonian of New York. By g. c. esker hi 498 Studies on Triassic fossil plants from Argentina. IV. The leaf genus Dicroidium and its possible relation to Rhexoxylon stems. By S. archangelsky 500 Astrocystites distans, sp. nov., an edrioblastoid from the Ordovician of Eastern Australia. By B. D. webby 513 Chubbina, a new Cretaceous alveolinid genus from Jamaica and Mexico. By E. ROBINSON 526 Famennian ammonoids from New South Wales. By t. ,b. h. jenkins 535 Aquilapollenites in the British Isles. By a. r. h. martin 549 Discohelix (Archaeogastropoda, Euomphalacea) as an index fossil in the Tethyan Jurassic. By J. wendt 554 Pedicellariae of two Silurian echinoids from western England. By D. b. blake 576 Macrocystella Callaway, the earliest glyptocystitid cystoid. By c. R. c. PAUL 580 A revision of some Upper Devonian foraminifera from Western Australia. By j. e. conkin and Barbara m. conkin " 601 A new medusoid (?) from the Silurian of England. By isles strachan 610 The atrypidine brachiopod Dayia navicula (J. de C. Sowerby). By E. v. tucker 612 Planicardinia, a new septate dalmanellid brachiopod from the Lower Devonian of New South Wales. By n. m. savage • 627 Probable dispersed spores of Cretaceous Equisetites. By D. J. batten 633 PRINTED IN GREAT BRITAIN AT THE UNIVERSITY PRESS. OXFORD BY VIVIAN R1DLER, PRINTER TO THE UNIVERSITY VOLUME 11 • PART 5 Palaeontology DECEMBER 1968 PUBLISHED BY THE PALAEONTOLOGICAL ASSOCIATION LONDON Price £3 THE PALAEONTOLOGICAL ASSOCIATION The Association was founded in 1957 to further the study of palaeontology. It holds meetings and demonstrations, and publishes the quarterly journal Palaeontology. Membership is open to individuals, institutions, libraries, etc., on payment of the appropriate annual subscription : Institute membership . . . . . £7. Os. (U.S. $20.00) Ordinary membership £5. Os. (U.S. $13.00) Student membership . . . . . £3. 0s. (U.S. $8.00) There is no admission fee. Student members are persons receiving full-time instruction at educational institutions recognized by the Council ; on first applying for member- ship, they should obtain an application form from the Secretary or the Treasurer. All subscriptions are due each January, and should be sent to the Membership Treasurer, Dr. A. J. Lloyd, Department of Geology, University College, Gower Street, London, W.C. 1, England. Palaeontology is devoted to the publication of papers (preferably illustrated) on all aspects of palaeontology and stratigraphical palaeontology. Four parts at least are published each year and are sent free to all members of the Association. Members who join for 1969 will receive Volume 12, Parts 1 to 4. All back numbers are still in print and may be ordered from B. H. Blackwell, Broad Street, Oxford, England, at £3 per part (post free). A complete set, Volumes 1-11, consists of 43 parts and costs £129. Special Papers in Palaeontology is a series of substantial separate works published by the Association. The subscription rate is £6 (U.S. $16.00) for Institute Members and £3 (U.S. $8.00) for Ordinary and Student Members. Subscriptions and orders by members of the Association should be placed through the Membership Treasurer. The following Special Papers are available. Members may obtain them at reduced rates through the Membership Treasurer. Non-members may obtain them from B. H. Blackwell, Broad Street, Oxford, England, at the prices indicated. Special Paper Number One (for 1967): Miospores in the Coal Seams of the Carboniferous of Great Britain, by A. H. V. Smith and M. A. Butterworth. 324 pp., 72 text-figs., 27 collotype plates. Price £8 (U.S. $22.00), post free. Special Paper Number Two (for 1968): Evolution of the Shell Structure of Articulate Brachiopods, by Alwyn Williams. 55 pp., 27 text-figs., 24 collotype plates. Price £5 (U.S. $13.00). Special Paper Number Three (for 1968): Upper Maestrichtian Radiolaria of California, by Helen P. Foreman. 82 pp., 8 collotype plates. Price £3 (U.S. $8.00). Special Paper Number Four (for 1969): Lower Turonian Ammonites from Israel, by R. Freund and M. Raab. 83 pp., 1 5 text-figs., 10 plates. Price £4 (U.S. $11.00). Available January 1969. Typescripts on all aspects of palaeontology and stratigraphical palaeontology are invited. They should conform in style to those already published in this journal, and should be sent to Mr. N. F. Hughes, Department of Geology, Sedgwick Museum, Downing Street, Cambridge, England, who will supply detailed instructions for authors on request (these are published in Palaeontology, 10, p. 707-12). © The Palaeontological Association, 1968 THE STRUCTURE OF VERTEBRARIA INDICA ROYLE by DIVYA DARSHAN PANT and R. SHANKER SINGH Abstract. Pulls of carbonised substance of Royle’s type of V. indica and 250 other axes referable to the same species have been studied. The other specimens were collected from the Raniganj and Giridih coalfields (Per- mian; Upper Damuda Group, Gondwana System). They all show the same kinds of secondary xylem, large parenchyma, and phloem-like cells. The different kinds of pitting and their frequency in the tracheids, the fre- quency of uniseriate rays of different heights, their length, and the range of variation in xylem character, have been determined. It has been found that characters of xylem described for V. raniganjensis are within the range of V. indica and this species may be regarded as a synonym of V. indica. None of our several axes of V. indica shows any trace of pith. Royle (1833) founded two species of Vertebraria from India ( V. indica and V. radiata), that were later recognized as lateral and sectional views of the same fossil. By a strange coincidence McCoy (1847) and Dana (1849) made the same error in giving two names ( V. australis and Clasteria australis) to laterally exposed and sectional views respectively of the same fossil. Arber (1905) and others have pointed out that all axes of this kind belong to a single species, Vertebraria indica, but Surange and Maheshwari (1962) have recognized two new species of the genus among compressions and Schopf (1965) has established a fourth species for a petrified axis, described earlier by Kraiisel. However, a review of the literature on Vertebraria gave us the impression that the characters of the various species of the genus were neither sufficiently known nor clearly demarcated. MATERIAL AND METHODS A large number of specimens of Vertebraria were collected from localities in the coal- fields of Raniganj and Giridih. About 300 celloidin pulls of coaly material were pre- pared from 250 compressions. The pulls were mounted in Canada balsam and examined in ordinary transmitted light under a compound microscope and also under phase contrast and dark field illumination. Averages have been generally calculated from 300 random counts but where this was not possible, the smaller number of counts has been mentioned against the averages. All figured specimens and slides of this paper form part of Divya Darshan Pant Collection of plant fossils, located in the Botany Department of the Allahabad University. DESCRIPTION OF EXTERNAL FEATURES Compressions of Vertebraria axes are either found lying almost vertical to the bedding plane and seen in sectional views or they lie horizontally in the layers of the Lower Gondwana rocks and are exposed in a lateral view (PI. 124, PI. 125, figs. 1-2; text-fig. 1a-c). Sectional views of Vertebraria axes show rays of carbon, all joined and radiating from a centre and separated from each other by rather wide bays of rock matrix. The [Palaeontology, Vol. 11, Part 5, 1968, pp. 643-53, pis. 124-7.1 C 6055 U u 644 PALAEONTOLOGY, VOLUME 11 rays are simple and almost equal, their ends may be slightly flattened, occasionally the surface of a bay may be covered by a film of carbon which joins two adjacent rays (PI. 125, fig. 1). The largest vertically preserved Vertebraria in our collection is 2-7 cm. in diameter. Lateral views of horizontally compressed Vertebraria axes are more com- mon. The thickest specimen in our collection is 9 cm. wide. The axes show 2 to 4 series of rectangular areas with 1 to 3 intervening longitudinal furrows or ridges. As a rule the rectangular areas are of almost equal size but sometimes they may appear very unequal (text-fig. 1b). Branching axes are not uncommon. As described by Oldham (1897) the branching of these axes takes place in two ways. Either the branching axis divides into two or more almost equal branches or the axes produce thinner branches from their sides (text-fig. 1a-c). The branches may show similar rectangular areas but some of the thinner axes have a central longitudinal groove or ridge and only occasional obscure horizontal marks (PI. 124, fig. 5). Some of these axes show attached roots (PI. 124, fig. 3). An attached root is also seen in a vertically compressed axis (PI. 125, fig. 1). As suggested by Pant (1956) the preservation of horizontally compressed Vertebraria is not fundamentally different from that of vertically compressed axes, except that their xylem rays radiating dorsally and ventrally become shorter or narrower in the radial direction by undergoing greater radial compression, while the radial dimension of their horizontally placed rays remains practically uncompressed. Naturally such axes are usually exposed along the plane of easier splitting, i.e., along the wider horizontally placed rays and these appear as two series of typical rectangular areas on either side of a median ridge or furrow representing the central longitudinal core of Vertebraria and the transverse sides of the rectangles represent the upper and lower limits of the wide parenchymatous bays. Occasionally the rock is fractured along one of the dorsally or ventrally placed vertically compressed rays (Pant 1956, text-fig. 2a, where towards the top left a few narrower rectangular areas are those of a vertically compressed ray). Additional longitudinal rows of rectangular areas may be seen at a point where a Vertebraria branches (text-fig. 1b; PI. 124, fig. 2). Many of our axes lie among leaves of Glossopteris, Rhabdotaenia, Pteronilssonia, and other specimens. In one of these a Glossopteris leaf at first sight appeared to be connected with a Vertebraria but a closer examination reveals that the apical end of the leaf was actually facing the axis. INTERNAL STRUCTURE Horizontally compressed axes Celloidin pulls of carbon from horizontally compressed axes and their branches with or without typical rectangular areas usually show the same kind of secondary xylem as EXPLANATION OF PLATE 124 Figs. 1-5. Vertebraria indica Royle. 1, Forked axis, No. 1804, xf. 2, Trifurcate axis showing wide rectangular areas in two series below and triseriate areas above; rectangular areas are not seen for some distance near the lower end; No. 1756, xf. 3, Axis with elongated rectangular areas and a few thin branches (roots) with scalariform xylem (cf. Lithorhiza tenuirama Pant), No. 1805, xf. 4, Fragment of a thick axis showing tetraseriate rectangular areas, No. 1808, X 5, Axis with a median furrow but its rectangular areas are ill defined due to the presence of only a few rather obscure transverse marks. No. 3052, xf. Palaeontology, Vol. 11 PLATE 124 PANT and SINGH, Permian Vertebraria PANT AND SINGH: STRUCTURE OF VERTEBRARIA INDICA ROYLE 645 s' text-fig. 1, a-c. Vertebraria indica. A, Axis forked into two almost equally thick branches. No. 1804, X2. b, Trifurcate axis showing short and wide rectangular areas above; rectangular areas not seen near the lower end (also PI. 124, fig. 2), No. 1756, X c, Axis showing thinner branch, No. 1789b, X 2. 646 PALAEONTOLOGY, VOLUME 11 is seen in a radial section. Naturally, in compressed wood, such pulls also show the radial or tangential walls of the tracheids often overlapping each other. At a few places, the xylem is even seen in a tangential view or in a partially radial and partially tangential view (text-fig. 2h; PI. 127, fig. 6). Radial views of xylem. (a) Royle's type. A pull from Royle’s type specimen No. V. 4189 in the British Museum (Natural History) shows the typical xylem of Vertebraria as if in a radial longitudinal section (PI. 125, fig. 3). The tracheids have uniseriate to tri-seriate oval to circular pits placed far apart or contiguous. Biseriate and triseriate pits are usually opposite but alternately arranged pits have also been observed (text-fig. 3d, e). Rays are uniseriate and ray fields show pits usually without any border (text-fig. 3f). A few large parenchyma cells like those reported by Pant (1956) are mostly broken down. ( b ) Specimens from Raniganj and Giridih coalfields. All our pulls of the carbon from horizontally preserved specimens of Vertebraria axes show well-preserved secondary xylem covering the entire surface of the rectangular areas (where carbon is preserved), from one side of the axis through the median ridge or furrow right up to the opposite side. On all our slides most of the xylem is seen as in a radial longitudinal section. A pith should have been visible in such radial views but no pull from the axes in our collection shows any trace of parenchyma cells in the centre, although ray parenchyma, phloem- like thin- walled cells and the typical large parenchyma cells of Vertebraria are usually well preserved. Instead, numerous pulls show undisturbed tracheids filling the central parts of the axes. Primary xylem. Primary xylem tracheids consisting of scalariform, annular, or spiral elements were described from pulls of Vertebraria by Walton and Wilson (1932) and Pant (1956). Pant described them as occurring towards the periphery of the axes, where they could even have belonged to associated or attached roots. Secondary xylem. The wood is mostly secondary and pycnoxylic. The tracheids are almost all of a uniform type without growth rings. The length of most tracheids is indeterminable in the pulls and this may suggest that they were usually longer than the size of our pulls. However, 60 complete tracheids have been observed; their length is 77-679 /a, (average 250 p , 8 105). As both ends of most tracheids are not seen, we presume that these sizes may represent the shorter elements in the xylem of Vertebraria. The width of the tracheids ranges from 10-56 p (average width of tracheids is 31 p, 8 20 5). The trachei- dal walls are up to 4 p thick. The ends of the tracheids taper gradually to a narrow point, the end walls being very oblique (PI. 126, fig. 1). The tracheids over the rectangular areas on the surface of the axes are straight and vertical but in the region of the horizontal ridges or furrows they are slightly curved. Arrangement of pits. The pits on radial walls are uniseriate, or multiseriate up to 5 series of pits (text-fig. 2 b, c, and g). A single tracheid with 6 seriate pits was also seen (text-fig. 2 e, PI. 126, fig. 2). EXPLANATION OF PLATE 125 Figs. 1-6, Vertebraria indica Royle. 1, Vertically compressed axis towards the top right-hand corner; part of transverse diaphragm (black) and an attached root also seen; No. 1375, X 11. 2, Vertically compressed axis showing marks of peripheral tissues, No. 1810, x2. 3, Pull from Royle’s holotype showing biseriate and triseriate pits and a two-cell high ray, V. 4189, x300. 4, Xylem of root attached to axis in Plate 124, fig. 3, showing scalariform tracheids, Slide No. 1805, X 350. 5, Portions of tracheidal walls of axis showing crossed pit pores, Slide No. 1809 Ba, X 500. 6, Tracheids show- ing uniseriate pits or mixed uniseriate and biseriate opposite pits, Slide No. 1808c, X 800. Palaeontology, Vol. 11 PLATE 125 PANT and SINGH, Permian Vertebraria PANT AND SINGH: STRUCTURE OF VERTEBRARIA INDICA ROYLE 647 text-fig. 2, a-h. Vertebraria indica. A, Portion of tracheid showing scalariform pitting (also PI. 127, fig. 4); Slide No. 586, X 1,000. b, Part of tracheid with uniseriate non-contiguous pits; Slide No. 1784 Da, X750. c, Portion of tracheid showing uniseriate and biseriate pits; Slide No. 1784 Da, X750. d. Irregularly-pitted tracheid showing oval pit-pores, some with crossed pit-pores; Slide No. 1807a, X 1,000. e, Portion of wide tracheid showing 6-seriate pits (also PI. 126, fig. 2); Slide No. 1807a, X 750. F, Portion of tracheid showing oval groups of pits with a central pit; Slide No. 1765 C2, X750. g, Tracheid showing tetraseriate and pentaseriate pits; Slide No. 1767c, X750. h, Uniseriate 3-cell-high ray in tangential view; Slide No. 1784 Da, x450. 648 PALAEONTOLOGY, VOLUME 11 The pits may be contiguous or far apart (text-fig. 2 b; PI. 125, fig. 5) and the portions of the same radial wall may be pitted in some regions and unpitted for varying distances in other regions. Multi- seriate pits may be opposite, araucarioid (i.e. with crowded alternate pits having hexagonal outlines) or irregularly arranged (PI. 126, fig. 3). As a rule unpitted portions of the wall are seen in tracheids with irregular pits. Some of these may show circular or oval groups of 3 to 9 pits (text-fig. 2 d, f). The same tracheid may show different kinds of pitting in different regions i.e., uniseriate or multiseriate, opposite or alternate, contiguous or non-contiguous (text-fig. 2 c; PI. 125, fig. 6; PI. 126, fig. 3). A random count of 300 tracheids showed about 11% with uniseriate pits, 47% opposite multiseriate pits, 12% alternate crowded multiseriate pits of araucarioid type, 28% irregular pits, and 2% showed portions of walls without pits. Out of the 47% tracheids with opposite pits 19% are biseriate, 18% triseriate, 8% tetraseriate, and 2% pentaseriate. Shape and size of pits. The typical shape of bordered pits may be circular or oval (text-fig. 2 b). Oval pits in multiseriate tracheids are placed horizontally but where uniseriate they may be horizontal or oblique. The pit pores are generally oval but occasionally rounded. The two opposite pores of a pit pair are usually crossed (text-fig. 2d; PI. 125, fig. 5). In the tracheids of the thinner axes pit pores are wide with a very thin border. In some regions of the pulls, the tracheids show large horizontally extended oval pits where a border is not discernible. Such pits may be uniseriate or multiseriate. Where uniseriate they almost look like scalariform thickenings (text-fig. 2a; Plate 125, fig. 4; Plate 127, fig. 4) and where multiseriate they often appear transitional between typical scalariform and pitted elements. Pant (1956) described such tracheids as in contact with large parenchyma cells. The circular pits are 4-8-5 p (average 5 p, 8 0-5). The length of the oval pits is 8-5-26-5 p (average length is 13 p, 8 2-5) and the breadth is 4-10 p (average breadth is 6 p, 8 1). The length of the oval pits whose borders are not clear is 10-31 p (average length is 21-5 p, 8 1-5) and the breadth is 4-10 p (average breadth is 7 p, 8 1). Secondary xyleni rays. Only uniseriate secondary xylem rays have been observed. These are 1-13 cells high (PI. 126, fig. 5; PI. 127, fig. 1), but more commonly they are only 1-4 cells high: Percentage of uniseriate rays of differing heights Height of rays in number of cells 1 2 3 45 6 7 13 Percentage 33 29 17 10 6 2 2 1 Frequency of rays per 10 mm.2 is 4-18 averaging 9 (60 readings, S 4). The rays are 1-48 cells long (average 1 1 cells, 8 4). Crossfield pits. Crossfields of V. indica where clearly visible show 1-3 elongated oval pits (text-fig. 3a; PI. 127, fig. 2). The oval pits are horizontally placed and seemingly simple. On the basis of our observa- tions, we believe that the numerous bordered crossfield pits described by Walton and Wilson (1932) and by Surange and Maheshwari (1962) are pits of tracheidal wall, other than the common wall between a ray cell and a tracheid. In a compression, this wall may overlap a crossfield and then the pits can be mistaken for crossfield pits. The length of the crossfield pits ranges from 11-32 p (average 18 p, 8 2) and breadth from 4 to 14 p (average 5 p, 8 1). Phloem like cells. Many pulls show thin elongated cells by the side of xylem which look like phloem (text-fig. 3 c; PI. 127, fig. 3). The length of these cells is from 17 to 238 p (average 61 p, 8 17) and breadth 4-35 p (average 10 p, 8 3). The thickness of the walls of these cells is up to 3 p. EXPLANATION OF PLATE 126 Figs. 1-5. Vertebraria indica Royle. 1, Xylem showing tapering ends of two tracheids and longitudinal walls of others with multiseriate pits, Slide No. 1807a, X 800. 2, Portion of wide tracheid (on the right side) showing six-seriate pits, Slide No. 1 807a, X 800. 3, Xylem showing triseriate pits, opposite in some regions but alternate and araucaroid in other parts, Slide No. 1754c, X 350. 4, Pull from thinner axis with obscure rectangular areas showing typical Vertebraria type of xylem, Slide No. 1784 Db, X 140. 5, Xylem showing a 7-cell-high ray. Slide No. 1807a, X 200. Palaeontology, Vol. II PLATE 126 PANT and SINGH, Permian Vertebraria PANT AND SINGH: STRUCTURE OF VERTEBRARIA INDICA ROYLE 649 text-fig. 3, a-f. Vertebraria indica. a, Portion of xylem showing simple, horizontally elongated, oval pits in cross-fields; Slide No. 1809 Ba, x 600. b. Pull showing xylem with simple pits overlapped by broad parenchyma cells, which tend to be in radial series; Slide No. 586, x 150. c, Pull showing thin- walled phloem like cells; Slide No. 1767a, X 150.D, Portion oftracheid from holotype showing triseriate opposite pits; V. 4189, X750. e, Pull from holotype showing tracheid with uniseriate, biseriate, and irregularly arranged pits; V. 4189, X 750. F, Xylem showing horizontally elongated, simple oval pits in crossfield; Holotype, V. 4189, X750. 650 PALAEONTOLOGY, VOLUME 11 Large parenchyma cells. In addition, sheets of large rectangular to polygonal parenchyma cells were seen overlying xylem almost in all preparations. Very often the cells tend to arrange in radial rows (text-fig. 3 b; PI. 127, fig. 5). The length of these cells is 21-228 p (average 79 p, 8 49) and breadth 21-84 p (average 34 p, 8 17). Tangential views of xylem. In portions of the pulls where the xylem appears in a partially or wholly tangential view, the xylem rays are seen and they are invariably uniseriate (text-fig. 2h). Where the rays are obliquely compressed, presenting a combination of tangential and radial views, the ray field pits appear to be simple and horizontally oval. The tangential walls of the tracheids are devoid of pits and their surface presents a uniformly granular texture (PI. 127, fig. 6). The xylem of axes with ob- scure areas is essentially similar, showing identical pitting, uniseriate rays, crossfield pits, phloem-like cells and large parenchyma cells. Vertically preserved axes Xylem and other cells from vertically preserved axes are also indistinguishable from those of horizontally compressed Vertebraria except for the presence of short tracheids with tapering or truncated ends over the transverse side of the rectangular areas. COMPARISON AND DISCUSSION Surange and Maheshwari (1962) described a new species V. raniganjensis, with xylem differing specifically from that of V. indica as described by Walton and Wilson (1932), Pant (1956), and Sen (1958). Surange and Maheshwari (1962) assumed that Royle’s type specimen of V. indica, No. V. 4189 in British Museum (Natural History) was without carbon. However, since one of us (D. D. P.) had examined Royle’s type of V. indica and seen carbon in it, we obtained through the courtesy of Dr. Pettitt a small pull from Royle’s type. Therefrom, we have been able to confirm that it shows exactly the same kind of xylem as has been described by Walton and Wilson (1932), Pant (1956), Sen (1958), and by us here from other specimens assignable to V. indica. This xylem is also similar to that of Surange and Maheshwari’s V. raniganjensis. Surange and Maheshwari described the following peculiarities of the xylem of V. raniganjensis : (a) The pits are ‘often’ six seriate. ( b ) The pits are ‘more than often’ arranged irregularly and in groups of 2 to 9 which are sometimes of stellate shape. (e) ‘Xylem rays if not common are also not scarce as in V. indica’ . They are usually 1-5 cells high, rarely up to 8 cells high. (d) Crossfield pits are 7-9 in number, bordered and with an oval pore and some fields show no pits. (e) Absence of large parenchyma cells. They did not, however, mention the frequency of tracheids with six seriate or irregu- larly arranged pits or of the xylem rays. The frequent occurrence of irregularly pitted EXPLANATION OF PLATE 127 Figs. 1-6. Vertebraria indica Royle. 1, Xylem showing 13-cell-high ray, Slide No. 1777b, x150. 2, Xylem showing horizontally elongated, simple oval pits in crossfields, Slide No. 1809 Ba, x800. 3, Pull showing thin-walled phloem-like cells, Slide No. 1767a, X 300. 4, Portion of tracheid showing scalariform pits, Slide No. 586, X 350. 5, Pull showing broad parenchyma cells, tending to be in radial series, Slide No. 1781c, x40. 6, Portion of xylem showing unpitted tangential walls, Slide No. 1789c, x 230. Palaeontology, Vol. 11 PLATE 127 PANT and SINGH, Permian Vertebraria PANT AND SINGH: STRUCTURE OF VERTEBRARIA INDICA ROYLE 651 elements of xylem and the occurrence of six seriate tracheids (even though rare) in axes whose xylem is otherwise exactly like that of V. indica suggests that such tracheids may not be used to distinguish another species. The frequency of rays in V. indica is rather variable. Calculated on the basis of figures and photographs given by Walton and Wilson (1932), Pant (1956), and Surange and Maheshwari (1962), the frequency of rays in the axes described by these authors is within the range determined for V. indica. As mentioned above we believe that there is an error in the description of crossfield pits by Walton and Wilson and by Surange and Maheshwari. The only remaining difference is the absence of large parenchyma cells in V. raniganjensis and their presence in V. indica ; however, the negative evidence of the absence of such cells in a pull may be mis- leading for even in two pulls of the same axis, such cells are often seen to be preserved in some regions and not in other portions. We have also re-examined Pant’s original pulls of V. indica from Mhukuru coalfield, Tanganyika, and they show the same kind of the xylem as the type. Accordingly we are convinced that axes described as V. rani- ganjensis and V. indica belong to the same ‘species’. Surange and Maheshwari (1962) described another new species, V. myelonis , which differed from V. indica in having a pith in the centre of the axes but no parenchymatous tissue was recorded in the ‘pith’ region. None of our pulls from the Raniganj and Giridih coalfields shows any trace of a parenchymatous pith although the xylem and parenchyma cells are usually well preserved. Surange and Maheshwari based their species on a single specimen which does show a narrow space (pith) in the centre between the two lateral series of rectangular areas, but they did not describe any parenchyma from this region and the space possibly represents the width of the solid central part of the xylem because we have actually observed solid xylem in this region in numerous specimens and also in the centre of vertically preserved axes. In the absence of any positive evidence of a pith it would be simpler, if this species too is regarded as a synonym of V. indica. MORPHOLOGY OF VERTEBRARIA Structural features Oldham believed that Vertebraria had a solid central core joined to a peripheral bark by a series of radiating longitudinal septa which could be branched (Oldham 1 897, pi. 4, fig. 2), and according to his description eight of these septa could usually be seen in a sectional view in vertically preserved axes. He thought that the longitudinal septa were joined at intervals by transverse partitions so that the axes of Vertebraria were visualized as having a ring of air chambers around a central solid core. In his view the transverse partitions are responsible for the septate appearance in horizontal preserva- tion. Zeiller (1896) thought that Vertebraria axes had an externally ridged form with transverse septa between the longitudinal ridges as in the rhizomes of the living fern Strutbiopteris. Both Oldham and Zeiller agreed in regarding the rectangular areas as representing empty spaces (internal or external). Walton and Wilson (1932), however, suggested that the axes of Vertebraria had a rayed stele practically without a pith; each carbonaceous ray in a vertically preserved Vertebraria was formed by compression of compact secondary xylem having uniseriate parenchymatous rays between its tracheids. These rays of xylem were joined with adjacent xylem rays on either side; rock-filled spaces 652 PALAEONTOLOGY, VOLUME 11 (bays) between the carbonaceous rays were not empty but originally occupied by a soft tissue. Pant (1956) agreed with Walton and Wilson on the basis of his study of a Stigmaria with rays similarly preserved, and he described large parenchyma cells which occupied the bays. An interpretation of the features of Vertebraria which is similar to that given earlier by Oldham has recently been given by Plumstead (1962), but she agrees with Walton and Wilson (1932) in believing that the chambers were filled with a soft tissue. However, her reconstruction visualizes a solid ‘central column communicating by means of large oval cavities with the rest of the axis’, and also a woody bark. Recently, Schopf (1965) has generally confirmed the reconstructions of the structural features of Vertebraria suggested by Walton and Wilson (1932). Our investigation of the extensive new material of Vertebraria is fully consistent with the views of Walton and Wilson (1932). Pulls from rays of carbon in vertically preserved axes show the same kind of xylem as the surface of the rectangular areas in horizontally compressed axes. Celloidin pulls of carbon from rectangular areas of horizontally preserved axes show tracheids which tend to be short, wide, and curved near the trans- verse portions but elsewhere the tracheids run uniformly in the longitudinal direction. We have also confirmed that the transverse connections joining adjacent carbonaceous rays are also formed by xylem. As already described by Pant (1956) xylem and phloem over rectangular areas is often overlapped by large parenchyma cells, which are clearly the remnants of a soft tissue which originally filled the bays. The absence of rectangular areas in thinner branches of Vertebraria leads us to the conclusion that the parenchy- matous bays may have been formed after secondary growth as in the axes of some modern lianas like Aristolochia. This may also explain the regular seriated arrangement of the large parenchyma cells. Stem or root Vertebraria has been generally recognized as the stem or rhizome of Glossopteris after Zeiller (1896) and Oldham (1897) described axes showing apparent attachment of Glossopteris leaves, and an axis described by Dolianiti (1954) seemingly confirms this. On the contrary Glossopteris bearing axes of Glossopteris described by Etheridge (1894), Seward (1910), and Walton and Wilson (1932), do not show any Vertebraria features. An explanation for this inconsistency has been offered by Arber (1905) and Pant (1956, 1962) who suggest that the features of Vertebraria are only those of its vascular cylinder and accordingly axes where the peripheral tissues are intact would appear different. Walton and Wilson (1932), and Thomas (1952) have, however, doubted the attachment of Glossopteris leaves on Vertebraria axes. On the basis of his study of some roots called Lithorhiza tenuirama , Pant (1958) sug- gested by implication that Vertebraria could be a root, and pointed out that the roots of L. tenuirama could sometimes be found attached to axes of Vertebraria and sometimes to axes which combined the features of roots as well as those of a Vertebraria. Such axes are somewhat thicker than undoubted roots but thinner than axes showing typical Vertebraria features. Roots arise from them endogenously. Unlike roots they show large parenchyma cells characteristic of Vertebraria but they entirely lack its rectangular areas. Lately, Schopf (1965) has strongly supported the root nature of Vertebraria on the basis of a petrified axis. In this connection he emphasizes the lack of leaf traces or PANT AND SINGH: STRUCTURE OF VERTEBRAR1A INDICA ROYLE 653 nodes by Vertebraria axes and also the endogenous insertion of its roots (but adventi- tious roots in stems are also endogenous). However, Vertebraria axes are sometimes dichotomised or trichotomised like stems or adventitiously branched roots, although Pant (1958) has also described some similarly forked roots. Pant (1967) has recently described a number of axes lacking rectangular areas of Vertebraria type but showing attached leaves of Glossopteris and this may also serve as a negative evidence favouring the root character. An investigation of the structural features of Glossopteris bearing stems which appear to differ from Vertebraria axes may help in solving the problem. Acknowledgements. We are grateful to Dr. J. M. Pettitt of the British Museum (Natural History), London, for sending us a pull from Royle’s type specimen and to Drs. D. D. Nautiyal, G. K. Srivastava, B. K. Varma, and Mr. K. B. Singh for help in the collection of Vertebraria axes and Dr. (Miss) P. F. Kidwai for help in taking some enlargements. We thank the State Council of Scientific and Industrial Research, Uttar Pradesh, for financial assistance. REFERENCES arber, E. A. N. 1905. Catalogue of the fossil plants of the Glossopteris flora in the Department of Geology, British Museum ( Natural History). London. DANA, J. D. 1849. Wilkes' United States exploring expedition, during the years 1838-1842 under the Command of Charles Wilkes, U.S.N. 10, Geology, New York. dolianiti, e. 1954. A Flora do gondwana inferior em Santa Catarina, IV O genero Vertebraria. Div. Geol. Min. Notas Prel. E. Estudos, 81, 1-5. etheridge, r. jun. 1894. On the mode of attachment of the leaves or fronds to the caudex in Glossop- teris. Proc. Linn. Soc. N.S.W., Ser 2, 9, 228-58. mccoy, F. 1847. On the fossil botany and zoology of the rocks associated with the coal of Australia. Ann. Mag. Nat. History, 20 (132), art 15, 145-57; 20 (134), art 28, 298-312. oldham, r. d. 1897. On a plant of Glossopteris with part of the rhizome attached, and on the structure of Vertebraria. Rec. Geol. Surv. India, 30 (1), 45-50. pant, d. d. 1956. On the two compressed Palaeozoic axes: Stigmaria ficoides in Gymnostrobus con- dition and Vertebraria indica. Ann. Bot. N.s. 20 (79), 419-29. 1958. Structure of some roots and spores from the Lower Gondwana (Permo-Carboniferous) of East Africa. Vijnana Parishad Anushandhan Patrika, 1 (4), 231-44. 1962. Some recent contributions towards our knowledge of the Glossopteris flora. Proceedings of the Summer School of Botany, Darjeeling, 302-10. 1967 (In press). On stem and attachment of Glossopteris leaves. Phytomorphology, 17. plumstead, E. p. 1962. Fossil flora of Antarctica. Trans-Antarctic Expedition {London), Rept. 9, Geology, pt 2, 154 p. royle, J. F. 1833. Illustrations of the botany and other branches of natural history of the Himalayan mountains, and of the flora of Cashmere. London . schopf, j. m. 1965. Anatomy of the axis in Vertebraria. Geol. and Paleont. of the Antarctic. Antarctic Research Series, 6, 217-28. sen, j. 1958. Further studies on the structure of Vertebraria. Bot. Not. 3, (2), 436-48. seward, a. c. 1910. Fossil plants, vol. 2. Cambridge. surange, k. r. and maheshwari, h. k. 1962. Studies in the Glossopteris flora of India-1 1 . Some observa- tions on Vertebraria from the Lower Gondwanas of India. Palaeobotanist 9, (1, 2), 61-7. thomas, h. H. 1952. A Glossopteris with whorled leaves. Ibid. 1, 435-8. walton, j. and wilson, j. a. r. 1932. On the structure of Vertebraria. Proc. Roy. Soc. Edinburgh, 52, pt. 2, 200-7. zeiller, r. 1896. Etude sur quelques plantes fossiles, en particulier Vertebraria et Glossopteris, des environs de Johannesburg (Transvaal). Bull. Soc. Geol. France, Paris, ser. 3, 24, 349-78. DIVYA DARSHAN PANT R. SHANKER SINGH Department of Botany, The University, Allahabad, India Typescript received 11 January 1968 THE GRAPTOLITE ASSEMBLAGES AND ZONES OF THE BIRKHILL SHALES (LOWER SILURIAN) AT DOBB’S LINN by P. TOGHILL Abstract. The graptolite assemblages and lithologies of the Birkhill Shales (Lower Silurian) at Dobb’s Linn are described. The condensed 141 ft. (43 m.) euxinic sequence is divided into eight graptolite zones in descending order : Rastrites maximus, Monograptus sedgwickii, M. convolutus, M. gregarius, M. cyphus, Cystograptus vesi- culosus, Akidograptus acuminatus, and Glyptograptus persculptus. These zones are compared with the original work of Lapworth (1878), and with equivalent zones elsewhere in Great Britain. The base of the Middle Llan- dovery at the type locality for the graptolitic Llandovery (Valentian) is taken at the base of the gregarius Zone ( triangulatus Zone in Wales), and not at the base of the magnus Zone, as is the current practice in Wales. The classic locality of Dobb’s Linn, and the surrounding Moffat area, have been left relatively untouched by research workers since the admirable account of Lapworth (1878). Because of the lack of new information from the Moffat area, recent reviews of Southern Uplands geology (Walton 1963, 1965) have had to discuss the Llandovery geology of the area using information obtained in, and unaltered since, the late 1800s, and thus their discussions are rather speculative. The present work does not, however, suggest any major errors by Lapworth. It is only the application of knowledge obtained since 1878 which has allowed alterations and additions to his classic results. The founda- tion for a Llandovery graptolite zonal scheme was laid down by Lapworth (1878, 1880), and his original research was quickly followed up in the Lake District by Marr and Nicholson (1888). Table 2 is a correlation of the large number of zonal schemes which have been used in Great Britain for the Llandovery (Valentian) up to the base of tur- riculatus Zone. The most recent scheme given by Curtis (1961) is in fact a compilation of most zones recognized in Wales, and based on the original research of O. T. Jones (1909), and Jones and Pugh (1916, 1935). The present work reappraises the graptolite fauna of the condensed 141 ft. (43 m.) euxinic Birkhill Shale sequence at the type locality, redefines the Llandovery graptolite zones exposed at Dobb’s Linn, and compares them with equivalent zones elsewhere. The condensed sequence at Dobb’s Linn is almost completely fossiliferous and thus is the most suitable place to define Llandovery grap- tolite zones, not only because it is the type locality for the graptolitic Llandovery (Valentian) but also because the equivalent sequence in Wales is a thick series (1,000 ft.) of coarse sediments in which the graptolite horizons are separated by considerable thicknesses of unfossiliferous strata (Jones 1909, p. 504). POSITION OF SECTIONS Dobb’s Linn lies 10 miles north-east of Moffat, and at the head of the Moffat Water. A complicated isoclinal fold structure, with Caledonoid trend, brings the graptolitic con- densed sequence of Moffat Shales (Caradoc-Llandovery) to the surface from below [Palaeontology, Vol. 11, Part 5, 1968, pp. 654-68.] TOGHILL: GRAPTOLITE ASSEMBLAGES AND ZONES OF BIRKHILL SHALES 655 the overlying Gala Greywackes (Upper Llandovery). This work is only concerned with the highest (Birkhill) division of the Moffat Shales, from the base of the Silurian ( persculptus Zone) up into the maximus Zone of the Upper Llandovery. The south-east limb of the fold is highly contorted and no detailed measurements are possible. The north-west (inverted) limb however shows an excellent section of the Birk- hill Shales, particularly in the Linn Branch, and on the North and Main Cliffs (text- fig. 1). A large waterfall has developed in the Linn Branch at the junction of Gala Greywackes and Birkhill Shales, and below this all but the lowest beds of the Birkhill Shales are exposed, the beds being inverted and dipping downstream at between 40° and 70°. The basal beds are cut out by a strike fault, and there is also a considerable amount of strike faulting in the convolutus, gregarius , and upper cyphus Zones. The Corrie and North Cliff show a continuous section from the lower sedgwickii Zone down to the base of the Silurian with little or no strike faulting. The Main Cliff (text-fig. 1) provides an excellent section across the Ordovician-Silurian boundary, and is by far the best place for measuring up the basal beds of the Birkhill Shales. Using these sections a detailed lithological sequence (text-fig. 2) has been worked out for the Birkhill Shales. Descending from the Gala Greywackes, the maximus and sedgwickii Zones have been measured up in the Upper Linn Branch (section A, text-fig. 1). Conspicuous lithological horizons in the lower sedgwickii Zone then allowed the section to be transferred to the Corrie and North Cliff (section B, text-fig. 1) where the remainder of the sequence was measured down to the base of the persculptus Zone. The lower part of the North Cliff' section ( persculptus to cyphus Zones) was correlated with, and supplemented by, collections from the Lower Linn Branch section (C) in the cyphus, vesiculosus and upper acuminatus Zones, adjacent to the West Fault, and the Main Cliff section (D) in the persculptus , acuminatus and vesiculosus Zones. THE LITHOLOGIES OF THE BIRKHILL SHALES Three main lithological types are present: massive black pyritic graptolitic mudstone, massive barren grey mudstone, and beds of soft, pale grey to orange, clay-like deposits. Lapworth (1878) referred to this latter conspicuous lithology under a variety of names including, ‘white clay bands’, ‘soft shaly clay’, and ‘white lines’, but the term claystone has recently been applied to these deposits by Walton (1965, p. 185). These beds, which often contain lines of calcareous nodules, weather out very easily and may be of volcanic origin (Walton 1963, pp. 85, 89; 1965, p. 185). This lithology makes up a large propor- tion of the Birkhill Shales and also occurs in the underlying Hartfell Shales, and persists into the Gala Greywackes. If volcanic in origin it indicates a period of vulcanicity from the Caradoc into the Upper Llandovery. The lower 65 ft. of the succession is made up of alternations of black mudstone and pale soft claystone, with occasional nodule bands, but without any grey mudstone at all. These lower Birkhill Shales follow directly on top of grey mudstones of the Hartfell Shales. The black mudstones are generally between 3 in. and 1 ft. thick and the intervening claystones are usually between 1 in. and 3 in. thick, but occasionally up to 1 ft. Within the upper acuminatus, vesiculosus, and lower cyphus Zones is a conspicuous 10-ft. thickness of massive black mudstone with very few claystones. These massive beds, Lapworth’s vesiculosus Flags, are a conspicuous unit of the lower Birkhill Shales well exposed in the Lower Linn Branch. The upper 76 ft. of 656 PALAEONTOLOGY, VOLUME 11 text-fig. 1. Locality map of Dobb’s Linn, showing position of sections. TOGHILL: GRAPTOLITE ASSEMBLAGES AND ZONES OF BIRKHILL SHALES 657 o o o 2E55-I Block mudstone Grey mudstone with cloystone Claystone/with calcareous nodules 0 12 3 4 Vertical scale in metres text-fig. 2. Vertical section through the Birkhill Shales, showing the position of fossiliferous horizons, etc. Although an attempt has been made to indicate the frequency of claystones, the section is diagram- matic, particularly below 55 ft. The positions of all nodular claystones, and of individual fossiliferous beds above 55 ft. are precise. the Birkhill Shales is made up of alternations of grey and black mudstones, varying in thickness from mere bedding plane bands to beds of 1 ft., and many soft claystones with occasional nodule bands. The inter-relationships of grey and black mudstones are shown in text-fig. 2. The highest graptolitic horizons in the Birkhill Shales ( maximus Zone) occur as thin bands | in. to 1 in. thick in a dominantly grey mudstone sequence. About 12 ft. above the highest black band the first greywacke appears in the Upper Linn Branch, 658 PALAEONTOLOGY, VOLUME 11 below the waterfall. It is 6 in. thick and is taken as the base of the Gala Greywackes, as indicated by Lapworth (1878, p. 322). This bed is followed by 14 ft. of massive grey mudstones, thin greywackes, and claystones, before the massive greywackes at the base of the waterfall are reached. From the above description, and from text-fig. 2, it can be seen that within the Birkhill Shales there is a gradual change from the exclusively euxinic facies of the lower Birkhill Shales, through the grey and black mudstones of the convolutus and sedgwickii Zones, into the dominantly grey mudstones of the rnaximus Zone, and finally into grey- wacke facies. Throughout the Moffat area the base of the Gala Greywackes has been found to be highly diachronous, and this will be further explained when describing the fauna of the R. maximus Zone. THE CHARACTERISTIC GRAPTOLITE FAUNA OF THE BIRKHILL SHALES The various graptolite zones have been divided up into fossiliferous horizons (1-46) which are indicated on text-fig. 2. The detailed fauna of any horizon can be ascertained from Table 1 . Zone of Glyptograptus persculptus This zone is well exposed on the Main CliiT (section D, text-fig. 1), but is less accessible at the top of the North Cliff (base of section B). On the Main Cliff the black mudstones of the Birkhill Shales follow barren grey mudstones of the anceps Zone of the Hartfell Shales. The highest fossiliferous band with D. anceps is approximately 10 ft. below the base of the Birkhill Shales. The persculptus Zone (horizon 46, text-fig. 2) comprises 3-|- ft. (T06 m.) of black mudstone with thin claystones. The mudstones are often weathered to a drab grey colour and the fauna is then extensively destroyed. At the base of the zone is a 4-in. bed crowded exclusively with Climacograptus scalar is normalis. The remainder of the zone is characterized by this species with the zone fossil, and Climacograptus medius, C. scalaris miserabilis and Diplograptus modestus sd. Zone of Akidograptus acuminatus 16| ft. (5-0 m.) of black mudstones with many thin claystones are assigned to this zone (horizons 43-5). It is best exposed on the Main Cliff (section D), but the highest beds are well exposed in the Lower Linn Branch (section C). The whole zone is exposed on the North Cliff (section B) but is rather inaccessible. The lower part of the zone weathers in a similar fashion to the persculptus Zone, but the highest beds are hard and massively bedded, and form the lowest part of Lapworth’s vesiculosus Flags. The base of the zone is marked by the appearance of Akidograptus and the zone is characterized by A. acuminatus sd. and A. ascensus, neither of which occur outside the zone. The latter occurs commonly before the main burst of A. acuminatus, which com- mences in the middle of the zone. Glyptograptus persculptus is common in the lower part and Climacograptus trifilis in the middle of the zone. Cystograptus vesiculosus appears in the middle and is common at the top of the zone, and Diplograptus modestus occurs throughout. Climacograptus scalaris normalis and C. medius are common throughout, and C. rectangularis appears in the upper part. TOGHILL: GRAPTOLITE ASSEMBLAGES AND ZONES OF BIRKHILL SHALES 659 Zone of Cystograptus vesiculosus This zone is represented by 4 ft. 3 in. (1*3 m.) of massive black mudstone with a few claystones (horizons 41-2), and is best exposed in the Lower Linn Branch (section C). The base is marked by the most conspicuous horizon in the whole of the Silurian, the appearance of the Monograptidae, together with Dimorphograptus and Rhaphidograptus. The zone is characterized by the earliest monograptids, with the zone fossil, and dimor- phograptids with long uniserial portions: D. elongatus and R. extenuatus. One particular bedding plane at the base of the zone is crowded with Monograptus cypluts praematurus Toghill (1968), which with M. atavus Jones is the earliest monograptid. The latter is common throughout the zone and occurs in the upper part with M. acinaces and M. incommodus. Climacograptus innotatus appears at the base of the zone and occurs throughout as does Diplograptus modestus sd. Glyptograptus tamariscus sd. appears and climacograptids are abundant including C. medius, C. rectangular is, and C. scalaris normalis. Zone of Monograptus cyphus This zone is represented by 24 ft. (7-3 m.) of massively bedded black mudstones with yellow and orange claystones, which are rare towards the base. The zone (horizons 34-40) has been measured up in the Lower Linn Branch (section C) and on the North Cliff (section B). The higher parts of the zone in the Lower Linn Branch are extensively strike-faulted. The overlying gregarius Zone has at its base a most conspicuous 1 ft. claystone with large calcareous nodules, and this, the lowest nodular claystone, is well exposed on the North Cliff. At the base of the zone dimorphograptids with short uniserial portions appear, including: D. confertus, D. confertus swanstoni, D. erectus, D. decussatus sd., and D. longissimus. Of these the first three are the commonest, and the first two persist into the upper parts of the zone. D. physophora is rare but probably occurs throughout as does Rhaphidograptus toernquisti. Monograptus cyphus is rare at the base of the zone but soon becomes abundant and ranges throughout the zone as do M. sandersoni, M. incom- modus, M. atavus, and M. acinaces. M. gregarius and M. revolutus appear in the middle of the zone and are common at the top. Cystograptus vesiculosus and Climacograptus innotatus are limited to the basal beds where the latter is common. Climacograptus medius and C. rectangularis are abundant in the lower part of the zone, but do not reach the top, and neither does C. scalaris normalis. Glyptograptus tamariscus sd. occurs throughout the zone. Zone of Monograptus gregarius 26 ft. (7-9 m.) of strata are assigned to this zone (horizons 29-33) as measured on the North Cliff (section B). Black mudstones with thick nodular claystones occupy the lower 18 ft. but these grade upwards into banded grey and black mudstones with many claystones. The base of the zone is marked by the appearance of Monograptus triangulatus sd. which occurs throughout the zone and is particularly common in the lower part. M. gregarius and M. revolutus occur throughout the zone, but M. sandersoni, M. incom- modus, and M. atavus are limited to the lower part. M. communis is characteristic of the upper part of the zone, and Rastrites longispinus and R. peregrinus appear in the middle. x x C G055 660 PALAEONTOLOGY, VOLUME 11 A conspicuous horizon (30) 8 ft. from the top of the zone yields abundant specimens of Monograptus fimbriatus with many petalograptids, including Petalograptus minor , P. palmeus latus, P. palmeus ovato-elongatus, as well as a few specimens of Diplograptus magnus. This is the only horizon with D. magnus and M . fimbriatus, and the main burst of the latter occurs above that of M. triangulalus s.l., as suggested by Sudbury (1958). This horizon is presumably equivalent to the magnus Band in Wales, although here it is only 3 in. thick. The unfossiliferous beds (text-fig. 2) which overlie horizon 30 but underlie the highest fossiliferous horizon (29) of the gregarius Zone might be considered part of the magnus Zone. This 3-ft. thickness of beds compares with 27 ft. assigned to the magnus Zone in Wales, although there the actual magnus Band at the base is only 6 in. thick (Jones and Pugh, 1935, p. 275), the remainder of the zone being relatively unfossiliferous. In horizon 29 Monograptus leptotheca is common, M. denticulatus appears, and this is the only horizon with M. argenteus. It can be equated with the argenteus Zone of the Lake District (Marr and Nicholson 1888), and the leptotheca Zone of Wales, which are here considered to be equivalent. M. argenteus in abundance seems to be restricted to the Lake District, its place in Wales and Scotland being taken by M. leptotheca. Zone of Monograptus convolutus The foot of the North Cliff (section B) provides an excellent section of this zone (horizons 22-8) which is 17J ft. (5-35 m.) thick and made up of two groups of highly fossiliferous black mudstones with claystones separated by barren grey mudstones and claystones (text-fig. 2). The upper group of fossiliferous beds (horizons 22-5) is, in fact, Lapworth’s original clingani Band, and the lower group his cometa Zone. The detailed fauna of any horizon can be ascertained from Table 1. Faunally the base of the zone is marked by the appearance of Monograptus lobiferus in abundance, with M. convolutus. The former is limited to the basal beds of the zone but the latter persists, with M. clingani , throughout the zone and is found as high as the lower maximus Zone. M. triangulatus, M. communis, and M. leptotheca are all present in the basal beds, but M. gregarius has disappeared. Cephalograptids, and petalograptids with protracted proximal ends are characteristic including in the lower part, Petalo- graptus folium and Cephalograptus tubulariformis, and C. cometa in the upper part of the zone. Rastritids are common and include R. hybridus, R. peregrinus, R. longispinus, and R. approximatus geinitzi. In the higher group of black mudstones the following monograptids are common: M. clingani, M. limatulus, M. crenularis, M. convolutus, and M. decipiens, together with Cephalograptus cometa, Glyptograptus tamariscus incertus, and Orthograptus bel/ulus. Of these species M. crenularis is very typical of the upper convolutus Zone. Zone of Monograptus sedgwickii 27J ft. (8-4 m.) of strata are assigned to this zone (horizons 8-25) which is best exposed in the Upper Linn Branch (section A), and the Corrie under the North Cliff (section B). Lithologically the zone can be divided into three units. A lower 7 ft. of barren grey mudstones with claystones, best seen in the Corrie; a middle 6 ft. of fossiliferous black mudstones with claystones (horizons 14—21) exposed in the Corrie and Upper Linn Branch; and an upper 14J ft. of grey mudstones with claystones and a few fossiliferous TOGHILL: GRAPTOLITE ASSEMBLAGES AND ZONES OF BIRKHILL SHALES 661 Q O A.ac. u M. cyphus M.gregarius M. convolutus M. sedgwick ii R.max imus Fossiliferous horizons 5 - "t o ■s ft ft 5 s ft 5 o ft s in ft a 5 £ ft 5 o ft 2 £ 2 2 2 2 2 - - - - - - - o r r • o c r r r o 9 o • r © o r o o r o • o O , © o o o o O , O r vor. normal is Lapw. o o o o o o miserabilis E.and W. o o o Cystograptus vesiculosus (Nich.) o • 9 o o o 9 Orthograptus bellulus ( Tornq.) cyperoides (Tbrnq.) r o o o O o O r insectiformis ( Nich.) r mutabilis E.and W. o o o truncatus var abbreviatus E.and W. Glyptograptus persculptus (Salter) sinuatus ( Nich J tamariscus (Nich.) var. incertus E.andW. serratus E.and W. var. barbatus E.andW. Diplograptus modestus L a p w. var. pa r v ul 0 s ( H . La pw.) Petalograptus palmeus (Barr.) var. tenuis (Barr.) latus (Barr.) ovato- elongatus (Kurck) minor Elies fol ium ( H i s.) Cepha lograptus cometa ( Gei n .) tubulariformis (Nich.) Akidograptus acuminatus s.l.(Nich.) ascensus Davies Rhaphidograptus toernquisti (E.andW.) extenuatus (E.andW.) Dimorphograptus confertus (Nich.) var swanstoni Lapw. erectus E.and W. decussatus E.and W. var partiliter E.andW. elongatus Lapw. longissimus ( Kurck) physophora (Nich.) Rastrites maximus Carr. linnaei Barr. distans Lapw. fugax Barr. hybridus Lapw. peregrinus Barr. longispinus Perner approximatus var. geinifzi Tornq. Monograptus cyphus Lapw. var. praematurus Toghill gregarius Lapw. acinaces TiJrnq. leptotheca Lapw. regularis TcJrnq. jaculum Lapw. nudus Lapw. concinnus Lapw. revolutus Kurck difformis Tdrnq. argenteus ( Nich.) limatulus Tttrnq. atavus Jones sandersoni Lapw. incommodus Tornq. argutus Lapw. crenularis Lapw. gemmatus ( Barr.) attenuatus (Hopk.) sedgwickii ( Port 1 . ) ha 1 1 i (Barr.) lobi ferus ( M'C oy ) clingani (Carr.) millipeda (M'Coy) convolutus ( H is.) decipiens Tornq. triangulatus s.l. (Hark.) denticulatus Tbrnq. spiralis (Gein.) involutus Lapw. circularis E.and W. communis Lapw. fimbriatus (Nich.) . i nter m ed ius L apw. Diploqraptus maqnus H Lapw. ® . • ? 0 9 ? 9 ? o o o o r o o o o o 9 o o O r r r r o 9 • o r o o o o o o o o o r r r f r r r r r o O r o o o • • r o r 9 r o o o • o o o o • o • o I r r 1 o o © r • • • v • o o o o o o r r r r r > v , o O 9 9 r 9 9 r O o o r o o o o o o r r o r r • o 9 © © 9 • 9 9 r ? o o o o ? © o o ? r r o 9 © o o « 9 9 r • 9 o o r r r o 9 9 o o 9 9 9 o r O ? r o O r ? O 7 © o o o o o o i , o o r © 9 r • © © © o © o r r r © • 9 o « o o ® • o o o o o o • 9 i o O o o r o o 9 o 9 © 9 9 r r r 9 o e o r r r o o : e o o o r • r © o o o o 9 o 9 9 O 9 2. r r o o 9 9 9 O o r r o • • • o o r o o © o O r r r o o o o o ? - 9 © © o o r O • o • o r O r r r r 9 9 o • common o occurs r rare ? questionably present § 3 3 - o ?; s 3. con CN - o a IS o £ s ;; ft O o- co K o «n * r> CN - o ■> - o - - table 1. The vertical distribution of the Graptoloidea in the Birkhill Shales at Dobb’s Linn. 662 PALAEONTOLOGY, VOLUME 11 black seams (horizons 8-13) exposed in the Upper Linn Branch. The middle unit is characterized by an abundance of Monograptus sedgwickii, M. regularis, and M. jacu- lum., with triangulate monograptids in the lower part, including M. convolutus, M. involutus, and M. decipiens. Glyptograptids are common and Petalograptus palmeus temds occurs with Climacograptus sealaris. The upper 144 ft. contain a similar but poorer fauna in which M. spiralis and M. attenuatus appear. Zone of Rastrites maximus The highest 214 ft. (6-55 m.) of the Birkhill Shales is assigned to this zone. Grey mudstones and claystones predominate, and grade up into the Gala Greywackes, the base of which is taken at the first greywacke seen in the Upper Linn Branch (section A), as indicated by Lapworth (1878, p. 322). Black mudstones are restricted to the lowest 9 ft. Seven separate fossiliferous horizons occur, and of these the two lowest yield a large fauna (Table 1). In these basal beds Monograptus sedgwickii, M. regularis, M. nudus, M. convolutus, and M. circularis are common, but the latter two species only occur commonly on one particular bedding plane. These common species occur with M. spiralis, M. attenuatus, M. gemmatus, M. decipiens, M. intermedius, Climacograptus extremus, and Petalograptus palmeus tenuis. At this lower level Rastrites maximus and M. halli are rare. The highest fossiliferous horizon of the Birkhill Shales at Dobb’s Linn is a 4-in. composite band of four black mudstones up to 1 in. thick separated by thin claystones. It yields Rastrites maximus, R.fugax, R. linnaeil, R. distansl, R. hybridus, Monograptus halli, M. sedgwickii, M. convolutus, M. spiralis, and M. intermedius. Owing to the diachronism of the overlying Gala Greywackes (see below), higher horizons not represented by fossiliferous strata at Dobb’s Linn, but occurring south east of the Caledonoid strike line through Dobb’s Linn, are characterized by the following common species: Rastrites maximus, Monograptus halli, and M. spiralis, with M. run- cinatus, M. nudus, and M. turriculatus minor. THE LLANDOVERY GR APTOLITE ZONES AT DOBB’S LINN The graptolite zonal scheme used here has been arrived at by considering the original scheme of Lapworth (1878, 1880), as modified by Elies and Wood (1913), and the large number of schemes erected in Wales by Jones (1909), Jones and Pugh (1916), Davies (1926), and others, as summarized by Curtis (1961). Table 2 shows an attempted cor- relation between the zones used here, and other schemes used elsewhere. The Glyptograptus persculptus Zone was defined at Dobb’s Linn for the first time by Toghill (1968), although the species was first recorded from there by Davies (1929). The validity of the zone is emphasized by the absence of Akidograptus, which does not appear until the overlying acuminatus Zone. The Akidograptus acuminatus Zone as defined here is approximately the same as defined by Lapworth (1878, pp. 250, 252, 318-20), although the basal beds have been separated off as the persculptus Zone. The base of the Cystograptus [Orthograptus] vesiculosus Zone has been taken at the incoming of the monograptids, as suggested by Elies (1925, p. 344). The re- stricted thickness (4 ft. 3 in.) used here is only the lower part of the zone as originally TOGHILL: GRAPTOLITE ASSEMBLAGES AND ZONES OF BIRKHILL SHALES 663 defined by Lapworth (1878, pp. 250, 252, 319-20), but the zone has been strictly defined on the ranges of the earliest monograptids, as given by Elies (1925), and dimorpho- graptids. The fact that the zone is so thin compared with the 16 ft. assigned to it by Lapworth is due to the more precise fixing of monograptid ranges, particularly those of M. cyphus, M. sander soni, M. incommodus, and M. atavus. Lapworth (1878, p. 319) did not give a very detailed monograptid fauna from his vesiculosus Zone and lower gre- garius Zone, recording only M. ‘ tenuis' and M. attenuatus from the vesiculosus Zone, and of course M. incommodus and M. atavus were not described until later. Elies (1925) did not give any thicknesses for the vesiculosus Zone, and thus this is the first time that the zone has been defined at Dobb’s Linn (the type locality) using a comprehensive mono- graptid fauna. The monograptid ranges agree with Elles’s definition of the zone (1925), but she recorded Dimorphograptus confertus and D. confertus swanstoni as being charac- teristic of the vesiculosus Zone. These are here taken as indicating the base of the over- lying cyphus Zone together with the appearance of M. cyphus and M. sandersoni. Thus, the vesiculosus Zone as used here is only the lower part of the zone as defined by Lapworth (1878), Elies and Wood (1913), and Elies (1925). It is probably equivalent to the atavus Zone in Wales (Jones 1909, Curtis 1961). The cyphus Zone includes the remainder of Lapworth’s vesiculosus Zone and part of his gregarius Zone. It is equivalent to the cyphus and acinaces Zones in Wales (Jones, 1909). From Table 1 it can be seen that the ranges of M. atavus and M. acinaces are almost identical, the former having the larger range, and the latter occurs, in the later part of its range, with M. cyphus. Although the vesiculosus Zone as defined here is equivalent to the atavus Zone in Wales, the former title is preferred on historical grounds, Dobb’s Linn being the type locality. The M. gregarius Zone as defined here is equivalent to the triangulatus, magnus, and leptotheca Zones in Wales (Curtis, 1961). The base is taken at a conspicuous faunal horizon, the appearance of triangulate monograptids, and is probably the base of Lap- worth’s Upper Division of his gregarius Zone (1878, pp. 319, 321), where he records M. triangulatus ‘in profusion’. The main horizon (29) with M. leptotheca is only 1 ft. thick, and that (30) with Diplograptus magnus only 3 in. thick. The former horizon con- tains M. argenteus and is presumably equivalent to the argenteus Zone in the Lake District. M. triangulatus s.l. ranges right through the gregarius Zone as defined here as does the zone fossil, and the title gregarius Zone is preferred on historical grounds, although it must be noted that it is equivalent to the triangulatus s.l. Zone of W. D. V. Jones (1945). Jones and Pugh (1935, pp. 275, 276) stated that the magnus Zone was made up of 27 ft. of mainly unfossiliferous strata with a 6-in. magnus Band at the base. Similarly the leptotheca Zone was given as 20 ft. of relatively unfossiliferous beds with a 2|-ft. leptotheca Band at the base. The practice of erecting zones based on thin graptolite bands separated by barren strata is not considered correct, particularly as one has no idea of the fauna of the intervening beds. Although the horizons with M. leptotheca and Diplograptus magnus at Dobb’s Linn provide direct correlation with the sequence in Central Wales, they should only be regarded as bands and not zones. Dr. R. B. Rickards records (in lift .) that the lithologies and faunas at the level of the gregarius Zone in the Howgill Fells (eastern Lake District) are very similar to Dobb’s Linn. He recognizes the three zones triangulatus, magnus, and argenteus ( leptotheca ), and reports that there is a marked lithological change from black to grey mudstones at the level of the magnus Zone, a state of affairs comparable with the sequence at Dobb’s Linn 664 PALAEONTOLOGY, VOLUME 11 table 2. A correlation of Llandovery (Valentian) graptolite zonal schemes, up to the base of the M. turriculatus Zone. TOGHILL: GRAPTOLITE ASSEMBLAGES AND ZONES OF BIRKHILL SHALES 665 where the first grey mudstones appear immediately above the 3-in. band with Diplo- graptus magnus. The M. convolutus Zone as used here is equivalent to the same zone as defined by Elies and Wood (1913), and Elies (1925), the cometa Band being included in the zone. In some cases, in Wales, the cometa Band has been treated as a separate zone (Jones, 1909, 1929; Curtis, 1961). The zone is equivalent to the cometa Zone of Lapworth (1878, pp. 322-3) together with the clingani Band of his sedgwickii Zone. The base of the M. sedgwickii Zone as defined here is equivalent to the same horizon in Wales, being taken at the appearance of M. sedgwickii. The clingani Band of Lapworth’s sedgwickii Zone has been relegated to the convolutus Zone. The base of the Rastrites maximus Zone has been taken at a similar level to Lapworth (1878, pp. 322, 325-7), but the two species which according to Lapworth characterize the zone at Dobb's Linn, R. maximus and M. hal/i, have not been found to be at all common in the zone, but occur abundantly in localities south-east of Dobb’s Linn (see below). The zone is considered to be equivalent to the halli Zone in Wales (Jones and Pugh 1916, W. D. V. Jones 1945). Some authors have suggested in the correlation charts of review papers (Jones 1921, 1929; Curtis 1961) that a maximus Zone has been defined in Wales above the halli Zone. In fact this has never been done, although within the halli Zone around Machynlleth ( Jones and Pugh 1916) there is a separate Rastrites Band above a halli Band. The only use of a R. maximus Zone in Wales was by Elies (1909) in the Conway area where it lay between the sedgwickii and crispus Zones. Localities south-east of the Caledonoid strike-line through Dobb’s Linn (Craigmichan Scars, Thirlstane Scar, and Beldcraig Burn) expose fossiliferous horizons which are higher in the maximus Zone than any fossiliferous strata at Dobb’s Linn. Lapworth com- pared these sections directly with the maximus Zone at Dobb’s Linn (1878, pp. 326-7) as if they were exactly the same horizon. He was unaware of the diachronous nature of the Gala Greywackes south-east of Dobb’s Linn which meant that their base had ‘stepped up’ in relation to Dobb’s Linn. He was, however, quite aware that the Gala Greywackes ‘stepped down’ north-west of Dobb’s Linn. These localities south-east of Dobb’s Linn yield Rastrites maximus and Monograptus halli in abundance with M. runcinatus, M. spiralis and rarely M. turriculatus minor, and further substantiate the suggestion that the halli and maximus Zones are equivalent. The latter name should take priority on historical grounds. Higher horizons in the Gala Greywackes south-east of Dobb’s Linn yield a fauna which is typical of the Monograptus turriculatus Zone, including, M. exiguus, M. barrandei, M. nodifer, and the zone fossil. THE TERMS LOWER, MIDDLE, AND UPPER LLANDOVERY APPLIED TO THE BIRKHILL SHALES The terms Llandovery (Murchison 1859) and Valentian (Lapworth 1876) are now considered to be synonymous, and the former term has priority. However, the Valentian is usually considered to be the graptolitic Llandovery, and as the divisions of the Llan- dovery are based on a shelly fauna and unconformities in the type area (Jones 1925), it is unlikely that the divisions can be readily applied to the graptolitic sequence. When the Valentian was further defined by Jones (1921) the base was taken at the base of the G. persculptus Zone (1921, pp. 173-4), this zone having been originally 666 PALAEONTOLOGY, VOLUME 11 defined in the graptolitic sequence of Central Wales (Jones 1909, p. 482). This level is now accepted as the base of the Llandovery (Curtis 1961, p. 129). The base of the Middle Llandovery is usually equated to the base of the magnus Zone in the graptolitic sequence of Central Wales (Curtis 1961, p. 130), but this is entirely based on lithological considera- tions (Jones, 1925, p. 366; 1947, p. 3; 1954, p. 252; Jones and Pugh 1935, pp. 274-5). The only variation from this opinion was by Jones in 1929 when he included the magnus Zone in the Lower Llandovery (1929, p. 97). The highest graptolites recorded from the Lower Llandovery at Llandovery are Monograptus incommodus and Climacograptus hughesi from the top of A4 (Jones 1925, p. 360; 1929, p. 94), and these were considered to indicate thetop of the acinaces Zone (Jones 1925, p. 360). The Middle Llandovery at Llan- dovery has yielded Monograptus decipiens, M. cf. lobiferus, and M. cf. regularis. These were collected from near to the top of the division and considered by Jones (1925, p. 366) to indicate the convolutus Zone. The abundance of M. sedgwickii at 1 50 ft. above the base of the Upper Llandovery at Llandovery (Jones 1925, p. 370) suggests that the present accepted level for the base of the Upper Llandovery in the graptolitic sequence, i.e. the base of the sedgwickii Zone, is quite acceptable. However, there is no such fixed level for the base of the Middle Llandovery in the graptolitic sequence which could, on the evidence stated above, lie as low as the base of the cyphus Zone, or even as high as the base of the convolutus Zone (Jones 1925, p. 366). This is an unsatisfactory situation, and the problem cannot really be solved until more graptolite evidence is obtained from the Llandovery area, or shelly evidence from the graptolitic sequence. Therefore as the type area cannot offer any faunal evidence which will fix the Lower-Middle Llandovery boundary with any certainty in the graptolitic sequence, there can be no harm in con- sidering the graptolite fauna of the Llandovery (Valentian) as a whole, in order to suggest an equivalent horizon in the graptolitic sequence as a working base for the Middle Llandovery. Within the graptolite fauna of the Birkhill Shales are four distinct divisions cor- responding to those erected by Elies (1922) and further elaborated by Bulman (1958). The lowest part of the Birkhill Shales below the incoming of monograptids ( persculptus - acuminatus Zones) is equivalent to the Ortho-Climacograptid sub-fauna of Bulman (1958, pp. 170-1), the highest division of Elles’s Diplograptid Fauna. The next division of the Birkhill Shales, above the incoming of monograptids, but below the appearance of triangulate monograptids, vesiculosus-cyphus Zones, ( atavus-cyphus Zones in Wales) represents the lowest division, A, of Bulman’s Monograptid Fauna. The next division, gregarius- convolutus Zones, (triangulatus- convolutus Zones in Wales), represents the second division, B, of Bulman’s Monograptid Fauna, and the final division, sedgwickii- maximus Zones represents part of the third division, C, of the Monograptid Fauna (Bulman 1958, pp. 170-1). The base of the Llandovery is well fixed in the graptolitic sequence at the base of the persculptus Zone, which was defined for the first time in Scotland by Toghill (1968). It is worth noting that Elies (1922, p. 195) suggested that the lowest two zones of the Llandovery ( persculptus-acuminatus Zones) being devoid of monograptids could well be relegated to the Ordovician on faunistic grounds, but this has never been accepted. The base of the Middle Llandovery if taken at the base of the magnus Zone is some way above the appearance of triangulate monograptids, and some way above the base of the gregarius Zone as defined here. The appearance of triangulate monograptids at the base TOGHILL: GRAPTOLITE ASSEMBLAGES AND ZONES OF BIRKHILL SHALES 667 of the gregarius Zone ( triangulatus Zone in Wales) corresponds to the base of the second division, B, of the Monograptid Fauna. It would be more convenient if the base of the Middle Llandovery in the graptolitic sequence were to be taken at the base of the triangulatus Zone in Wales, and thus equate with the base of the gregarius Zone at Dobb’s Linn. If acceptable the base of the Middle Llandovery would then correspond to the base of the second division of Bulman’s Monograptid Fauna (Bulman 1958, pp. 170-1). The base of the Upper Llandovery taken at the base of the sedgwickii Zone is a convenient horizon as it is the base of the third division of the Monograptid Fauna. Thus the stage boundaries of the Llandovery in the graptolitic sequence would now correspond to significant faunal horizons, though within the Lower Llandovery is the most conspicuous horizon of all, the appearance of the monograptids. In conclusion, the Lower Llandovery at Dobb’s Linn is represented by the zones of Glyptograptus persculptus, Akidograptus acuminatus, Cystograptus vesiculosus, and Mono- graptus cyphus, and is 48 ft. (14-6 m.) thick. The Middle Llandovery includes the zones of M. gregarius and M. convolutus and is 44 ft. (13-4 m.) thick. Only part of the Upper Llandovery is present, represented by the zones of M. sedgwickii and Rastrites maximus. These two zones are 49 ft. (15 m.) thick and pass up into the great thickness of the Gala Greywackes which (together with the Hawick Rocks?) represent the remainder of the Upper Llandovery. Acknowledgements. I am extremely grateful to Dr. I. Strachan for his constant advice throughout the research, a grant for which was obtained from the Department of Scientific and Industrial Research (now N.E.R.C.). I would also like to thank Professor O. M. B. Bulman for the loan of supplementary specimens from the Sedgwick Museum. REFERENCES bulman, o. m. b. 1958. The sequence of graptolite faunas. Palaeontology , 1, 159-73. curtis, m. l. K. 1961. In Lexique Stratigraphique International, 1, Europe, (3aV), Silurian. davies, K. a. 1926. The geology of the country between Drygarn and Abergwesyn (Breconshire). Q. Jl geol. Soc. Lond. 82, 436-64. 1929. Notes on the graptolite faunas of the Upper Ordovician and Lower Silurian. Geol. Mag. 66, 1-27. elles, G. l. 1909. The relations of the Ordovician and Silurian rocks of Conway (North Wales). Q. Jl geol. Soc. Lond. 65, 169-94. — — 1922. The graptolite faunas of the British Isles. A study in evolution. Proc. Geol. Ass. Lond. 33, 168-200. 1925. The characteristic assemblages of the graptolite zones of the British Isles. Geol. Mag. 62, 337-47. and wood, e. m. r. 1901-18. A monograph of British graptolites. Pa/aeontogr. Soc. [Monogr.]. jones, o. t. 1909. The Hartfell-Valentian succession in the district around Plynlimon and Pont Erwyd (North Cardiganshire). Q. Jl geol. Soc. Lond. 65, 463-537. 1921. The Valentian series. Ibid. 77, 144-74. 1925. The geology of the Llandovery district: Part I — The southern area. Ibid. 81, 344-88. 1929. Silurian. In Handbook of the geology of Great Britain, j. w. evans and c. J. Stubblefield (eds.), London, 88-127. 1933. The lower Palaeozoic rocks of Britain. Int. geol. Congr. 16th session, Washington, 1, 463-84. 1947. The Llandoverian graptolite succession in Britain. Ball. Mus. r. Hist. nat. Belg. 23, (22), 1-4. 1954. The use of graptolites in geological mapping. Lpool. Manchr. geol. J. 1, 246-60. and pugh, w. j. 1916. The geology of the district around Machynlleth and the Llyfnant Valley. Q. Jl geol. Soc. Lond. 71, 343-85. 668 PALAEONTOLOGY, VOLUME 11 jones, o. T. and pugh, w. j. 1935. The geology of the districts around Machynlleth and Aberystwyth. Proc. Geol. Ass. Lond. 46, 247-300. jones, w. d. v. 1945. The Valentian succession around Llanidloes, Montgomeryshire. Q. Jl geol. Soc. Lond. 100, 309-32. and rickards, r. b. 1967. Diplograptus penna Hopkinson 1869, and its bearing on vesicular structures. Palaont. Z. 41, 173-85. lapworth, C. 1876. The Silurian system in the south of Scotland. In Catalogue of the Western Scottish fossils, j. Armstrong et al. (ed.), Glasgow, 1-28. 1878. The Moffat Series. Q. Jl geol. Soc. Lond. 34, 240-346. 1880. On the Geological Distribution of the Rhabdophora. Ann. Mag. nat. Hist. (5), 5, 45-62, 358-69; 6, 16-29, 185-207. lapworth, h. 1900. The Silurian sequence of Rhayader. Q. Jl geol. Soc. Lond. 56, 67-137. marr, j. e. and nicholson, h. a. 1888. The Stockdale Shales. Ibid. 44, 654-732. murchison, r. i. 1859. Siluria. 2nd ed. London. sudbury, m. 1958. Triangulate Monograptids from the Monograptns gregarius Zone (Lower Llan- dovery) of the Rheidol Gorge (Cardiganshire). Phil. Trans. R. Soc. B. 241, No. 685, 485-555. toghill, p. 1968. The stratigraphical relationships of the earliest Monograptidae, and the Dimor- phograptidae. Geol. Mag. 105, 46-51. walton, e. k. 1963. Sedimentation and structure in the Southern Uplands. In The British Caledonides, m. r. w. Johnson and f. h. stewart (eds.), Edinburgh, 71-97. 1965. Lower Palaeozoic rocks— stratigraphy, palaeogeography and structure. In The geology of Scotland, g. y. craig (ed.), Edinburgh, 161-227. P. TOGHILL Department of Palaeontology, British Museum (Nat. Hist), Cromwell Road, Typescript received 5 February 1968 London S.W. 7 NUMMULITES (FOR AMINIFE R A) FROM THE UPPER EOCENE KOPILI FORMATION OF ASSAM, INDIA by B I M A L K. SAMANTA Abstract. Four species of Nummulites are described and illustrated front the Kopili Formation, Garo Hills, Assam, India. This is the first account of the genus Nummulites from the Upper Eocene Pellatispira- bearing horizon in the Indian region. Outcrops of marine Upper Eocene rocks with larger foraminifera are known to occur in three areas in the India-Pakistan region (Samanta 1968, fig. 1): Surat-Broach in Western India (Rao 1941), the Sulaiman Range in West Pakistan (Eames 1952) and Assam in Eastern India (Nagappa 1951, Samanta 1965). Nummulites has been reported to occur in association with the typical Upper Eocene genus Pellatispira in all three areas but so far there is no published account of the genus from this horizon. In Assam the Kopili Formation contains a rich Upper Eocene larger foraminiferal assemblage including such stratigraphically important genera as Asterocyclina, Dis- coeyclina, Nummulites , and Pellatispira. An investigation of the larger foraminifera of the Kopili Formation in the Garo Hills has been carried out by the writer and an account of the genus Nummulites is given in the present paper. KOPILI FORMATION Evans (1932, pp. 173—5) first called this unit the Kopili alternations ‘Stage’ and some- times Kopili ‘Stage’. Later workers have changed the name to Kopili Formation, since by original designation it is basically a rock unit. In the type section (Kopili River section of the Kopili-Khorungma region) the succession is reported to be about 450 m. thick and consists of alternations of sandstone, mudstone, shales, carbonaceous rocks, and shell-bearing sandstone. It conformably overlies the Sylhet Limestone and is apparently conformably overlain by the Barail group of rocks. The formation outcrops along the southern fringe of the Shillong Plateau, from the Garo Hills in the west to the Mikir Hills in the east. In the Garo Hills the Kopili Formation is best exposed in the Simsang River section between Siju Artheka (90° 41' E., 25° 20' N.) and Matmagitik (90° 40' E., 25° 18' N.). It conformably overlies the Siju Limestone and is apparently conformably overlain by Barail-equivalent rocks (Samanta 1968, p. 128, table 1). The lower part of the formation is richly fossiliferous and contains abundant larger foraminifera, including such strati- graphically important genera as Asterocyclina , Discocyclina, Nummulites , and Pellati- spira. Of these, Discocyclina is the most abundant. Because of their much larger size in comparison to other larger foraminifera, discocyclines constitute the most conspicuous element of the foraminiferal fauna. Nummulites is represented by small to medium sized striate and reticulate forms and occurs in almost all foraminiferal samples. In contrast [Palaeontology, Vol. 11, Part 5, 1968, pp. 669-82, pis. 128-9] 670 PALAEONTOLOGY, VOLUME 11 to Discocyclina and Nummulites, Asterocyclina and Pellatispira occur in fewer samples and are much less abundant in numbers of individuals. The following larger foraminifera are identified from the Kopili Formation, Garo Hills (see also Samanta 1968, p. 129, table 2): Asterocyclina matanzensis Cole Discocyclina archiaci (Schlumberger) D. assamica Samanta D. august ae Weijden D. dispansa (Sowerby) I). eamesi Samanta D. javana (Verbeek) D. omphalus (Fritsch) D. pygmaea Henrici D. sella (d’Archiac) D. sowerbyi Nuttall D. sp. cf. D. trabayensis Neumann Nummulites chavannesi de la Harpe N. sp. aft. N. chavannesi de la Harpe N. fabianii (Prever) N. pengaronensis Verbeek Pellatispira inflata Umbgrove P. sp. cf. P. irregularis Umbgrove P. madaraszi (Hantken) P. sp. cf. P. orbitoidea (Provale) Of these, D. augustae, D. sella, N. chavannesi, N. fabianii, and P. madaraszi are recorded from the Priabonian of North Italy, while A. matanzensis, D. javana, D. omphalus, D. pygmaea , D. sella, N. pengaronensis, and the four species of Pellatispira are abundantly represented in the T b of the Indonesian region. The larger foraminiferal assemblage, therefore, indicates a definite Upper Eocene age for the lower part of the Kopili Formation. Material. The material was collected from five localities in the Garo Hills, previously described (Samanta 1965, p. 416, text-fig. 3). All the four species of Nummulites are represented by sufficient material. Presence of free specimens permits a detailed study of these forms. Table 1 shows the distribution of the species in the Garo Hills. table 1. Distribution of Nummulites in the Kopili Formation, Garo Hills, Assam Localities Species Sa Rn Rgt N K Nummulites chavannesi X de la Harpe N. sp. atf. N. chavannesi X X X de la Harpe N. fabianii (Prever) X X X N. pengaronensis Verbeek X X X X X Acknowledgements. The author is indebted to Dr. J. R. Haynes for critically reading the manuscript; Dr. F. E. Eames for helpful discussions; Professors H. Hagn and E. Montanaro Gallitelli for com- parative material; Drs. F. Bieda and V. Roveda for literature; Professor Alan Wood for providing facilities in his Department; and Mr. H. Williams for help in preparing the plates. SYSTEMATIC PALAEONTOLOGY Family nummulitidae de Blainville 1825 Subfamily nummulitinae de Blainville 1825 Genus nummulites Lamarck 1801 Nummulites chavannesi de la Harpe Plate 128, figs. 11, 12; Plate 129, figs. 9-14; text-fig. 1 1877 Nummulites chavannesi de la Harpe, p. 232 ( nom . nud.). 1883fl Nummulites bouillei var. riitimeyeri de le Harpe, pi. 6. figs. 5-11. 1883fl Nummulites chavannesi de la Harpe, pi. 6. figs. 22-41. 18836 Nummulites riitimeyeri de la Harpe, pp. 162, 163, pi. 30, figs. 9-11. SAMANTA: NUMMULITES, KOPILI FORMATION, ASSAM 671 1883 b Nummulites chavannesi de la Harpe; de la Harpe, pp. 163, 164, pi. 30, figs. 12-18. 19116 Nummulites chavannesi de la Harpe; Boussac, pp. 37, 38. 1934 Nummulites cf. chavannesi de la Harpe; Flandrin, pp. 254, 255, pi. 14, figs. 15, 16. 1934 Nummulites rutimeyeri de la Harpe; Flandrin, p. 254, pi. 14, fig. 17. 1938 Nummulites rutimeyeri de la Harpe; Flandrin, pp. 34, 35, pi. 3, figs. 9, 10. 1951 Nummulites rutimeyeri de la Harpe; Daci, pp. 209, 210, pi. 2, figs. 7, 8. 1951 Nummulites chavannesi de la Harpe; Daci, pp. 210, 211, pi. 2, fig. 9. 1957 Nummulites chavannesi de la Harpe; Bieda, pp. 46, 47, pi. 4, figs. 8, 9. 1960 Nummulites chavannesi de la Harpe; Hagn, p. 70, pi. 1, fig. 2; pi. 2, figs. 4, 5. 1961 Nummulites chavannesi de la Harpe; Roveda, pp. 177-81, pi. 14, figs. 1-8. 1963n Nummulites chavannesi de la Harpe; Bieda, pp. 71, 72, 186, pi. 6, figs. 5-7 ; pi. 7, figs. 1-3. text-fig. 1. Nummulites chavannesi de la Harpe. a, Part of the equatorial section of a microspheric specimen. x20 approx, b, Equatorial section of a megalospheric specimen, X 25 approx, c, Axial section of a megalospheric specimen, X 25 approx. All from locality K. Material. Megalospheric form — 25 specimens examined externally, 5 specimens studied in equatorial section, and 5 in axial section. Microspheric form — 6 specimens examined externally, 3 specimens studied in equatorial section, and 2 in axial section. Description. Megalospheric form. Test small, lenticular, with slightly elevated polar region surrounded by sloping peripheral part; margin acute. Surface ornamented with well-developed polar pustules from which thin, straight to gently curved septal filaments radiate. Diameter of test varies from T9 to 3-4 mm., thickness from 0-9 to T4 mm., ratio of diameter to thickness from 2-2 to 2-7, and diameter of polar pustules from 0-4 to 0-6 mm. About 4\ to 6 regularly coiled whorls open rapidly. Spiral lamina thin and in outer 672 PALAEONTOLOGY, VOLUME 11 whorls height of spiral cavity about 4 to 6 times thickness of spiral lamina. Septa nearly perpendicular to spiral lamina, straight with sharp curvature near distal end. About 8-11 septa occur in 1st whorl; 15-20 in 2nd; 19-28 in 3rd; 24-9 in 4th; and 28-32 in 5th. Small, subcircular first chamber followed by subequal, reniform second chamber. Separating wall either straight or curved outwards. Diameters of first chamber vary from 0-055 x 0-050 mm. to 0-095x0-075 mm. and those of second chamber from 0-055 x 0-045 mm. to 0-100x0-060 mm. Distance across both chambers varies from 0-105 to 0-180 mm. Equatorial chambers quadrate in shape and about twice as high as long. In axial section first chamber circular and about 0-05 mm. in height. Chamber cavity triangular in shape. Alar prolongations wide open. Marginal cord distinct. Wedge- shaped polar plugs always very conspicuous and about 0-6 mm. in diameter near surface. Microspheric form. Test small with well-developed, slightly elevated polar pustules; margin acute. Septal filaments thin, radiate, nearly straight. Diameter of test varies from 3-8 to 5-0 mm., thickness from 1-8 to 2-0 mm., ratio of diameter to thickness from 2-1 to 2-4 mm., and diameter of polar pustules from 0-8 to TO mm. There are about 9 whorls in diameter of 4-2 mm. Whorls regularly coiled and open rather rapidly. In outer whorls height of spiral cavity about 3 times thickness of spiral lamina. Septa nearly perpendicular and straight with sharp curvature at top. Chambers quadrate, about 2 to 3 times higher than long. In axial sections alar prolongations wide open. Marginal cord distinct. Well-developed polar plugs wedge-shaped, about TO mm. in diameter near surface. Remarks. The presence of well-defined polar pustules, high equatorial chambers between almost straight septa, and wide alar prolongations distinguish this species from the associated nummulites in the Kopili Formation. The Assam specimens have been com- pared with European material, provided by Professor H. Hagn. They are closely similar to the Priabonian material described and illustrated by Roveda (1961). Distribution. N. chavannesi has been reported from Italy, Spain, France, Switzerland, Poland, Hungary, Albania, Turkey, Algeria, Egypt, and Somaliland. Its known range is from Upper Eocene to Oligocene. In the Garo Hills N. chavannesi occurs only in the Upper Eocene Kopili Formation (Table 1). There is no report of its occurrence in the other Upper Eocene localities in India and adjacent countries. The present record of N. chavannesi extends its geographic distribution considerably. EXPLANATION OF PLATE 128 Figs. 1-10. Nummulites pengaronensisNe rbeek. 1, External view of microspheric specimen, x6. 2, 3, External views of megalospheric specimens; 2, inflated variety, x6; 3, compressed lenticular variety, x9. 4, 5, Equatorial sections of megalospheric specimens, x 15; 4, inflated variety; 5, len- ticular variety. 6-8, Axial sections of megalospheric specimens showing variation in transverse views of tests ; 6, X 1 5 ; 7, 8, X 2 1 . 9, Axial section of microspheric specimen, X 9. 10, Equatorial section of microspheric specimen, x9. 1, 3, 8, 10, from locality Sa; 2, 7, from locality K; 4, 5, from locality Rn; 6, 9, from locality N (see Samanta 1965, p. 416). Figs. 11, 12. Nummulites chavannesi de la Harpe. 11, Axial section of megalospheric specimen, x21. 12, Equatorial section of megalospheric specimen, x21. Both from locality K. Figs. 13-15. Nummulites sp. aff. N. chavannesi de la Harpe. 13, External view of megalo-spheric specimen, X 15. 14, Equatorial section of megalospheric specimen, x21. 15, Axial section of megalo- spheric specimen, x 21. All from locality K. Palaeontology, Vol. 11 PLATE 128 SAMANTA, Upper Eocene Nummulites SAMANTA: NUMMULITES, KOPILI FORMATION, ASSAM 673 Nummulites sp. aft'. N. chavannesi de la Harpe Plate 128, figs. 13-15; text-fig. 2 Material. Only megalospheric specimens were observed. 1 5 specimens examined externally, 5 specimens studied in equatorial section, and 2 in axial section. Description. Megalospheric form. Test very small, lenticular, with acute margin. Septal filaments thin and radial. Diameter of test varies from T6 to 2-0 mm., thickness from 0-56 to 0-80 mm., and ratio of diameter to thickness from 2-0 to 3-4. Spire is regular with about 44 to 54 whorls increasing regularly in height. Spiral lamina thin, about } height of spiral cavity in thickness. Septa are perpendicular to wall and text-fig. 2. Nummulites sp. aft'. N. chavannesi de la Harpe (megalospheric form.) A, Equatorial section, b. Axial section. Both from locality K, x25 approx. nearly straight with curvature at top. About 8-10 septa occur in 1st whorl; 15-18 in 2nd; 19-25 in 3rd; 24-6 in 4th; and 24-30 in 5th. Subcircular first chamber followed by subequal, crescentic second chamber. Separat- ing wall gently curved outwards. Diameters of first chamber vary from 0-05 x 0 03 mm. to 0-08. x 0-07 mm.; those of second chamber from 0-055 x 0-035 mm. to 0-10x0-055 mm. Distance across both chambers varies from 0-09 to 0-15 mm. Equatorial chambers quadrate and higher than long. In axial section first chamber circular, about 0-06 mm. in height. Alar prolongations wide open. Marginal cord distinct. Remarks. The present specimens are closely similar to N. chavannesi in internal charac- ters, but can easily be distinguished from the latter by the shape of the test and the absence of polar pustules. They are provisionally identified as N. sp. aff. N. chavannesi. Distribution, The species occur in localities Rn, N, and K in the Garo Hills (Samanta 1965, p. 416). The presence of similar forms has also been noticed by the writer in the Upper Eocene of Surat- Broach, Western India. Nummulites fabianii (Prever) Plate 129, figs. 1-8; text-fig. 3, 4 1905 Bruguieria fabianii Prever in title, Fabiani, pp. 1805, 1811, 1824. 1905 Bruguieria sub-fabianii Prever in lift.; Fabiani, pp. 1811, 1824. 674 PALAEONTOLOGY, VOLUME 11 1906 Nummulites fabianii (Prever in Fabiani); Boussac, pp. 88-90, pi. 1, figs. 1-5, 7-9; pi. 3, fig. 6. 191 la Nummulites fabianii Prever in Fabiani; Boussac, pp. 40, 72, pi. 10, figs. 1, 2, 28; pi. 17, figs. 8, 11, 13. 19116 Nummulites fabianii Prever in Fabiani; Boussac, pp. 79-84, pi. 1, figs. 6, 13; pi. 4, figs. 9, 10. 1928 Nummulites fabianii Prever; de Cizancourt, p. 294, pi. 2, fig. 10. 1930 Nummulites fabianii Prever; de Cizancourt, pp. 209, 210, pi. 22, figs. 4, 7; pi. 23, fig. 5. 1934 Nummulites fabianii Prever; Flandrin, p. 259, pi. 1, fig. 20. 1951 Nummulites fabianii Prever; Daci, pp. 221, 222, pi. 3, figs. 1, 2. 1951 Nummulites subfabianii Prever; Daci, pp. 222-4, pi. 3, figs. 4-7. 1957 Nummulites fabianii Prever; Bieda, p. 30, pi. 5, fig. 5. 1959 Nummulites retiatus Roveda, pp. 201-7, pi. 1, figs. 1-11. 1960 Nummulites fabianii (Prever); Hagn, p. 149, pi. 2, figs. 2, 3, 7. 1961 Nummulites fabianii (Prever)', Roveda, pp. 161-9, pi. 15, figs. 15, 16; pi. 17, figs. 8, 9; pi. 18, figs. 4, 5; pi. 19, figs. 1, 6-8, 14-16. 1963a Nummulites fabianii Prever; Bieda, pp. 101-4, 195, 196, pi. 15, fig. 9; pi. 16, figs. 1-4. 19636 Nummulites fabianii Prever; Bieda, pp. 201-4, 214-15, pi. 13, figs. 3, 4. 1965 Nummulites fabianii Prever; Bozorgnia and Kalantari, pp. 17, 18; pi. 20, figs. 1-7. Material. Megalospheric form — 20 specimens examined externally, 5 studied in equatorial section, and 7 in axial section. Microspheric form — 10 specimens examined externally, 4 studied in equatorial section, and 1 in axial section. Description. Megalospheric form. Test small, lenticular, with subacute margin. Surface ornamented with spirally arranged rectangular meshes produced by intersections of radial filaments with raised spiral line. Spirally arranged granules joined together by ‘transverse lamina’ produce raised spiral line. In some specimens granules cluster at poles to form polar pustules. Diameter of test varies from 1-8 to 3-0 mm., thickness from 1-25 to 1-85 mm., and ratio of diameter to thickness from 1-4 to 1-9. About 5 to regularly coiled whorls occur, increasing slowly in height. Spiral lamina thick, and in some inner whorls may be as thick as height of spiral cavity. Near periphery height of spiral cavity about 2 to 3 times thickness of spiral lamina. Septa slightly inclined to spiral wall, straight to gently curved in their course. About 6-7 septa occur in 1st whorl; 9-13 in 2nd; 12-16 in 3rd; 15-20 in 4th; and 16-22 in 5th. Subcircular first chamber followed by smaller, semicircular to reniform second cham- ber. Separating wall either straight or curved outwards. Diameters of first chamber vary from 0-130x0-095 mm. to 0-20x0-20 mm.; those of second chamber from 0*10x0-05 mm. to 0-175 X 0-095 mm. Distance across both chambers varies from 0-16 to 0-28 mm. Chambers quadrate in shape. Near centre, chambers almost as long as high, but in ontogeny chambers become considerably longer, so that in outer whorls chambers become twice as long as high. EXPLANATION OF PLATE 129 Figs. 1-8. Nummulites fabianii (Prever). 1, External view of microspheric specimen, X 6. 2, 3, Equatorial sections of microspheric specimens, x9. 4, External view of megalospheric specimen, Xl5. 5, 6, Axial sections of megalospheric specimens, x21. 7, Equatorial section of megalospheric specimen, x 15. 8, Axial section of microspheric specimen, x7-5. All from locality Rn (see Samanta 1965, p. 416). Figs. 9-14. Nummulites chavannesi de la Harpe. 9, External view of microspheric specimen, x6. 10, 11, External views of megalospheric specimens, x 12. 12, Axial section of megalospheric speci- men, x21. 13, Equatorial section of megalospheric specimen, x21. 14, Equatorial section of microspheric specimen, X 15. All from locality K. Palaeontology, Vol. 11 PLATE 129 SAMANTA, Upper Eocene Nummulites 675 SAMANTA: NUMMULITES , KOPILI FORMATION, ASSAM text-fig. 3. Nummulites fabianii (Prever) (megalospheric form). A, Equatorial section, b. Axial section. Both from locality Rn, x25 approx. text-fig. 4. Nummulites fabianii (Prever). Part of the equatorial section of a microspheric specimen from locality Rn, x 18 approx. In axial section first chamber circular, about 0-10 to 0-13 mm. in height. Spiral lamina rather thick. There may be reduction in thickness of spiral lamina at periphery. Alar prolongations narrow to moderately open. Marginal cord distinct. Pillars well-developed, start from marginal cord of each whorl, and of uniform thickness throughout length. At poles, pillars cluster together to form polar plugs. Diameter of pillars varies from 0-050 to 0-075 mm. and polar plugs from 0-25 to 0-50 mm. Microspheric form. Test medium-sized, lenticular, with subacute margin. Surface of vy C 6055 676 PALAEONTOLOGY, VOLUME 11 test ornamented with thin, reticulate septal filaments. In young individuals rectangular meshes are discernible but in adult specimens branching filaments produce complex network. Diameter of test varies from 5-1 to 8-4 mm., thickness from 2-7 to 4-2 mm., and ratio of diameter to thickness from 1 -9 to 2-2. In equatorial section about 9 to 13 whorls occur, coiled regularly and increasing slowly in height during ontogeny. Spiral lamina rather thick. In adult whorls height of spiral cavity usually greater than thickness of spiral lamina. Septa inclined to whorl wall, and straight to gently curved in their course. Equatorial chambers longer than high, and in outer whorls 3 to 4 times as long as high. In axial section, alar prolongations narrow to moderately open. Marginal cord dis- tinct. Pillars moderately developed. Each pillar starts from marginal cord and extends up to surface. Diameter of pillars varies from 0-05 to 0-15 mm. In polar region pillars cluster together to form polar plug-like structures about 0-8 mm. in diameter near surface. Remarks. Both in external and internal features the present form is distinctive. The reticulate ornamentation, the long equatorial chambers, and the pillared axial section enables the species to be distinguished from the associated nummulites in the Kopili Formation. The Assam specimens were compared with those of N. fabianii from North Italy provided by Professor Montanaro-Gallitelli. Because of their distinctive morphological features and wide geographic distribution in the rocks of Upper Eocene to Oligocene age, the reticulate Nummulites have received particular attention and several species have been described. But at present there is considerable difference of opinion about the validity of a number of these forms (see Eames et al. 1959; Bieda 196 3b); consequently, application of reticulate Nummulites species in the finer biostratigraphic zonation of Upper Eocene-Oligocene rocks is lacking. Distribution . N. fabianii is one of the most widely distributed representatives of the genus, reported from the Upper Eocene of Italy, Spain, France, Switzerland, Poland, Hungary, Albania, Rhodes Island, Turkey, Morocco, Algeria, Tunisia, Libya, Egypt, and Iran. In the Garo Hills N. fabianii occurs in the Kopili Formation at localities Sa, Rn, and K (Table 1). It occurs also in two other Upper Eocene localities in the Indian region; in the Sulaiman Range its presence has been noted by Bayliss (1961), while the writer has observed it in Surat-Broach in associa- tion with Pellatispira spp., etc. There is no authentic record of reticulate Nummulites from the Upper Eocene of the Malayan Archipelago (Cole 1963, Adams 1965). The only report of an occurrence in association with a typical Upper Eocene assemblage from this region was that by Cole (Cloud and Cole 1953, p. 323) who later (1963, pp. E4, E14) postulated that the Upper Eocene species in the assemblage are reworked specimens and that reticulate Nummulites do not occur in the Eocene of the Malayan Archipelago. Thus, Assam is the easternmost locality with N. fabianii. Nummulites pengaronensis Verbeek Plate 128, figs. 1-10; text-figs. 5, 6 1871 Nummulites pengaronensis Verbeek, pp. 3-6, pi. 1, figs. 1 a-k. 1892 Nummulites nanggoelani Verbeek, pp. 116, 118. 1896 Nummulites nanggoelani Verbeek; Verbeek and Fennema, p. 1152, pi. 8, figs. 111-13. 1896 Nummulites pengaronensis Verbeek; Verbeek and Fennema, pp. 1153, 1 154. 1912 Nummulites pengaronensis Verbeek; Douville, pp. 284, 285, pi. 24, fig. 6. SAMANTA: NUMMULITES, KOPILI FORMATION, ASSAM 677 1921 Nummulites cf. pengaronensis Verbeek; Yabe, pp. 104, 105, pi. 18, fig. 8. 1929 Nummulites pengaronensis Verbeek; Vlerk, pp. 20, 21, figs. 12, 35a, b. 1932 Camerina pengaronensis (Verbeek); Doornink, pp. 283, 284, pi. 4, figs. 1-3; pi. 6, fig. 12. 1934 Camerina pengaronensis (Verbeek); Henrici, pp. 29, 30, pi. 1, fig. 10. 1934 Camerina cf. pengaronensis (Verbeek); Caudri, p. 52. 1953 Camerina saipanensis Cole, pp. 20, 21, pi. 2, figs. 7-19. 1957 Camerina pengaronensis (Verbeek); Cole, pp. 753, 754, pi. 231, figs. 1-17. 1959a Nummulites pengaronensis Verbeek; Nagappa, pp. 163, 166, pi. 10, figs. 3-5. 1965 Nummulites cf. saipanensis (Cole); Adams, p. 313, pi. 23, fig. c. text-fig. 5. Nummulites pengaronensis Verbeek (megalospheric form), a-c, Axial sections, x25 approx.; a, from locality Sa; b, from locality K; c, from locality N. d, Split specimen from locality Sa, x 20 approx, e, Equatorial section from locality Rn, x 25 approx. 678 PALAEONTOLOGY, VOLUME 11 Material. Megalospheric form — 75 specimens examined externally, 5 studied in equatorial section, and 12 in axial section. Microspheric form — 12 specimens examined externally, 3 studied in equatorial section, and 4 in axial section. In addition, 23 split specimens of megalospheric form and 3 of micro- spheric form were studied in equatorial view. Description. Megalospheric form. Test small, compressed, lenticular to globose, with acute margin. Surface marked by thin, radial septal filaments, straight to gently curved at ends. Diameter of test varies from 2-0 to 5-2 mm., thickness from 0-8 to 3-3 mm., and ratio of diameter to thickness from 1-3 to 3-6. Spire more or less regular with about 5 to 7 whorls increasing regularly in height, with exception of last whorl, which may be narrower than preceding one. Height of spiral cavity in outer whorls about 3 times thickness of spiral lamina. Septa nearly perpendicular to inclined at their base, straight for about half their course, then curve sharply backwards. Thickness of septa decreases considerably from proximal to distal end. About 6-7 septa occur in 1st whorl; 1 1-14 in 2nd; 17-19 in 3rd; 23-6 in 4th; 24-7 in 5th; 26-9 in 6th; and 29-32 in 7th. First chamber circular to elliptical in equa- torial section, followed by usually smaller, cres- centic to reniform second chamber. Separating wall curved outwards. Diameters of first chamber vary from 0-125 xO-120 mm. to 0-275x0-225 mm., those of second chamber from 0-105x0-050 mm. to 0-175x0-075 mm. Distance across both chambers varies from 0-200 to 0-325 mm. Equatorial chambers subquadrate to falciform in shape, usually higher than long, although reverse also quite common. In axial section first chamber circular, about 0-100 to 0-175 mm. in height. Alar prolongations extremely narrow. Appreciable reduction in thickness of spiral lamina at periphery. Mar- ginal cord distinct. Traces of weakly developed pillar-like structures in polar region. Microspheric form. Test small- to medium-sized, lenticular, with acute margin. Septal filaments radial, straight. Diameter of test varies from 4-6 to 8-9 mm., thickness from 1-9 to 3-3 mm., and ratio of diameter to thickness from 2-0 to 3-3. In equatorial section about 9 to 13 whorls occur, regularly increasing in height. Height of spiral cavity greater than thickness of spiral lamina. Septa nearly perpendi- cular at base, straight for about half their course, then curve sharply backwards. Thick- ness of septa decreases considerably from proximal to distal end. Chambers sub- quadrate to falciform, usually higher than long. In axial section chamber cavity triangular in shape. Alar prolongations extremely narrow. Considerable reduction in thickness of spiral lamina at periphery. Marginal cord weakly developed. In polar region traces of pillar-like structures. text-fig. 6. Nummulites pengaronensis Ver- beek. Part of the equatorial section of a microspheric specimen from locality Sa, x 18 approx. Remarks. Although the Assam specimens show considerable variation in external form SAMANTA: NUMMULITES, KOPILI FORMATION, ASSAM 679 of the test, they are identical in internal structures and are included here under one species. They are characterized externally by the presence of radial septal filaments and the absence of distinct polar pustules, and internally by the characters of the septa in equatorial section and very narrow alar prolongations in axial section. Nummulites saipanensis (Cole), originally described from Saipan, Mariana Islands (Cole and Bridge 1953), was later considered by its author (Cole 1957) to be synonymous with N. pengaronensis Verbeek. This is accepted here. Cole (op. cit.) also included Nummulites semiglobula (Doornink), described from Java (Doornink 1932), in the synonymy of N. pengaronensis. However, although N. semiglobula , as well as N. gerthi, bear some resemblance to N. pengaronensis, a more detailed study based on the topo- type material of these two species is needed before considering them as junior synonyms of N. pengaronensis. Among the European species, Nummulites stel/atus Roveda described from the Priabonian of North Italy (Roveda 1961), is very closely similar to N. pengaronensis in internal morphology. Sen Gupta (1965), while working on some Middle Eocene Nummulites from Western India, regarded N. pengaronensis as a junior synonym of Nummulites beaumonti d’Archiac and Haime. Under the remarks on N. beaumonti , Sen Gupta (op. cit., p. 93) wrote: ‘Another synonym of N. beaumonti is N. pengaronensis Verbeek, a wide-spread Indo-Pacific form. It shows a tight coiling of spiral wall, which is almost uniformly thick, and small embryonic chambers, as does typical N. beaumonti. These features are clearly seen in the figures of N. pengaronensis published by Cole (1957m pi. 231, figs. 1-17) and have been confirmed by an examination of Cole’s material from Eniwetok.’ Sen Gupta, therefore, neither examined the type or topotype materials nor consulted the type description and illustrations of N. pengaronensis to support his remarks. Further, he did not compare his Indian specimens of N. beaumonti with ones from the Indian region identified as N. pengaronensis by other workers. (There are good illustrations of micro- spheric and megalospheric specimens of N. pengaronensis from Eastern India (Nagappa 1959a, b). In both these publications, there are good illustrations of N. beaumonti too, and in the latter Nagappa has pointed out (p. 158, pi. 21, figs. 1, 2) the conspicuous difference in the character of the septa in these two species.) A thorough comparison between the two species, including such taxonomically important features as the charac- ters of the equatorial chambers and the septa as seen in equatorial section, has not been made and Sen Gupta’s remarks do not appear to be justified. A comparison of the description and illustrations of N. beaumonti provided by Davies (1940) with those of N. pengaronensis given by Verbeek (1871), Doornink (1932), and others, shows clearly that they are two distinct species. In equatorial section they can always be separated by the characters of the septa and the equatorial chambers, and in axial section by the width of the alar prolongations and the degree of development of polar plugs. The writer believes that these two species are not only distinct but that they belong to two different groups of species. If N. pengaronensis is considered to be a synonym of N. beaumonti, or in other words, if the morphological differences between them are not considered to be of specific importance, then the usefulness of species of Nummulites in the stratigraphic analysis and correlation of the Lower Tertiary will be greatly reduced. With N. pengaronensis as a junior synonym, the stratigraphic range of N. beaumonti would be from Middle Eocene to Oligocene (not Middle to Upper 680 PALAEONTOLOGY, VOLUME 11 Eocene as mentioned by Sen Gupta (1965, p. 92)) and it would then be difficult to use it as a ‘key’ species in stratigraphy. Distribution. N. pengaronensis is a widely distributed Indo- Pacific form and has been reported from the Central Pacific Islands, the East Indies, Burma, Eastern India, and Western Pakistan. Its known stratigraphic range is from the upper part of the Middle Eocene to Oligocene. In the Garo Hills N. pengaronensis ranges from the Upper Member of the Siju Limestone (Middle Eocene) to the overlying Kopili Formation (Upper Eocene). It is the most abundant representative of the genus in the Kopili Formation and occurs in all the five localities (see Table 1). Although it is known to occur in the Sulaiman Range (Eames 1952), there is no report of the species from the Upper Eocene of Surat-Broach, Western India. GENERAL REMARKS The four species of Nummulites recorded from the Kopili Formation belong to three different groups. N.fabianii (Prever) belongs to the reticulate group of forms charac- teristic of Upper Eocene to Oligocene. This is the only pillared form in the present assemblage. Of the three remaining striate forms, N. pengaronensis with its strongly curved septa and very narrow alar prolongations is distinctly different from N. cliavan- nesi and N. sp. aff. N. chavannesi , characterized by rapid opening of the whorls, straight septa, and wider alar prolongations. In all these four forms the marginal cord is only moderately developed and the size of the tests does not exceed 9 mm. The striate forms are much more abundant than the reticulate one, and occur in almost all foraminiferal samples. The assemblage of Nummulites in the Kopili Formation is markedly different from that in the underlying Upper Member of the Siju Limestone (Samanta 1968). In the latter horizon the assemblage is characterized by the presence of large, highly evolved species showing three ‘parts’ in the spire as recognized by Schaub (1963). These forms are totally absent in the Kopili Formation. Also, the number of species of Nummulites is fewer in the Kopili Formation than in the underlying Siju Limestone. Throughout the range of the genus in the Indian region the most striking change in the assemblages occurs at this horizon. The total absence of the typical representatives of the genus characterizing the older horizons, together with the appearance of a new group of forms in the Kopili Formation, makes the assemblage more akin to that of the Oligocene than to that of underlying Middle Eocene horizon. Indeed, in the absence of reticulate species it is difficult to distinguish the Nummulites assemblage of the Upper Eocene from that of the Oligocene. In the presence of N. fabianii and abundant small to medium striate forms, the present assemblage is closely comparable to that recorded in the Priabonian of Europe. It is distinguished from the latter essentially by the absence of the Nummulites striatus- garnieri group of forms, which are common in the European Upper Eocene. The Num- mulites assemblage in the Kopili Formation is, however, quite distinct from that known from the Upper Eocene of the Far East. N. pengaronensis is the only species common to the two regions. The Nummulites yawensis-djokdjokartae group of forms described from the Upper Eocene of the Malayan Archipelago are absent in the Kopili Formation of Assam. The absence of the well-known and widely distributed Upper Eocene reticulate Nummulites in the Far East constitutes the most striking difference between the Upper Eocene Nummulites assemblages of the two regions. SAMANTA: NUMMULITES , KOPILI FORMATION, ASSAM 681 REFERENCES adams, c. G. 1965. The Foraminifera and stratigraphy of the Melinau Limestone, Sarawak, and its importance in Tertiary correlation. Q. Jl geol. Soc. Loud. 121, 283-338, pi. 21-30. bayliss, d. d. 1961. An investigation of certain larger fossil foraminifera from Pakistan. Unpubl. Ph.D. thesis, Univ. Wales, 1-253, pi. 1-26. bieda, f. 1957. Die Fauna Grosser Foraminiferen im Obereozaen der Slowakei. Geol. Sb., Bratislava, 8, 28-71, pi. 2-6. 1963a. Larger foraminifers of the Tatra Eocene. Inst. Geol. Peace, 37, 1-215, pi. 1-26. 19636. Septieme niveau de grands Foraminiferes dans le Flysch des Karpates Polonaises. Ann. Soc. geol. Pologne, 33, 189-218, pi. 11-14. boussac, J. 1906. Developpement et morphologie de quelques foraminiferes de Priabona. Bull. Soc. geol. Fr. (4) 6, 88-97, pi. 1-3. — 1911a. Etudes stratigraphiques et paleontologiques sur le Nummulitique de Biarritz. Ann. Hebert, 5, 1-95, pi. 1-24. 19116. Etudes stratigraphiques sur le Nummulitique alpin. Mem. Carte geol. Fr. 1 122, pi. 1-5. bozorgnia, f. and kalantari, A. 1965. Nummulites of parts of Central and East Iran — A thin-section Study. 1-28, pi. 1-24. Tehran, Iran. caudri, c. m. b. 1934. Tertiary deposits of Soemba. Diss. Leiden Natl. Mas. geol., Amsterdam, 1-224, pi. 1-5. cizancourt, m. de. 1928. Sur quelques nummulites du flysch Karpatique et sur leur signification pour la stratigraphie des Karpates. Kosmos, Warsz. 53, 287-312, pi. 1, 2. - 1930. Sur la stratigraphie et la faune Nummulitique du Flysch de l’Albanie. Bull. Soc. geol. Fr. (4) 30, 195-212, pi. 22, 23. cloud, p. e. jun. and cole, w. s. 1953. Eocene Foraminifera from Guam, and their implications. Science, N.Y. 117, 323-4. cole, w. s. 1957. Larger Foraminifera from Eniwetok drill holes. Prof. Pap. U.S. geol. Surv. 260- V, 743-84, pi. 231-49. 1963. Tertiary larger Foraminifera from Guam. Ibid. 403-E, 1-28, pi. 1-11. and bridge, j. 1953. Geology and larger Foraminifera of Saipan Island. Ibid. 253, 1-45, pi. 1-15. daci, a. 1951. Etude paleontologique du Nummulitique entre Kiigukgekmece et Catalca. Rev. Fac. Sc. Univ. Istanbul, (B) 16, 89-246, pi. 1-5. davies, l. m. 1940. The Upper Khirthar beds of north-west India. Q. Jl geol. Soc. Lond. 96, 199-230, pi. 9-12. doornink, h. w. 1932. Tertiary Nummulitidae from Java. Verb, geol.-mijnb. Genoot. Ned. 9, 267-315, pi. 1-10. douville, h. 1912. Quelques foraminiferes de Java. Samml. geol. Reichsmus. Leiden, (1) 8, 279-94, pi. 22-24. eames, f. e. 1952. A contribution to the study of the Eocene in Western Pakistan and Western India: D. Discussion of the faunas of certain standard sections, and their bearing on the classification and correlation of the Eocene in Western Pakistan and Western India. Q. Jl geol. Soc. Lond. 107, 173-96. clarke, w. j. and banner, f. t. 1959. Nummulites retiatus, a synonym of Nummulites fabianii. Revue Micropaleont . 2, 113. evans, p. 1932. Explanatory notes to accompany a table showing the Tertiary succession in Assam. Trans. Min. geol. metall. Inst. India, 27, 155-260. fabiani, r. 1905. Studio geo-paleontologico dei Colli Berici. Atti 1st. veneto Sci. 64, 1797-1839. flandrin, J. 1934. La faune de Tizi Renif pres Dra el Mizan (Algerie). Bull. Soc. geol. Fr. (5) 4, 251-71, pi. 14-16. 1938. Contribution a l’etude paleontologique du Nummulitique algerien. Mater. Carte geol. Alger. (1) Paleont. 8, 1-155, pi. 1-15. hagn, h. 1960. Die stratigraphischen, palaeogeographischen und tektonischen Beziehungen zwischen Molasse und Helvetikum im ostlichen Oberbayern. Geologica Bav. 44, 1-208, pi. 1-12. harpe, p. de la. 1877. Note sur les Nummulites des Alpes occidentales. Verb, schweiz. naturf. Ges. 60, 227-32. 682 PALAEONTOLOGY, VOLUME 11 harpe p. de la. 1883a. Etude des Nummulites de la Suisse et revision des especes eocenes des genres Nummulites et Assilina. Abh. schweiz. palaont. Ges. 10, 141-60, pi. 3-7. 18836. Monographic der in Aegypten und der libyschen Wiiste vorkommenden Nummuliten. Palaeontographica, 30, 157-216, pi. 30-35. henrici, H. 1934. Foraminiferen aus dem Eozan und Altmiozan von Timor. Ibid. Suppl. 4, pi. 4, 1-56, pi. 1-4. nagappa, y. 1951. The stratigraphical value of Miscellanea and Pellatispira in India, Pakistan and Burma. Proc. Indian Acad. Sci. (B) 33, 41-8. 1959a. Foraminiferal biostratigraphy of the Cretaccous-Eocene succession in the India-Pakistan- Burma region. Micropaleontology, 5, 145-92, pi. 1-11. 19596. Note on Operculinoides Hanzawa 1935. Palaeontology, 2, 156-60, pi. 21-23. rao, s. r. n. 1941. The Tertiary sequence near Surat and Broach (Western India), with description of foraminifera of the genus Pellatispira from the Upper Eocene of this region. Half-yrly J. Mysore Univ. N.s. (B) 2, 5-17, pi. 1, 2. roveda, v. 1959. Nummulites retiatus nouvelle espece de nummulite reticulee des Abruzzes (Italie). Revue Micropaleont. 1, 201-7, pi. 1. 1961. Contributo alio studio di alcuni macroforaminiferi di Priabona. Riv. ital. Paleont. Stratigr. 67, 153-224, pi. 14-19. samanta, b. k. 1965. Discocyclina from the Upper Eocene of Assam, India. Micropaleontology, 11, 415-30, pi. 1-4. 1968. The Eocene Beds of Garo Hills, Assam, India. Geol. Mag. 105, 124-35. schaub, h. 1963. Uber einige Entwicklungsreihen von Nummulites und Assilina und ihre stratigra- phische bedeutung. Evolutionary trends in Foraminifera, 282-97. Amsterdam. sen gupta, b. k. 1965. Morphology of some key species of Nummulites from the Indian Eocene. J. Paleont. 39, 86-96, pi. 15-17. verbeek, r. d. m. 1871. Die Nummuliten des Borneo-Kalksteines. Neues Jb. Miner. Geol. Palaont. Jahr. 1871, 1-14, pi. 1-3. 1892. Voorloopig bericht over Nummuliten, Orbitoiden en Alveolinen van Java, en over den ouderdom der gesteenten waarin zij optreden. Natuurw Tijdschr. Ned.-Indie, 51, 101-39, pi. 1. and fennema, r. 1896. Description geologique de Java et Madoura. 2 vols. 1-1183, pi. 1-11. Amsterdam. vlerk, i. m. van der. 1929. Groote foraminiferen van N. O. Borneo. Meded. Dienst Mijnb. Ned.-Indie 9, 1-44. yabe, H. 1921. Note on some Eocene Foraminifera. II. Notes on two Foraminiferal Limestone from E. D. Borneo. Sci. Rep. Tdhoku Univ. II Series (Geol.), 5 (4), 100-6, pi. 17, 18. BIMAL K. SAMANTA Department of Geology University College of Wales Typescript received 3 January 1968 Aberystwyth, Wales A NEW PLANT FROM THE LOWER OLD RED SANDSTONE OF SOUTH WALES by DIANNE EDWARDS Abstract. A new plant is described from the Senni Beds of the Lower Devonian of South Wales. The naked axes were pseudomonopodially and dichotomously branched and contained protosteles in which the protoxylem was central. Terminal fructifications consisted of sporangia alternating with sterile bracts; the plant was homo- sporous. A comparison is made with other Devonian genera, which show similar organization in the fertile regions, and it is concluded that the plant should be placed in a new genus, Krithodeophyton, assigned to the Barinophytaceae (Incertae sedis). The fossils described in this paper were among those collected by Croft from the Brecon Beacons Quarry, also called the Storey Arms Quarry (Nat. Grid Ref. SO 972208) and are now in the Department of Palaeontology, British Museum (Natural History). The quarry is in the Senni Beds, which form the lower part of the Breconian Stage of the Lower Old Red Sandstone in South Wales (Croft 1953) and are probably equivalent to the Siegenian of Europe. Among the plants previously described from this locality are Gosslingia breconensis, Zosterophyllum Hanover anum, and Drepanophycus spinae- formis (Heard 1927 and 1939, Croft and Lang 1942, Edwards and Banks 1965, Edwards 1967). The majority of the plants were preserved as compressions in a fine-grained, blue-grey sandstone, but parts of the axes were sometimes petrified. Small pieces of cuticle were recovered after bulk maceration of the rock in commercial strength (40%) hydrofluoric acid. These were then treated with Schulze’s solution (con- centrated nitric acid and potassium chlorate) for 1-8 hours and, when the carbon had oxidized and the outlines of cells were visible, the fragments were washed, immersed in Diaphane solvent and mounted in Diaphane (Distributors: Will Scientific Inc., New York 52, N.Y., U.S.A.). Pieces of carbon were also picked oft' both axes and sporangia with steel needles, macerated in Schulze’s solution and mounted in the same way. Film pulls were made from those fossils, which were exposed on the surface of the rock. A solution of cellulose nitrate in amyl acetate was poured over the fossil and left to dry overnight. The resulting rough, transparent film was peeled off and mounted in Harleco Synthetic Resin (H.S.R.). (Distributors: Arthur H. Thomas Company, Philadelphia, Pa., U.S.A.). The anatomy of the pyritised axes was investigated using a modification of the method described by Beck in 1955. A small piece of rock containing a petrified fossil was first embedded in the synthetic resin, Ceemar, because the rock matrix tended to crumble when sawn (Leclercq and Noel 1953). The transparent block of plastic was trimmed to size and cut into sections about a millimetre thick. These were ground smooth on a glass plate with grade 600 carborundum powder, washed and then treated in chromic oxide powder until a good polish was obtained. After washing they were dried, soaked in xylol and mounted in H.S.R. The sections of the axes were then examined using a Leitz Ultropak microscope. All the preparations have been deposited in the Department of Palaeontology, British Museum (V26578, V26579, V521 37-54). [Palaeontology, Vol. 11, Part 5, 1968, pp. 683-90, pis. 130-32.] 684 PALAEONTOLOGY, VOLUME 11 SYSTEMATIC DESCRIPTION Family barinophytaceae Krausel and Weyland 1961 Genus krithodeophyton gen. nov. Type species. Krithodeophyton croftii sp. nov. Diagnosis. Plant consisting of naked, pseudomonopodially branching axes with some dichotomous branching in distal parts. Simple protostele composed of mainly scalari- form, but some reticulate tracheids; protoxylem central. Sporangia in terminal spikes. Dichotomous branching within the base of the fertile region or just below it. Oval sessile sporangia borne in two rows, one on either side of the axis. Narrow sterile appendages, attached at right angles to the axis, alternate with the sporangia. Bracts straight throughout their length or with distal parts curving downwards. Sporangium wall composed of isodiametric cells. Plant homosporous. Spores assignable to the dis- persed spore genus Apiculiretusispora ( sensu Streel 1964). Krithodeophyton croftii sp. nov. Plate 130, figs. 1-12; Plate 132, figs. 1-10 Diagnosis. Plant, at least 10-0 cm. high, consisting of naked, pseudomonopodially branching axes, 1 -5-4-3 mm. wide, with some dichotomous branching in the distal parts; wide angles (< 80°) at branching points. Small bud-like lateral branches about 4-0 mm. long. Simple,(circular protosteleTjaverage diameter 0-5 mm., composed of mainly scalariform, with some reticulate, tracheids; smallest elements at the centre of the xylem. Outer cortex composed of elongate, thick-walled cells, 36-60 /a in diameter. Epidermis composed of short, fusiform cells. Sporangia aggregated into terminal spikes, 2-5-3-0 mm. wide and up to 1-3 cm. long. Dichotomous branching within the base of the fertile region or just below it. Oval sessile sporangia, 1-25-1-5 mm. long and 0-8-1 -0 mm. wide borne in two rows, one on either side of the axis. Narrow sterile appendages up to 2-5 mm. long, given oft' at right angles to the axis, alternate with the sporangia. Bracts straight throughout their length or with distal parts curving downwards. Sporangium wall composed of isodiametric cells, of average diameter 27 p. Homosporous. Spores approximately circular in outline, average diameter is 60 /a; simple trilete f-f of the spore radius long; variable wall elements up to 3 p high (the majority being under 1 /a); ornament reduced or absent in contact area. Assignable to the dispersed spore genus Apiculiretusispora. EXPLANATION OF PLATE 130 Figs. 1-12. Krithodeophyton croftii sp. nov. 1,4, Rock bearing fertile spikes, xl-1 (V26578 and V26579 respectively). 2, 5, Sterile axes, x 0-9 (V52154 and V52152 respectively). 3, Film pull showing epider- mal cells, X 80 (V52151, BMP 25). 6, Film pull from axis showing small lateral branch. (Inset = specimen before treatment), x6 (V52154, BMP 12). 7, Film pull from axis showing cortical cells, x 50 (V52144, BMP 1). 8, Central strand after treatment with Schulze’s solution, x 200 (V52137, BMM 22). 9, Central strand on film pull, X 100 (V52145, BMP 2). 10, T.S. axis showing xylem strand before a division, x 108 (V52150, RSI/8). 11, T.S. pyritised xylem strand with centre not preserved, x 108 (V52138, RSI/4). 12, T.S. part of outer cortex, X 108 (V52150, RSI/5). Palaeontology, Vol. II PLATE 130 EDWARDS, Lower Old Red Sandstone plants DIANNE EDWARDS: A NEW PLANT FROM SOUTH WALES 685 Locality. Brecon Beacons Quarry, abandoned roadside quarry on the A470 between Brecon and Merthyr, approximately 1\ miles south of Brecon. Horizon. Senni Beds, Breconian, Lower Old Red Sandstone of South Wales (= Siegenian). Holotype. Specimen V26579, Department of Palaeontology, British Museum (Natural History), London. Description of vegetative parts. The over-all height, branching frequency, and basal parts of the plant are unknown. The most complete fertile specimen found was 10 cm. long, with axes 2 mm. in diameter. There were two branching points (3-4 cm. apart) in the vegetative region, and a further dichotomy occurred within the base of the spike of sporangia. The diameter of the sterile axes ranged between T5 and 4-0 mm. and axes of similar size were always found together. Hence it is probable that larger ones formed the basal, and small axes the more terminal, parts of the plant. Branching was pseudo- monopodial in these wider axes, and the narrower ‘lateral’ branch formed a wide angle with the ‘main’ axis (PI. 130, fig. 2). These axes tended to be slightly flexuous and, in addition to a central line, 0-5 mm. in diameter, faint longitudinal striations were some- times visible. Small, lateral bud-like structures up to 4-0 mm. long, which tapered in width from base to apex were also present (PI. 130, fig. 6). Each was supplied with a central strand of xylem. The axes associated with the fertile regions were much narrower (< 20 mm. wide) and branched dichotomously. Film pull and maceration preparations revealed some cellular detail of the outer layers of the axes. The outer cortex was the tissue most frequently recovered and con- sisted of very long cells with thick walls, 36-44 p apart. Occasionally the tapering, over- lapping ends of these cells were seen (PI. 130, fig. 7). Scattered among the thick parallel walls were thinner, irregular lines probably representing the remains of underlying tissue. The outer cortical cells were represented in the transverse sections of petrified axes by 1 or 2 layers of angular cells (35-60 p in diameter) surrounding a structureless mass of pyrites, containing a circular protostele. The outer walls of the cortical cells were often broken down giving the surface of the axis a hairy appearance (PI. 1 30, fig. 1 2). In one instance only, a piece of cuticle was found where the cells had thinner walls, were fusiform in shape and relatively short (PI. 130, fig. 3). It is probable that this was the epidermis. The central strand of xylem was composed mainly of scalariform tracheids in which pits on adjacent walls were opposite (PI. 130, figs. 8 and 9). A few elements showed reticulate pitting. Sections through petrified xylem revealed similar anatomy. The central part was usually not preserved (PI. 130, fig. 11), but a decrease in tracheid diameter from the outside of the xylem inwards was always apparent. In a few cases, very small elements could be seen at the centre. The protoxylem is therefore considered to be central. The largest tracheids measured 35 p in diameter and the distance between the horizontal bars of the scalariform tracheids was sometimes as great as 10 p. In one axis only could stages in the division of the xylem be seen. A few millimetres below the branching point, the xylem became oval in cross section (PI. 130, fig. 10) and divided to give two circular strands of approximately the same diameter. The axis itself then divided into two. I have found well-preserved pyritised axes with similar anatomy to those described above at the same locality. I include a detailed account of them here to emphasize this 686 PALAEONTOLOGY, VOLUME 11 close anatomical and morphological similarity, but in the absence of any reproductive parts in the new specimens, I feel it would be unwise to conclude that the two sets of remains belong to the same plant. The axes were at least 10-0 cm. long and 2-0-3-0 mm. wide. They branched dichoto- mously and were slightly flexuous. One axis bore a small protuberance 2-0 mm. long and T5 mm. wide (PI. 131, fig. 1) comparable to the small lateral branch in Krithodeo- phyton. The surfaces of the axes were striated. Film pulls showed thick longitudinal walls similar to those described above and sometimes smaller fusiform cells were seen (PI. 131, fig. 4). Transverse and longitudinal sections through petrified axes showed a circular pro- tostele composed predominantly of scalariform tracheids (PI. 131, figs. 2 and 5). Some reticulate pitting was seen. The widest tracheids were found to the outside of the xylem. The cells at the centre were either not preserved or very small and crushed (PI. 131, fig. 5). Immediately outside the xylem was a narrow band of squashed thin-walled cells. One or two layers of thick-walled cells, circular to angular in cross-section, with an average diameter of 49 p, formed the outermost layer of the axis. The outer walls of these cells were eroded away so that the radial walls project into the matrix, giving the axis a hairy outline (PI. 131, fig. 6) which is also seen in Krithodeophyton. In longitudinal sections, the cortical cells had tapering ends (PI. 131, fig. 3). They were, on average, 350 p long. Stages in the division of the xylem below a branching point are illustrated in Plate 131, figs. 5-10. Description of fructifications. The sporangia were aggregated into terminal spikes. A few millimetres below the fertile region an axis dichotomised and this was followed by a further dichotomy immediately below or within the base of the fructification (PI. 130, fig. 4). Isolated fructifications were common in the matrix (PI. 130, fig. 1). The spikes were at least 1-3 cm. long and, on average, were 2-75 mm. wide. The majority were incomplete at the apex and 6-0-8-0 mm. long. At the base they were parallel-sided, but there was a gradual decrease in width in the distal parts and the apices were rounded (PI. 132, fig. 4). No organization was apparent in the distal parts. The sporangia were oval in outline, 1 -25-1 -5 mm. long and 0-8-1 -0 mm. wide (PI. 1 32, fig. 2). They were sessile on a central axis (0-3 mm. diameter) which was often obscured by the sporangia them- selves. The sporangia appeared to be arranged in two rows one on either side of the axis, each row containing at least eleven sporangia. It is unlikely that this arrangement was produced by compression of an originally spiral or whorled organisation as the spor- angia are quite distinct. At the base of the spike the long axes of the sporangia were orientated at right-angles to the central axis, in the distal part they were directed toward the apex. Alternating with the sporangia and extending beyond them out into the matrix were thin, possibly spine-like appendages up to 2-5 mm. long (PI. 132, figs. 3 and 5). EXPLANATION OF PLATE 131 Figs. 1-10. Krithodeophyton croftii sp. nov. 1, Branching sterile axes from Brecon Beacons Quarry (small lateral projection indicated by arrow), x 1 (DE 32/1). 2, Section through petrified axis, L.S. xylem strand composed of scalariform tracheids, x 108 (32-4-33). 3, As last, L.S. outer cortical cells, X 108 (32-4-35). 4, Film pull of surface of axis showing epidermal cells, X 80 (DE32. P77). 5-10, Series of sections of petrified axis showing division of xylem at a branching point, x51 (Series 32-4). Palaeontology, Vol. 11 PLATE 131 EDWARDS, Lower Old Red Sandstone plants DIANNE EDWARDS: A NEW PLANT FROM SOUTH WALES 687 These arose at right angles to the fructification axis and sometimes curved downwards at their tips. The exact relationship between sporangium and lateral appendage or bract could not be determined with certainty as the structures were very heavily carbonised. In some cases, the base of a sterile bract extended down the central axis partially covering the sporangium below, giving the impression that the sporangium and the bract above it formed a single unit (PI. 132, fig. 1). Plate 132, fig. 2 shows a fructification, where sporangia only are visible, while in Plate 132, fig. 6 only bracts are present. Where sporangia and bracts were slightly superimposed the sporangia appear beaked. At the moment it seems very likely that the appendages were not attached to the sporangia, but that both were borne separately on a central axis. Description of spores. Maceration and film pull preparations showed spores adhering to the sporangium wall. The cellular structure of the walls was not preserved, but some very fragmentary remains indicated that part of the wall was composed of isodiametric cells, around 27 p, in diameter. Imprints left by the spores could frequently be seen in the sporangium wall (PI. 132, fig. 7). A small number of spores were recovered by dissolving the film pull bearing a sporangium (PI. 132, figs. 8-10). They were all approximately the same size, with an average diameter of 58 /x (range: 55-68 /x). The spores were almost circular in outline. The position of the trilete was usually difficult to establish because of heavy folding, but in one case a simple Y-mark was observed. The length of the trilete was almost equal to the spore radius. The wall ornament was made up of rather variable elements ranging from coni to spinae, which were 1 /x or less high. Narrower parallel- sided elements with truncated tips occasionally reached a height of 3 /x. There was a well-developed contact area, where the ornament was much reduced or absent. These spores are assignable to the dispersed spore genus Apiculaliretusispora ( sensu Streel 1964) and are comparable to the species A. brandtii (Streel 1964). In conclusion it must be emphasized that no anatomy has been detected in the axes attached to the fertile regions. The sterile and fertile parts are thought to belong to the same plant because they are always found in close association, no other plant being present, and where fructifications terminate relatively long axes, the latter have the same dimensions, surface features, and branching angles as do the sterile axes. Discussion. Plants with a similar arrangement of sporangia in terminal spikes and, in some cases, with sterile and fertile appendages intermixed, are known from other Devonian deposits. These include Barinophyton, Protobarinophyton, Pectinophyton, and the possible reproductive parts of Enigmophyton. Hoeg (1942) described some fertile axes, which he considered were probably attached to Enigmophyton superbum. These plants came from the upper Middle Devonian of Spitzbergen. As in the Welsh plant, the sporangia, which alternated with sporophylls, were arranged in two rows, with dichotomous branching occurring both below and within the base of the fructification. Hoeg, however, thought that the two rows were produced by compression of an origin- ally verticillate or spiral arrangement of appendages. He considered the sporangia to be borne proximally on the upper surfaces of the sterile bracts. Unlike Krithodeophyton, this plant was heterosporous, and Vigran (1964), described the spores in detail. She identified the microspores as Phyllothecotriletes microgranulcitus (average diameter 72 /x) and called the megaspores Enigmospora simplex (200-70 p diameter). Although the microspores are comparable in size to those spores isolated from Krithodeophyton , the 688 PALAEONTOLOGY, VOLUME 11 ornament is quite different. Nothing resembling the fan-shaped leaves of Enigmophyton superbum has been found at the Brecon Beacons Quarry. Earlier, Hoeg (1935) had described a Middle Devonian plant from Norway, in which naked axes, with mainly pseudomonopodial branching had fructifications terminating the ‘lateral’ branches. The sporangia, or perhaps the organs bearing the sporangia, were arranged in two rows, both borne on the lower surface of the fructification axis. Hoeg called this plant Pectinophyton norvegicum and Ananiev (1957), described another species P. bipectinatum, where the axis of the fructification bore two rows of long, lateral appendages, twisted towards the axis. On the inside of the resulting coils were attached the sporangia. The axis had a central strand of annular tracheids. Whereas Petrosyan, in Lepekhina, Petrosyan, and Radchenko (1962) considered the above two species to be identical, Hoeg (1967) maintained that they were two distinct species on the basis of differences in sporangial shape and arrangement. Not enough is known of Hoeg’s plant to make a satisfactory comparison with Krithodeophyton , but P. bipectinatum with its apparently complex organisation in the sporangial region, is quite unlike the Welsh plant. The genus Barinophyton was erected by White (1905), when he re-investigated some plants from the Upper Devonian of Perry, Maine. He distinguished three species, Barinophyton richardsonii ( Lepidostrobus richardsonii Dawson 1861, Lycopodites richard- sonii Dawson 1863), B. obscurum ( Pecopteris (?) obscurum Dun 1898), and B. perryanum. The most completely described species in the literature are B. citrulliforme from the Upper Devonian of New York (Arnold 1939) and B. richardsonii, re-described by Pettitt (1965). Arnold showed that branching in the sterile region was pseudomonopodial but that the fructifications, much larger than in Krithodeophyton, terminated lateral branches. In both Barinophyton species the fructification axis bore on its dorsal surface two rows of appendages between which were found the sporangia. They were unlike Krithodeo- phyton in that the appendages did not extend beyond the sporangia and were discoidal in shape. In addition, in Pettitt’s re-construction two sporangia occurred between suc- cessive bracts in each row. Where the sporangia were borne in two rows, one on either side of the fertile axis, i.e. the normal Krithodeophyton arrangement, Arnold thought the arrangement was produced during fossilization. Both species were heterosporous. A Middle Devonian species, B. sibiricum (Lepekhina et al. 1962) differed from the Welsh plant in having long, lax, pendant spikes of sporangia. In 1954, Ananiev described B. obrutschevii from the Lower Devonian of the U.S.S.R. This differed from previously described species of Barinophyton in that branching was dichotomous not pseudomonopodial. On the basis of this difference alone, he transferred the plant to a new genus Protobarinophyton obrutschevii (Ananiev 1957). Although in 1954, Ananiev had briefly mentioned sporophylls associated with the sporangia, in later papers by him and other authors no mention is made of them. Certainly Protobarinophyton obrutschevii f. mucronata (originally Distichophytum mucronatum Ananiev) transferred EXPLANATION OF PLATE 132 Figs. 1-6. Fructifications of Krithodeophyton croftii sp. nov.; s, sporangium; b , bract; d, branching within the base of the fertile region. 1, x 5 (V26579). 2 and 6, x 2-3 (V26578). 3-5, x 5 (V26578). 7, Sporangium wall with spores attached, x 500 (V52146, BMM 12). 8-10, Spores isolated from sporan- gia. 8, x 500 (V52146, BMS 1). 9 and 10, x 1000 (V52146, BMS 1). Palaeontology, Vol. 11 PLATE 132 EDWARDS, Lower Old Red Sandstone plants DIANNE EDWARDS: A NEW PLANT FROM SOUTH WALES 689 to the genus by Lepekhina et a/. (1962) had no bracts in the fertile region. Ananiev (1963) stated that the sporangia of P. obrutschevii contained spores of one size only (70 p diameter). Krithodeophyton and Protobarinophyton resemble each other in many ways. Indeed, considering such characters as their similar age, the aggregation of oval, sessile sporangia into terminal spikes, homospory and dichotomous branching in the fertile regions, they might be included in the same genus. However, until more is known of the detailed organization of the spikes (e.g. whether or not bracts are present), about the spores and the anatomy of the axes of Protobarinophyton , I feel justified in erecting a new genus for the Welsh plant. Differences include the absence from Krithodeophyton of U- and H- branching, the absence of dehiscence lines on the sporangia and the presence of scalari- form not annular tracheids in the protostele. In addition, the spikes in Krithodeophyton were much smaller, more compact and had distinctly protruding bracts. The specimen illustrated in Plate 132, fig. 2, in which no bracts are visible, most closely resembles the Russian plant. It is possible that this specimen represents a late stage in the maturation of a spike, where the sporangia were large and completely obliterated the bracts. This explanation could be applied to the apparently bractless condition seen in Protobarino- phyton. Variation within the species has been indicated by Lepekhina et al., who dis- tinguished a P. obrutschevii f. minuta from P. obrutschevii f. typica , but little is known of the variation in fructification morphology within a single population. In conclusion it is stressed that not enough is known of the details of anatomy and morphology of any of the plants discussed above. It may be that ultimately they will all be placed in the same genus. Indeed Arnold believed that this was true for Pectinophyton and Barinophyton. On the information available at the moment, I propose to place this relatively completely described Lower Devonian plant in the new genus Krithodeophyton (Derivation: Krithodes — like barley). Classification. Ananiev (1963) included Protobarinophyton, with Zosterophyllwv and Bucheria in the Zosterophyllaceae. My own observations on the anatomy of Z. llano- veranum show that Krithodeophyton is not related to this species, because the latter had an elliptical exarch protostele, and therefore should not be placed in the same family. In his contribution to the second volume of the Traite de Paleobotanique (ed. Boureau 1967), Hoeg included the Barinophytales as an order, Incertae Sedis, containing two families, Barinophytaceae and Barrandeinaceae. The former included, in addition to Barinostrobus, all the genera mentioned above with the exception of Enigmophyton. It is suggested that Krithodeophyton should also be placed in this family. Acknowledgements. I thank Dr. K. R. Sporne for his help in preparing this paper, and Dr. J. B. Richardson and Dr. F. A. Hibbert for their assistance in identifying the spores. I am indebted to the Keeper of Palaeontology of the British Museum (Natural History) for the loan of specimens. This work formed part of a Ph.D. thesis submitted to the University of Cambridge and was carried out during the tenure of a N.A.T.O. Research Studentship. REFERENCES ananiev, a. r. 1954. On the Lower Devonian flora of the south-east part of Western Siberia. Voprosy Geologii Azii, 1, 287-324, pi. 1-5. [Russian Translation Service, National Lending Library.] — 1957. New fossil plants from the Lower Devonian deposits of the village of Torgashino, south- eastern zone of Western Siberia. Bot. Zh. Moscow, 42, 691-702, pi. 1-3. [Ibid.] 690 PALAEONTOLOGY, VOLUME 11 ananiev, a. r. 1963. Psilopsida in V. A. Vakhrameev (ed.) Osnovy Paleontologii, 14, 315-43, pi. 1-8 [In Russian.] Arnold, c. a. 1939. Observations on fossil plants from the Devonian of Eastern North America, IV. Plant remains from the Catskill Delta deposits of Northern Pennsylvania and Southern New York. Contr. Mus. Geol. Univ. Mich. Ann Arbor, 5, 271-313, pi. 1-10. beck, c. b. 1955. A technique for obtaining polished surfaces of sections of pyritized plant fossils. Bull. Ton. hot. Cl. 82, 286-91. croft, w. n. 1953. Breconian: a stage name in the Old Red Sandstone. Geol. Mag. 90, 429-32. and lang, w. h. 1942. The Lower Devonian flora of the Senni Beds of Monmouthshire and Breconshire. Phil. Trans. R. Soc. Loud. B221, 131-63, pi. 9-11. dawson, j. w. 1861. On the pre-Carboniferous flora of New Brunswick, Maine and Eastern Canada. Can. Naturalist, 6, 161-80. 1863. Further observations on the Devonian plants of Maine, Gaspe and New York. Q. Jl geol. Soc. Loud. 19, 458-69, pi. 17-19. dun, w. s. 1898. On the occurrence of Devonian plant-bearing beds on the Genoa River, County of Auckland. Rec. geol. Surv. N.S. Wales, 5, 117-21, pi. 10, 11. edwards, d. 1967. An investigation of certain Devonian plants. Ph.D. thesis. University of Cam- bridge. and banks, H. p. 1965. Branching in Goss/ingia breconensis. Am. J. Bot. 52, 536. heard, a. 1927. On Old Red Sandstone plants showing structure from Brecon (South Wales). O. Jl geol. Soc. Load. 85, 195-209, pi. 13-15. 1939. Further notes on Lower Devonian plants from South Wales. Ibid. 95, 223-9, pi. 12, 13. hoeg, o. a. 1935. Further contributions to the Middle Devonian flora of Norway. Norsk geol. Tidsskr. B15, 1-18, pi. 1-8. 1942. The Downtonian and Devonian flora of Spitzbergen. Skr. Svalb. og Ishavet 83, 1-228, pi. 1-62. 1967. Psilophyta. In Traite de Paleobotanique, vol. 2. Boureau ed. Paris. [In French.] krausel, r. and weyland, h. 1961. Uber Psilophyton robustius Dawson. Palaeontographica 108B, 11-21, pi. 4-5. leclercq, s. and noel r. 1953. Plastic: a suitable embedding substance for petrographic study of coal and fossil plants. Phytomorphology, 3, 222-3. lepekhina, v. G., Petrosyan, n. m., and Radchenko, g. p. 1962. Main Devonian plants of the Altay- Sayan Mountain region. Trudy vses. naucho-issled. geol. Inst. 70, 61-189, pi. 1-24. [Russian Transla- tion Service, National Lending Library.] pettitt, j. m. 1965. Two heterosporous plants from the Upper Devonian of North America. Bull. Brit. Mus. ( nat . hist.) Geol. 10, 83-92, pi. 1, 2. streel, m. 1964. Une association de spores du Givetien inferieur de la Vesdre, a Goe (Belgique). Annls. Soc. geol. Belg. 87 B1-B30, pi. 1, 2. vigran, j. o. 1964. Spores from Devonian deposits, Mimerdalen, Spitzbergen. Skr. norsk. Polarinst. 132, 1-33, pi. 1-6. white, d. 1905. In G. O. Smith and D. White. The geology of the Perry Basin in South Eastern Maine. Prof. Pap. U.S. geol. Surv., Washington, 35, 9-92, pi. 2-6. D. EDWARDS Botany School Downing Streel Typescript received 1 March 1968 Cambridge DENCKM ANNITES (TRILOBITA) FROM THE SILURIAN OF NEW SOUTH WALES by L. SHERWIN Abstract. The trilobite Denckmannites Wedekind is redefined and a new species, Denckmannites rutherfordi, is described from Siluro-Devonian shales near Orange in the Central West District of New South Wales. The genus Denckmannites has hitherto been recorded only from the Siluro-Devonian of Central Europe and the Devonian of Morocco (Richter, in Moore 1959, p. 0 467). The species described in this paper, Denckmannites rutherfordi , was found in the Wallace Shale (Stevens and Packham 1952) at a locality about 18 miles west of Orange in the Central West district of New South Wales. The two bands bearing the trilobites were separated by a thickness of 100 ft. of apparently unfossiliferous olive-green shales. Some uncinate monograptids accompanying the trilobites in both bands indicated either a late Silurian, or less probably, an early Devonian age. Other fossils in the lower band include Dictyonema sp., Encrinurus mitchelli Foerste, an indeterminate odontopleurid trilobite, orthoconic nautiloids, and small orthid and linguloid brachiopods. Apart from Monograptus sp., the only associates of Denckmannites in the upper band are Dictyonema sp. and Encrinurus sp. The morphological terms are as in the Treatise, Part O. SYSTEMATIC PALAEONTOLOGY Order phacopida Salter 1864 Genus denckmannites Wedekind 1914 1931 Phacopidella ( Denckmannites ) Wedekind; Richter and Richter, p. 143 (cum syn). 1959 Denckmannites Wedekind; Richter, p. 0 467. Type species. Phacops volborthi Barrande 1852; SD Vogdes 1925. Emended diagnosis. Cephalon with broad border and shallow border furrow along posterior edge, both continued antero-laterally as far as axial furrows; border furrow joins axial furrow in front of eye; in front of glabella border is reduced or absent. Mar- gin of cephalon sharp, upper surface meeting doublure at acute angle; 3p and 2p glabellar furrows vestigial, equal in depth, unconnected with axial furrows; lp furrows initially as deep as axial furrows, becoming much shallower axially. Intercalating ring tripartite, distinct from glabella. Eyes small, of cryptothalmus pattern. Entire posterior section of facial suture with forwardly directed convex curvature; genal angles rounded. Vincular furrow crenulate, developed only on postero-lateral doublure. Pygidium large, rounded, with narrow border and smooth surface. Remarks. In 1914 Wedekind erected the subgenus Glockeria ( Denckmannites ) without selecting a type, referring the definition instead to the ‘group of Phacops volborthi Barrande’ which also included P. miser Barrande and P.fugitivus Barrande. P. volborthi [Palaeontology, Vol. 11, Part 5, 1968, pp. 691-6, pi. 133.] C 6055 Z Z 692 PALAEONTOLOGY, VOLUME 11 was subsequently designated as the type by Vogdes in 1925; Richter and Richter (1931) confirmed this action by applying Art. 30, la of the then current Rules for Zoological Nomenclature. In defining G. ( Denckmannites ) Wedekind stated that the cephalon is totally surrounded by a border. However, this feature is found only in P. volborthi and possibly P. fugitivus, but not in P. miser, although Wedekind included them within his subgeneric conception. In addition, he stated that the Ip glabellar furrows did not coalesce, although they are shown connected in figures of P. volborthi and P. miser by Barrande (1852). Vogdes (1925) designated P. volborthi as the type for Phacopidella ( Denckmannites ), but retained Wedekind’s diagnosis unaltered. Richter and Richter (1931), in revising the subgenus, omitted any reference to the cephalic border and lp glabellar furrows. They distinguished Phacopidella ( Denckmannites ) from P. ( Phacopi- della) by its ’uniform glabellar surface anterior to the lp furrows as in Phacops (i.e., an undivided glabellar surface with only mere vestiges of 2p and 3p furrows), eyes of a cryptothalmus pattern, and a tripartite intercalating ring’ (p. 143, in German). Richter (in Moore 1959, p. O 467) raised Denckmannites to full generic status, at the same time modifying the 1931 diagnosis by referring to the intercalary furrow as tripartite instead of the intercalating ring. The genal angles were described as truncate, while the entire posterior section of the facial suture was specified as concave; i.e., with the concave side directed posteriorly. Alberti (1966) designated P. (£>.) micromma (A. Roemer) as the type species for Phacopidella ( Struveaspis ) thus restricting the scope of the genus Denck- mannites ( sensu Richter 1959). This new subgenus is characterized by fusion of the reduced central lobe of a tripartite intercalating ring to the base of the glabella by medial reduction of the lp furrows. For the sake of conformity in the two following lists, Struveaspis is regarded as being of the same generic status as Denckmannites. Species retained in genus Denckmannites : Denckmannites volborthi (Barrande) 1852 Denckmannites miser (Barrande) 1852 Denckmannites rutherfordi sp. nov. Species transferred to Struveaspis Struveaspis micromma (A. Roemer) 1852 Struveaspis fugitivus (Barrande) 1872 Struveaspis thuringica (Kegel) 1932 Struveaspis prantli (Ruzicka) 1946 Struveaspis n.sp. A. Alberti 1966 Denckmannites rutherfordi sp. nov. Plate 133 Source of name. After the Rutherford family, on whose property the fossils were found. Material. Six more or less complete exuviae preserved as internal moulds with fragments of the exo- skeleton adhering; USGD (University of Sydney Geology Department) 8914, 8916, 8934, and MMF (Geological and Mining Museum, Sydney) 14467. The external counterparts of USGD 8929 and 8934 have also been preserved. The only deformation is that caused by burial pressure, with crushing of the glabella. In addition, abundant imperfect specimens were also collected from the same locality. Holotype. USGD 8927 a, b. Paratypes. USGD 8929, 8934, MMF 14467. Locality. Portion 216, Parish Boree Nyrang, County Ashburnham. Sydney University Geology Department Locality catalogue MN/II/27. SHERWIN: DENCKMANNITES FROM NEW SOUTH WALES 693 Diagnosis. Denckmannites with very small eyes (about 20 lenses in each), deep trans- glabellar lp furrows becoming shallow axially, and exceedingly narrow anterior cephalic border; pygidium with 6-7 discrete axial rings. Description. The carapace possesses a long oval outline and, except for a strongly convex axis, has only moderate relief. The cephalon has a micro-granulose surface and a semicircular outline, with rounded genal angles which do not project noticeably beyond its posterior edge. The interior of the cephalon is pitted. The wide border is defined by a shallow border furrow, and is continuous along the posterior and lateral A B text-fig. 1. A. Reconstruction of cephalon showing inferred position of facial sutures, b. Reconstruction of ventral surface of cephalon, showing approximate shape of doublure. edges of the cephalon, maintaining a more or less uniform width. In front of the glabella the border narrows considerably; none of the specimens was sufficiently well preserved to determine if it disappears altogether. The border furrow runs parallel to the margin as far as the anterior corner of the cheeks, where it swings sharply into the axial furrow. Anterior to the lp furrows the axial furrows are straight, and are generally deeper than the border furrow. Near the cephalic margin they fade, and outline the glabellar curve to the edge of the cephalon before disappearing entirely. The glabella is large, pentagonal, and slopes forward gently to meet the doublure in an acute angle. The 2p and 3p furrows are weak, and do not meet the axial furrows. In different specimens the 2p furrows are represented by straight or curved furrows, the curve always being directed anteriorly. The posterior branches of the 3p furrows are generally more strongly curved than the 2p furrows; the 3p anterior branch is straighter and inclined at about 30 degrees to the axial furrows. The lp furrows cut deeply into the glabella for about one-third of their length, where they divide into two branches. The anterior branches swing forward decreasing in depth, being joined axially by a shallow depression. The posterior branch cuts the intercalating ring and joins the occipital furrow. The intercalating ring is thus trisected into a broad medium lobe and two ovoid lateral lobes. This division is reflected by the vaguely tripartite occipital ring. The cheeks have a slightly lower relief than the glabella; approximately triangular in shape, they support very small eyes at their anterior corners. Preservation does not indicate whether an anterior section of the facial suture exists or not; the posterior section has a forwardly directed curvature throughout, and intersects the cephalic margin just in front of the genal angles. 694 PALAEONTOLOGY, VOLUME 11 The eyes are oval in shape, supported on a reniform palpebral lobe. Preservation does not allow an accurate count, but each eye probably has between 1 9 and 2 1 lenses arranged in six vertical rows as follows (front-rear): 2, (?3), 3, 3, (?4), 4, 4, 3. In the holotype (USGD 8927) a portion of the exoskeleton partially covering one of the eyes shows no expression of the underlying lenses, unlike other species even in the same genus where the lenses are quite prominent surface features. Unless the exoskeleton in this region was transparent, it seems unlikely that the eyes could have functioned as light sensitive organs. The hypostome is triangular in shape with a slightly convex anterior margin. The anterior wings are particularly well developed, with appreciable thickening at their tips. On the ventral surface there is a small longitudinal ridge which runs from the lateral notch, fading at the anterior margin: on the dorsal surface the ridge is expressed as a shallow furrow. The border is continuous posterior to the lateral notches, and is broadest along the posterior edge. Any other details of the posterior border were obscured in all specimens by either cephalon or thorax. The median body is subcircular, apparently undivided, and strongly convex. No maculae were observed. The entire surface of the hypostome is finely granulose. The cephalic doublure is crescentic in shape, broadest beneath the glabella, with a paraboloid posterior margin which commences near the genal angles and passes just behind the eyes. The crenulate vincular furrow is widest near the genal angles, tapers forward, and disappears near the axial furrow. The doublure on small specimens seems comparatively rigid, as in such cases the posterior portion of the glabella, and to a lesser extent that of the cheeks, is crushed along an arc corresponding to the portion of the cephalon unsupported by the doublure. In large specimens, crushing of the cephalon is quite independent of the doublure. As regards ornament, the doublure seems comparable to the dorsal cephalic surface. The thorax is virtually parallel-sided, showing an appreciable taper in only the last three segments. The axial rings are strongly convex, and have a tripartite division into a central lobe, flanked by smaller, globose, lateral lobes. The central lobe carries a par- ticularly well-formed articulating half-ring, separated from it by a deep but broad transverse furrow. At the junction of central and lateral lobes, the extension of this EXPLANATION OF PLATE 133 Denckmannites rutherfordi sp. nov. Figs. 1 , 2, 5, 8, 9. Holotype USGD 8927. 1 , composite internal mould of cephalon, showing eye covered by exoskeleton which gives no indication of underlying lenses, X 5. 2, Composite internal mould of cephalon, X 1-5. 5, hypostome, x 3. 8, Composite external mould of cephalon, and internal mould of thorax and pygidium. Short hooks are visible on the fragments of exoskeleton adhering to the pleural extremities. 9, Silastic cast of external mould and hypostome, X 1-5. The anterior granulation is caused by pitting on the interior surface of the exoskeleton. Fig. 3, Paratype MMF 14467. Internal mould exhibiting phacopid or Salterian mode of moulting, x 2. Fig. 4, Paratype USGD 8934. Latex cast of external mould, showing border and interpleural furrows on pygidium, xl-5. Figs. 6, 7, Paratype USGD 8929a, b. 6, Internal mould, showing crushing of portion of cephalon unsup- ported by doublure, X 3. 7, Interior of dorsal carapace showing apodemes, x 3. An attempt to make a latex cast resulted in damage to the exoskeleton. Palaeontology, Vol. 11 PLATE 133 SHERWIN, Denckmannites SHERWIN: DENCKMANNITES FROM NEW SOUTH WALES 695 furrow projects downward to form a robust apodeme. These apodemes are flange shaped and taper downwards, with a slight expansion at their tips. The pleurae are more or less uniform in width throughout their length, with deep pleural furrows extending from axial furrow to articulating facet. These facets are large, extending about one half the length of the leading edge and a third the length of the posterior edge of the segments. The first five segments have rounded extremities; the remaining six show progressive development of a short spine or hook, possibly a feature associated with enrollment. The pygidium has a very rounded outline, marked by a flat border. On internal moulds this border is exaggerated by the imprint of the doublure which extends about one-third the distance across the pleural region. Articulating facets occupy about half the length of the leading ribs. The axis contains six discrete axial rings, a seventh vaguely discernible in some specimens, followed by a fused terminal piece. The first ring is partially obscured by the last thoracic segment. Both the first and second rings are transversely tripartite, though the division is much less pronounced than in the thoracic axial rings. Each pleural field bears five rather flattened ribs, separated by straight and narrow pleural furrows. Interpleural furrows are seldom observed except on external mounds, and are parallel to the pleural furrows. The surface is free of ornament. Dimensions (in mm.). Holotype USGD 8927: Width Length Cephalon 22 12 Thorax 24 20 Pygidium 17 9 Remarks. The bohemian species, Denckmannites miser (Barrande), most closely resembles D. rutherfordi, but has much larger eyes, and the 3p glabellar furrows, as shown in Barrande’s illustrations, appears to be whole, not bipartite. D. volborthi (Barrande) has eyes of about the same size, with a similar number of lenses in each, as D. rutherfordi , but its cephalon is wholly surrounded by a border, while the pygidium has almost twice as many axial segments as D. rutherfordi. Acknowledgements. I wish to thank Dr. J. W. Pickett of the Geological and Mining Museum, Sydney and Dr. B. D. Webby of Sydney University for a critical review of this paper; Dr. K. S. W. Campbell of the Australian National University, Canberra, for much helpful discussion and criticism; Dr. G. H. Packham of Sydney University for forwarding useful information concerning the stratigraphic position of this species. Permission to publish this paper was given by the Under-Secretary, New South Wales Department of Mines. REFERENCES alberti, g. k. b. 1966. Uber einige neue Trilobiten aus dem Silurium und Devon, besonders von Marokko. Senck. leth. 47, 1 1 1-21, pi. 4. barrande, J. 1852. Systeme Silurien du Centre de la Boheme: Iere Partie, Crustaces, Trilobites. 1, 935 pp., 51 pi. Prague-Paris. richter, r. 1959. In R. C. Moore (ed.), Treatise on invertebrate paleontology. Part O, Arthropoda 1, Lawrence, Kansas. and richter, E. 1931. Unterlagen zum Fossilium Catalogus, Trilobitae. V. Senckenbergiana, 13, 140-6. stevens, N. c. and packham, g. h. 1952. Graptolite zones and associated stratigraphy at Four Mile Creek, S.W. of Orange, N.S.W. /. Proc. R. Soc. N.S.W. 86, 95-8. 696 PALAEONTOLOGY, VOLUME 11 vogdes, a. w. 1925. Palaeozoic Crustacea. Pt. 11. A list of the genera and subgenera of the Trilobita. San Diego Soc. Nat. Hist. Trans. 3, 87-115. wedekind, r. 1914. Palaontologische Beitrage zur Geologie des Kellerwaldes. Abb. preuss. geol. Landesanst. 69, 1-84, pi. 1-5, figs. 1-26. L. SHERWIN Geological and Mining Museum 36-64 George Street North Sydney, New South Wales, 2000 Typescript received from author 18 April 1968 Australia MORPHOLOGY AND FUNCTION OF DICHOPORITE PORE-STRUCTURES IN CYSTOIDS by C. R. C. PAUL Abstract. The cystoids, a diverse and artificial group of extinct Palaeozoic echinoderms, are characterized by the possession of pore-structures developed in the thecal wall and composed of thecal canals which open in thecal pores. Thecal canals may occur singly to form dipores, or in sets perpendicular to plate sutures to form rhombs. Five basic types of pore-structure are recognised: pectinirhombs, cryptorhombs, humatirhombs, humatipores, and diplopores. The first two are composed of dichopores, thecal canals which connect external pores and through which sea-water flowed in life. The dichopores of pectinirhombs open in slits; those of cryptorhombs open in pores, one of which is sieve-like. Pectinirhombs and cryptorhombs are characteristic of, and confined to, the super-families Glyptocystitida and Hemicosmitida respectively; humatirhombs characterize the superfamily Caryocystitida. Five types of pectinirhombs and one type of cryptorhomb are recognized. Pectinirhombs and cryptorhombs agree closely with the paradigm of an exchange system and were respiratory structures. Evolution of pectinirhombs proceeded from less to more efficient types independently in all families of Glyptocystitida. The three rhombiferan superfamilies probably acquired their rhombs independently. Rhombifera and Diploporita are regarded as separate classes. The former contains two Orders: Dichoporita and Fistuliporita. The cystoids, an extinct group of Palaeozoic echinoderms, are characterized, inter alia, by the possession of pore-structures in the calcareous wall of the theca. Although various authors from Muller (1854) to Regnell (1945) and Kesling (1963) have em- phasized the importance of these pore-structures in cystoid morphology and taxonomy, they have not been the subject of complete morphological or functional analyses. Stain- brook (1941) and Sinclair (1948) considered the function of pectinirhombs in one genus each and Delpey (1942) the function of pectinirhombs and cryptorhombs. The last- named author suggested that pectinirhombs and cryptorhombs were balancing organs but most authors have accepted a respiratory function for all pore-structures. Function- ally all cystoid pore-structures fall into two groups: dichoporite and non-dichoporite. This paper deals with the dichoporite pore-structures (pectinirhombs and crypto- rhombs) after a brief review of the morphology of the five basic types of pore-structure recognized. Function in non-dichoporite pore-structures will be the subject of a future paper. The cystoids of this account correspond to the class Cystoidea of Kesling (1963) but only to the subclass Hydrophoridea of Regnell (1945). The hydrospires of blastoids and epispires of certain eocrinoids and crinoids are not considered nor are the peculiar pore-structures of paracrinoids like Comarocystites Billings. Knowledge of the mor- phology of cystoid pore-structures has grown piecemeal, virtually every researcher add- ing some contribution. One result of this is a rather imprecise terminology often related to ill-defined concepts. In reviewing the morphology of pore-structures an attempt has been made to introduce a systematic terminology and to define the concepts upon which this terminology is based. Previous usage has been followed as far as possible. Through- out this account the classification is that of Kesling (1963) unless otherwise stated. [Palaeontology, Vol. 11, Part 5, 1968, pp. 697-730, pis. 134^40.] 698 PALAEONTOLOGY, VOLUME 11 Acknowledgements. The author is grateful to the following for the loan of, or access to, specimens in their care: Dr. R. P. S. Jefferies and Mr. H. G. Owen, British Museum, Natural History (BMNH); Dr. I. Strachan and Dr. G. R. Coope, Birmingham University, Department of Geology (BU); Dr. T. E. Bolton and Dr. G. W. Sinclair, Canadian Geological Survey (CGS); Dr. R. V. Melville and Dr. A. Rushton, Geological Survey and Museum, London (GSM); Dr. W. D. I. Rolfe, Hunterian Museum, Glasgow (HM); Dr. J. C. Harper, Liverpool University, Department of Geology (LU); Dr. H. Mutvei, Naturhistoriska Riksmuseet, Stockholm (RM); Dr. C. D. Waterston, Royal Scottish Museum, Edin- burgh (RSM); Mr. A. G. Brighton, Sedgwick Museum, Cambridge (SM); Dr. K. E. Caster, University of Cincinnati, Department of Geology (UC); Mr. H. L. Strimple, University of Iowa, Department of Geology (UI); Dr. D. B. Macurda Jr., University of Michigan, Museum of Paleontology (UMMP); Dr. P. M. Kier and Mr. T. Phelan, United States National Museum, Washington (USNM); and Dr. Katherine G. Nelson, University of Wisconsin, Milwaukee, Department of Geology (UWM). Parts of this work were completed during the tenure of a Harkness Scholarship and a National Environmental Research Council research scholarship at the Sedgwick Museum, Cambridge; and a post-doctoral fellowship at the Museum of Paleontology, University of Michigan (NSF GB-3366). All three are gratefully acknowledged. Mr. P. Minton, Imperial College, Department of Civil Engineering and Dr. R. P. S. Jefferies con- tributed much useful discussion on the functioning of rhombs. Dr. Jefferies kindly reviewed the manu- script and suggested several improvements. MORPHOLOGY OF CYSTOID PORE-STRUCTURES A. GENERAL The term pore-structure embraces all the structures dealt with. All pore-structures are composed of thecal canals and thecal pores. These have not always been distinguished from each other in the past but Kesling (1963) noted that two separate concepts are involved in the terms pore and canal. Thecal pores are perforations in the surfaces of the thecal wall. They cannot have a separate existence without the thecal canals to which they give rise, i.e. thecal pores bear the same relationship to thecal canals that the two ends of a stick do to the stick. Pores in the external surface of the thecal wall are external pores (text-fig. 1) and correspond- ingly, internal pores (text-fig. 2) are developed in the internal surface of the thecal wall. External pores may be slit-like, and are then referred to as slits (text-fig. 12), or they may be either simple or sieve-like (text-figs. 3, 20). When simple a single thecal canal ter- minates in a single pore, when sieve-like the canal divides near the external surface of the thecal wall and opens in a cluster of small pores. Apparently internal pores are always simple and circular in cross-section. Thecal canals are tubular structures which connect , or terminate in, thecal pores. Thecal canals are analogous to the entire stick plus its ends. The portions of thecal canals which are approximately perpendicular to the surface of the thecal wall are termed perpendicular canals (text-fig. 2) and those tangential to the surface, tangential canals (text-fig. 2). All thecal canals are composed of one or more tangential canals which connect a pair of perpendicular canals and they all commence and terminate in two thecal pores, both of which lie in the same surface of the thecal wall (with the possible exception of haplopores). Tangential canals may be incompletely calcified or uncalcified, i.e. made of soft tissue only as in diplopores, in which case only the pair of perpendicular canals is preserved in fossils. Thecal canals may be compound (text-fig. 4) when several tangential canals connect a pair of perpendicular canals, or simple (text-figs. 1, 2) when one tangential canal connects the perpendicular canals. Here thecal canals are considered PAUL: DICHOPORITE PORE-STRUCTURES IN CYSTOIDS 699 as functional units rather than morphological units. Portions of canals composed entirely of soft tissue, as is the case with external papulae, are considered integral parts of the canal even though their former existence can only be inferred in fossils. Previously two types of pore-structures have been recognized: ‘pore rhombs’ (or more simply and accurately, rhombs ) and ‘diplopores’. The latter term has been used in both a specialized and a generalized sense. The new term dipore is proposed to cover all types of pore-structures within the Diploporita and diplopore is restricted to one type of dipore following current usage. 1 2 EP S TC text-figs. 1-4. Thecal pores and thecal canals. 1. Diagrammatic longitudinal section through a pectinirhomb dichopore to show external pores. 2. Diagrammatic longitudinal section through a simple fistulipore to show internal pores. 3. Diagrammatic longitudinal section through a cryptorhomb dichopore to show simple and sieve- like external pores. 4. Diagrammatic plan view of a compound fistulipore with surface layer removed to show two perpendicular canals connected by a pair of tangential canals. CP, sieve pore; EP, external pore; IP, internal pore; SP, simple pore; PC, perpendicular canal; TC, tangential canal; S, plate suture. In text-figs. 1-3 and all other figures unless otherwise stated, the external surface of the thecal wall is towards the top of the figure. Rhombs (fig. 5) may be defined as pore-structures composed of a set of thecal canals each of which arises in one thecal plate, crosses a plate suture and terminates in an adjacent plate , the whole set having a rhombic outline in plan view. Rhombs have their thecal canals developed more or less perpendicular to the plate suture and were randomly orientated with respect to other features in the theca since plate sutures occur in almost all possible orientations. Rhombs are therefore orientated and described with respect to their sutures (text-fig. 5). The outline of a rhomb may be described as indicated in text-figs. 6-11. Rhombs may be composed of two types of thecal canal. Jaekel (1899) proposed the term dichopore for the canals (and their terminal pores) of the Rhombifera, Jaekel’s Dichoporita. Sub- sequently dichopore has been restricted to the canals of pectinirhombs. A fundamental difference exists between canals which terminate in internal pores and those with external pores. No complete canal can have both. Here the term dichopore is restricted to rhomb canals with external pores (text-figs. 1, 3) and the new term fistulipore (fistula = canal) is proposed for rhomb canals with internal pores (text-figs. 2, 4). There are two types of rhombs composed of dichopores and a third composed of hstulipores. Pectinirhombs (text-figs. 15-19) are composed of dichopores which open in slits, whereas cryptorhombs 700 PALAEONTOLOGY, VOLUME 11 (text-fig. 20) are composed of dichopores which open in pores. Humatirhombs are rhombs composed of fistulipores. Dipores may be defined as pore-structures composed of a single thecal canal randomly orientated with respect to the thecal plates. Dipores are composed of one type of thecai canal, which resembles a fistulipore in connecting internal pores but differs in rarely crossing plate sutures. There were two types of dipore. A diplopore is a dipore composed text-fig. 5. Diagram of the principal features of a rhomb. Every rhomb is developed in two thecal plates (1 and 2). A plate suture divides the rhomb into two half-rhombs and the width is measured along it. Perpendicular to the suture lies the rhomb axis which divides the rhomb into two demi-rhombs and along which the length is measured. Thecal canals adjacent to the axis are adaxial, those away from the axis lateral and the outermost one or two are marginal. Positions and directions ( adsutural , adaxial, etc.) are defined with respect to the plate suture and the rhomb axis. of a simple thecal canal in which the tangential portion was uncalcified, i.e. the tangential portion formed a papula in life and only the two perpendicular portions of the canal are found fossil. Humatipores are dipores composed of compound thecal canals in which all the tangential canals were fully calcified. These five major types of pore-structure are recognized: pectinirhombs, crypto- rhombs, humatirhombs, diplopores, and humatipores. The first two differ fundamentally from the other three in that they are composed of thecal canals (dichopores) which connect external pores. Hudson (1911, 1915) recognized this distinction and described dicho- porite pore-structures as endothecal and non-dichoporite as exothecal. These general terms may be applied to all types of pore-structures in primitive echinoderms. A sixth type of pore-structure, haplopores, has been attributed to cystoids. Haplopores, PAUL: DICHOPORITE PORE-STRUCTURES IN CYSTOIDS 701 as here understood, consist of a single, blind (in life), perpendicular canal which opens in a single internal pore. No true haplopores have been observed in any cystoids examined by the writer but this does not imply that none ever existed. The first information on the structure of the thecal wall, which is intimately related to the morphology of pore-structures, was given by Barrande (1887). Barrande maintained that there were three layers to the thecal wall and that the outer and inner layers sealed text-figs. 6-11. Outlines of rhombs. Rhombs in which the length exceeds the width are compressed (6); the converse, depressed (7). Rhombs with unequal half-rhombs are unequal (8); the converse, equal (6, 7). Rhombs with unequal demi-rhombs are asymmetrical (9); the converse, symmetrical (6, 7, 8). These features occur in combina- tion. A symmetrical, equal rhomb which is neither depressed nor compressed is ideal (10). Rhombs with thecal canals oblique to the suture are oblique (11). A, rhomb axis (an imaginary line); S, plate suture. Text-figs. 6-11 represent disjunct pectinirhombs in plan view but the terms may be applied to all types of rhombs. Rhombs which are ‘depressed’ into the theca may be termed sunken to avoid confusion with those with depressed outlines. the thecal canals which were confined to the middle layer. Several subsequent authors have described an epithecal layer sealing canals externally but the present writer has found no evidence to support Barrande’s statement. Barrande described the surfaces of his layers (as preserved in moulds in most cases) and not cross-sections. All sub- sequent authors also described surfaces not sections. When the external surface was without pores Barrande claimed an epithecal layer was present; when pores were visible on the external surface he claimed the epitheca was eroded before preservation. How- ever, external surfaces with and without pores may result from the type of canal developed. A rhombiferan with dichopores will have pores in the external surface but one with fistulipores will not unless the outer walls of the tangential canals are eroded. 702 PALAEONTOLOGY, VOLUME 11 A similar distinction may be made between diplopores, with tangential canals which were of soft tissue and never preserved, and humatipores with fully calcified tangential canals. The description of sealed dichopores in cryptorhombs may possibly result from post-mortem calcification. However, it is more probably due to a misinterpretation of sieve pores as a progressive sealing of the canals. Finally sealed canals do not necessarily indicate the presence of a separate layer sealing them. All thin sections prepared by the author lack any indication of separate layers sealing the canals but the thecal wall is definitely not uniform in structure. text-fig. 12. Diagram to illustrate the principal features and measurements of pectinirhomb dicho- pores and slits. Dichopore and slit lengths and widths are measured in the same direction as rhomb lengths and widths. In this figure and text-figs. 15-20 the external surface of the thecal wall faces the top of the figures and part of the rhomb is removed to show internal features. B. THE MORPHOLOGY OF PECTINIRHOMBS Pectinirhombs are rhombs composed of dichopores which open externally in slits (text-fig. 12). The term pectinirhomb was first used by E. Forbes (1848, as 'pectinated rhomb’) for the rhombs of the superfamily Glyptocystitida. The various features of pectinirhomb morphology will now be described. The pores (text-figs. 1, 12) of pectinirhombs are slits. The principal measurements of slits and dichopores are shown in text-fig. 12. There may be a single conjunct slit along the entire length of a dichopore (text-figs. 15, 17). Alternatively the dichopore may open in a pair of disjunct slits, one at each end (text-figs. 16, 18). One disjunct slit is generally longer than the other and all the longer slits occur in one half-rhomb, a fact first noted by Stainbrook (1941, p. 93) for the pectinirhombs of Strobilocystites White. The slit lengths are also approximately proportional to the length of the dichopores to which they give rise and thus marginal slit lengths are less than adaxial slit lengths (text-fig. 32). PAUL: DICHOPORITE PORE-STRUCTURES IN CYSTOIDS 703 The canals (text-figs. 1, 12) of pectinirhombs are dichopores. Pectinirhomb dichopores are discrete (text-fig. 1 3) when each dichopore is a separate entity and the inter- dichopore space is covered by a thin strip of unmodified thecal plate. The inter-dicho- pore spaces may vary in width in a pectinirhomb with discrete dichopores, and the dichopores open externally in slits each provided with an individual slit rim. Confluent dichopores (text-fig. 14) are dichopores in which the walls of adjacent dichopores are con- fluent and the inter-dichopore spaces are covered with dichopore walls. In such cases the inter-dichopore widths are constant and equal to the dichopore widths. Confluent dichopores open in slits grouped into vestibules (text-figs. 17, 18) and collectively 13 14 text-figs. 13-14. Discrete and confluent dichopores. 13. Discrete. Each dichopore is separated from adjacent dichopores by a strip of unmodified thecal plate. 14. Confluent. The walls of adjacent dichopores are continuous and the pectinirhomb is made of isoclinal folds. Both figures represent sutural views of conjunct pectinirhombs. surrounded by a vestibule rim. In disjunct pectinirhombs with confluent dichopores one half-rhomb has a complete closed rim which surrounds the slits on both the ab- and ad-sutural sides. The other half-rhomb has an open rim on the ab-sutural side only (text-fig. 18). A number of types of pectinirhomb may therefore be recognized on the character of the slits and dichopores. Five types are in fact recognized: 1. Conjunct pectinirhombs with discrete dichopores (text-fig. 15) 2. Disjunct ,, „ ,, „ (text-fig. 16) 3. Conjunct ,, ,, confluent ,, (text-fig. 17) 4. Disjunct ,, ,, ,, ,, (text-fig. 18) 5. Multi-disjunct pectinirhombs (text-fig. 19) The structure of multi-disjunct pectinirhombs is not yet fully known. Conjunct pectinirhombs are much less common than disjunct pectinirhombs and discrete dichopores are only found in Ordovician genera. Conjunct pectinirhombs with discrete dichopores are only found in the genus Cheirocrinus s./.: seven species have this type of pectinirhomb. Conjunct pectinirhombs with confluent dichopores occur in the genera Homocystites Barrande, Pleurocystites Billings, and Regulaecystis Dehm. Dis- junct pectinirhombs with discrete dichopores occur in the Cheirocrinidae, Rhombi- feridae, Cystoblastidae, and Echinoencrinitinae. The Scoliocystinae, Callocystitidae, and Praepleurocystis Paul have disjunct pectinirhombs with confluent dichopores. Multi- disjunct pectinirhombs occur in two Upper Ordovician species of Cheirocrinus s.l ., C.jamesi (McCoy), and C. interruptus (Jaekel). The transition from discrete to confluent dichopores seems to have been gradual and to have taken place independently in several 704 PALAEONTOLOGY, VOLUME 11 text-figs. 15-18. Diagram to illustrate the morphology of pectinirhombs. 1 5. Conjunct with discrete dichopores. 16. Disjunct with discrete dichopores. 17. Conjunct with confluent dichopores. 18. Disjunct with confluent dichopores. text-fig. 19. Diagram to illustrate the interpretation of the morphology of a multidisjunct pectinirhomb. PAUL: DICHOPORITE PORE-STRUCTURES IN CYSTOIDS 705 families. In particular the pectinirhombs of Glyptocystites spp. are intermediate in character. Isolated demi-rhombs only occur in species which have pectinirhombs with discrete dichopores. All five types of pectinirhomb are confined to and characteristic of the superfamily Glyptocystitida. Members of this superfamily have thecae composed of a definite number of plates arranged in five circlets. There are 4 basals (BB), 5 infra-laterals (ILL), 5 laterals (LL), 4, 5, or 6 radials (RR), and 7 orals (OO). The plates are numbered clockwise around the theca from the left of the periproct. The positions of pectinirhombs on the theca may text-fig. 20. Diagram of the morphology of a cryptorhomb. be denoted by citing the two plates in which the half-rhombs are developed. Thus pec- tinirhomb L4:R3 has one half-rhomb in L4 and the other in R3. Pectinirhombs are only developed across certain sutures and some positions are occupied more frequently than others. No orals ever bear pectinirhombs and there is no marked preference for pec- tinirhombs to develop across sutures between different circlets (inter-circlet sutures) as opposed to sutures between plates of the same circlet (intra-circlet sutures). C. THE MORPHOLOGY OF CRYPTORHOMBS Cryptorhombs are rhombs composed of dichopores which open externally in pores (text-fig. 20). The term cryptorhomb is proposed because the entire length of the tan- gential part of all the dichopores is hidden from external view so that in many species the rhombs are inconspicuous. The pores of cryptorhombs (text-figs. 3, 20) may be simple or sieve-like (compound pores of text-fig. 20). In almost all cases examined each dichopore opened in a sieve- pore at one end and a simple pore at the other. In all complete cryptorhombs one half- rhomb has sieve-pores only and the other has simple pores only. A few dichopores may 706 PALAEONTOLOGY, VOLUME 11 be reversed with respect to the others in some incomplete cryptorhombs. The dichotomous branching of the perpendicular canal which opens in a sieve-pore may proceed evenly in all branches to produce clusters of 2, 4, 8, 1 6, or 32 individual pores, or unevenly, produc- ing intermediate numbers of individual pores. The individual pores may be circular, oval, or dumb-bell shaped : this depends on the stage of branching. Individual pores are usually surrounded by pore rims and may be set in shallow depressions or raised in rounded tubercles. Within a sieve-pore, individual pores may be arranged randomly or in rows. Simple pores may be developed on tubercles, or in or beside ridges on the plates. Nearly all simple pores have strongly developed rims. Cryptorhomb dichopores (text-figs. 3, 20) resemble discrete dichopores of pectini- rhombs in the sense that the inter-dichopore spaces are covered by unmodified thecal plate. Dichopores may be evenly developed throughout a cryptorhomb, which then has a regular rhombic outline and is complete. Incomplete cryptorhombs have few, irregularly spaced, dichopores and irregular outlines. Cryptorhomb dichopores may be weakly calcified within the theca. Only one type of cryptorhomb (text-fig. 20), which corresponds in structure to a disjunct pectinirhomb with discrete dichopores, has been recognized. However, crypto- rhombs may be complete or incomplete and the unusual rhombs of Polycosmites Jaekel may represent a second type of cryptorhomb, corresponding to multi-disjunct pectini- rhombs in structure. Unfortunately no specimens are available. Cryptorhombs are confined to and characteristic of the superfamily Hemicosmitida, members of which have thecae composed of a definite number of plates arranged in 3 or 4 circlets. These are designated by the same symbols as in the Glyptocystitida but the arrangement of plates is less constant. There may be 3 or 4 basals, 6 or 10 infra-laterals, 8 or 9 laterals, and, in the Hemicosmitidae, 9 radials. In the Caryocrinitidae the oral area is covered by a special ‘tegmeiT which apparently replaces the radial circlet. Crypto- rhombs are not developed in tegminals but may be present in all other plates. In many species complete cryptorhombs are predominantly developed across inter-circlet sutures while incomplete cryptorhombs are confined to intra-circlet sutures of the lateral and infra-lateral circlets. There are thus rings of complete cryptorhombs which correspond to the inter-circlet sutures. Furthermore the arrangement of the pores of complete cryptorhombs is constant (text-fig. 21). In the BB:ILL ring sieve-pores occur in the basals; in the ILL: LL ring they are in the laterals; in the LL: RR ring (Hemicosmitidae only) they are in the radials. Incomplete cryptorhombs complicate this arrangement but functionally they are of minor importance. THE FUNCTION OF DICHOPORITE PORE-STRUCTURES A. THE BASIS OF FUNCTIONAL INTERPRETATIONS The principles on which this functional analysis of cystoid pore-structures has been based were outlined by Rudwick (19646) but a brief resume will now be given. It is necessary first to assume that any organ under investigation was functional and to suggest a function for it. This assumption may be tested in fossils only if it is possible to predict a structure which would serve the function efficiently and which is directly comparable with the preserved hard parts of the fossil. This prediction has been termed a paradigm by Rudwick (1960) and is assumed to have maximum efficiency within the PAUL: DICHOPORITE PORE-STRUCTURES IN CYSTOIDS 707 limits imposed by the materials of which the fossil structure is made. A close morpho- logical similarity between the fossil structure and the paradigm would be expected if the former functioned efficiently. Where such similarity is found it is probable that the postu- lated function was served in life but this cannot be said to have been proved absolutely. Three points arise in this connection. 1. The preserved hard parts of the fossil did not constitute the entire structure in life, i.e. at the time of functioning. It is necessary to reconstruct the unpreserved soft parts and in doing so it is assumed that the soft text-fig. 21. Diagrammatic representation of the distribution of simple and sieve-pores in the theca of Hemicosmites extraneus Eichwald. Only complete cryptorhombs are shown. F, arm facet; G, gono- pore; N, node in lateral plates; Pe, periproct. Based on RM. Ec5381. (unpreserved) parts of the fossil bore the same relationship to the preserved hard parts that the soft parts of related living organisms do to their hard parts. For example, Sinclair’s (1948, p. 306) suggestion that the dichopores of the pectinirhombs of Glypto- cystites were lined with ciliated epithelium may be accepted since dichopores are invaginations of the external surface of the thecal wall and nearly all modern echino- derms have an external ciliated epithelium. 2. The second assumption that the fossil structure functioned efficiently is only a working assumption and may be modified to fit any case. To test any postulated func- tion a close morphological similarity between the fossil structure and the paradigm is sought. If it is found, the fossil structure could have served the function efficiently. If not, it cannot be concluded that the structure did not serve the postulated function but only that it was inefficient if it did. The early functional morphologists evolved as an empirical law the idea that any given organ is only developed to the minimum efficiency conducive with survival C 6055 3 A 708 PALAEONTOLOGY, VOLUME 11 (i minimum survival efficiency ) and not the maximum potential efficiency. A low relative efficiency in an organ may result from a low minimum survival efficiency, perhaps due to a lack of competition or low selective pressures. At present this is an hypothetical consideration since evaluation of minimum survival efficiencies in fossils is impossible and there is little information from living animals. Any structure may form part of two (or more) functional systems, or may have two (or more) concomitant functions, each with different structural requirements. Such structures may become a compromise between the ideal cases. If it is possible to suggest the two (or more) functions, it may become possible to test whether departure from one paradigm is due to approximation to another. A low relative efficiency in one structure may be compensated for by a high efficiency elsewhere in the system. Equally minimum survival efficiency may be achieved by a few efficient organs or a large number of less efficient organs. Thus the second assumption that the fossil structure functioned efficiently is only a working assumption and con- siderable information on the functioning of a fossil structure may be gained even when there is not a close comparison with the paradigm. 3. There is always a danger of using teleological language in functional analyses since ‘function’ and ‘purpose’ are sometimes used synonymously. Yet no fossil ever evolved a structure to serve a purpose. To investigate the ‘effect’ of a structure need neither imply function nor purpose. Thus if a chance rock fall produces a narrow crack between two boulders through which the tide washes, the crack will prevent particles of a dia- meter greater than its width from passing between the boulders. The crack may be analysed in the same way as zigzag slits in brachiopods (Rudwick 1964#) or pectini- rhomb slits (p. 715) and will be found to have the same effect. Since it was a chance rock fall which produced the crack there can be no purpose behind it nor does it serve any function. Rudwick (1964fl) concluded that if zigzag slits in brachiopods were pro- tective they could have served this function efficiently. That is: the effect of zigzag slits in brachiopods was to reduce the maximum size of particles that could enter (or leave) the mantle cavity for any given gape. If fossil brachiopods behaved in a similar manner to living brachiopods, those with zigzag slits were better ‘protected’ from large particles than those without (for any given gape). This may have been beneficial to these brachio- pods. The ‘function’ of a structure may be considered as a ‘beneficial effect’ or better still as the most beneficial of its several effects. Although the purpose of functional analyses is to determine the ‘function’ of structures, it is the mode of functioning, i.e. the sum of the ‘effects’ of the detailed morphology of the structures, which is investigated. In the next sections attempts are made to interpret the effects of the detailed morphology of pectinirhombs and cryptorhombs, having assumed that they were functional in life. B. THE GENERAL FEATURES OF EXCHANGE SYSTEMS Most authors have accepted a respiratory function for all cystoid pore-structures. It is true that Delpey (1942) has suggested that pectinirhombs and cryptorhombs were balancing organs and Hyman (1955) among others, has suggested that some canals may have been nutritive. However, Sinclair (1948, pp. 307-8) has shown how unlikely it is that pectinirhombs or cryptorhombs were balancing organs. If thecal canals were nutritive, they would be evenly distributed or at least present in all thecal plates. This is not always the case with pectinirhombs, cryptorhombs, and diplopores but apparently PAUL: DICHOPORITE PORE-STRUCTURES IN CYSTOIDS 709 it is the case with humatipores and humatirhombs. The most plausible function for pore-structures is respiration. If pore-structures were respiratory this was achieved by diffusion of oxygen and carbon dioxide and the pore-structures formed part of an exchange system. To test this hypothesis the paradigm of an exchange system must be examined. The most essential feature of an exchange system is that only the ‘exchange substance’ or ‘substances’ should be exchanged; direct mixing must be prevented. There must be an exchange surface which allows the passage of the exchange substances while keeping the donor and acceptor fluids separate. In respiration the donor current for oxygen is also the acceptor current for carbon dioxide. Throughout donor current is used with reference to oxygen exchange. The rate of exchange will depend on the following factors: 1. The area of the exchange surface. The larger the area of the exchange surface, the greater the amount of exchange. The exchange surface should be as large as its strength will allow. Echinoderm skeletal calcite is a meshwork of calcite rods and soft tissue. Oxygen and carbon dioxide diffused not through the skeletal calcite but rather through the soft tissues. In echinoderms no more than half the area of an exchange surface was available for exchange because of this meshwork structure. 2. The resistance to exchange of the exchange surface. The resistance of an exchange surface will be proportional to its thickness and will depend on its nature. Although the walls of the thecal canals are heterogeneous only the soft tissue functioned as exchange surface. The thinner the walls are, the lower their resistance to exchange, and their strength, will be. The paramount function is to maintain separation between two fluids; rupture of the exchange surface could be fatal. The exchange surface may be as thin as is compatible with this stipulation. 3. The concentration gradient across the exchange surface. No exchange will take place unless there is a gradient across the exchange surface, i.e. unless the concentrations of the exchange substances in the donor and acceptor fluids are different. In a static system the exchange substances would flow from higher concentrations to lower until the con- centrations were equal when flow would cease. It is therefore necessary to have a current system to maintain the gradient. The most efficient system is a counter-current system (text-fig. 22). In a closed (internal) current system (text-fig. 24) an animal has control over the composition of the fluids circulating. In an open current system (text-fig. 24) however, devices to prevent recirculation of exhausted fluids and choking by extraneous particles are desirable. Thus the paradigm of an exchange system will have a large area of exchange surface which is as thin as is compatible with its strength and a counter-current system (with provisions to prevent recirculation and choking where necessary). This paradigm will fit any exchange system such as a radiator (heat exchange), a kidney (urea excretion) or a gill (oxygen and carbon dioxide diffusion). It has not been developed purely for the present analysis. C. DETAILED FUNCTIONAL INTERPRETATIONS From a purely functional point of view there are two types of cystoid pore-structures: dichoporite and non-dichoporite. In dichoporite pore-structures the current which flowed through the thecal canals was the donor current (for oxygen) and formed part of an 710 PALAEONTOLOGY, VOLUME 11 open current system. The exchange surfaces were within the thecal cavity (endothecal) and protected from mechanical damage. In this section the detailed morphology of dichoporite pore-structures (pectinirhombs and cryptorhombs) will be compared with the paradigm of an exchange system and estimates of the relative efficiencies of the various types of rhomb made. 22 D 100 0 100-)*90 T 80 4 70 60 1 40 ->30 20 -> 10 l 10 -• 0 t 100- • 40 t ,30 1 ■20 t 10 24 text-figs. 22-4. Current and exchange systems. 22. A counter-current system. In the ideal case all the exchange substance is transferred from the donor to the acceptor current. 23. A system with both currents flowing in the same direction. The maximum potential exchange is 50%. 24. Diagram to illustrate closed and open current systems. In the donor current there is a potential danger of the entrance becoming choked with particles and of re-circulation from the exit. A, acceptor current; D, donor current; ES, exchange surface. Figures represent concentrations of exchange substance; heavy arrows represent exchange; light arrows represent current directions. 1. The area of the exchange surface. An increase in the area of the exchange surface means an increase in the amount of exchange. In the simplest hypothetical case exchange would take place through the thecal wall. Oxygen require- ments would be proportional to the volume, and rate of exchange to the surface area, of the theca. Throughout growth oxygen requirements would increase faster than the amount of exchange, thus limiting maximum size. Evagi- nations or invaginations of the thecal wall can counteract this effect without involving a change of gross thecal shape. Both are found in the Glyptocystitida. Maerocystella Callaway, the earliest known form, has externally ridged thecal plates (text-fig. 25). The ridges are formed by folds in the plates and are evaginations of the thecal wall. They increased the area, and also the strength, of the plates which are uniformly very thin (OT mm. in M. mariae Callaway). Exchange probably took place through the entire area of the thecal wall in Maerocystella. Pectinirhombs first appear in Cheirocrinus Eichwald, a direct descendent of Macro - cy Stella (see Paul 1968). Dichopores are invaginations of the thecal wall which allow a differentiation of function between the thicker thecal plates (rarely if ever less than 0-5 mm.) and the thin dichopore walls (usually 0-01 mm.). Exchange was probably restricted text-fig. 25. An isolated thecal plate of Maerocystella mariae Callaway to show external ridges which are effectively evaginations of the thecal wall. PAUL: DICHOPOR1TE PORE-STRUCTURES IN CYSTOIDS 711 to the dichopore walls. These allow a large area for exchange within a small area of thecal surface. In Cheirocrinus granulatus (Jaekel) RM. Ec5384 (PI. 135, figs. 3-5) one half-rhomb has a ratio of exchange area to thecal surface area of 7-84 : 1 . This ratio indicates the relative exchange efficiencies of different types of rhomb. In dichoporite rhombs the closer the spacing of the dichopores and the greater their length and depth, the higher the ratio will be. Available measurements of spacing are sum- marized in Table 1. Dichopore width (W) would probably be given by W = 2 (h+1), where h = height of lining epithelial cells and 1 = length of cilia attached to the epi- thelial cells. At the limit of closeness of spacing the dichopores would be evenly spaced and the dichopore and inter-dichopore widths would be equal. This is the case in pectini- rhombs with confluent dichopores and in complete cryptorhombs but not always in pectinirhombs with discrete dichopores nor in incomplete cryptorhombs. These latter were less efficient and tend to have higher values for spacing (Table 1). Large pectinirhombs tend to have depressed outlines, a large number of dichopores, and to occur in large thecae. In any exchange system there is a distance within which all the exchange substance will be removed from the donor current as it flows along the exchange surface. This limits the length of a dichopore and the longest dichopores of both compressed and depressed pectinirhombs are about the same length (7-9 mm.). The exchange area of a pectinirhomb or cryptorhomb may be more efficiently increased by the addition of new dichopores than by the lengthening of existing dichopores. This may partly account for the tendency of large pectinirhombs to have depressed outlines (i.e. width greater than length). Maximum dichopore lengths in cryptorhombs are greater than those of pectinirhombs and oxygen exchange may have been less rapid in the former. To summarize, the structure of dichoporite rhombs compares closely with the para- digm of an exchange system as regards the area of the exchange surface. Pectinirhombs with confluent dichopores were relatively more efficient than those with discrete dicho- pores in terms of exchange area. 2. The resistance to exchange of the exchange surface. For maximum efficiency an exchange surface should be as thin as its strength will allow. Dichopore walls vary in thickness from less than 0-01 mm. up to 0-03 mm. in rare cases, whereas the thecal wall is 0T mm. thick in Macrocystel/a mariae Callaway, rarely, if ever, less than 0-5 mm. in Cheirocrinus and many Glyptocystitida have thecal walls 2-3 mm. thick. Dichopore walls are very much thinner than other skeletal elements and pectinirhombs were poten- tially weak areas of the theca. The morphology of dichopores, as thin tubes or isoclinal folds, allows great strength and rigidity for a thin surface. It also allows a large surface area without increasing the volume of the theca and thus satisfies both this and the pre- vious requirements. Pectinirhombs with conjunct slits were particularly susceptible to fracture or distor- tion by forces acting on the surface of the theca. It was essential to maintain rigidity for normal functioning and this could be achieved by solid ridges in the positions shown in text-fig. 26. Such ridges impart both strength and rigidity. Ridges are found in the positions suggested in some cystoids with conjunct pectinirhombs. In species of Cheiro- crinus with discrete dichopores the ridges are often incorporated in the plate ‘ornament’ of radiating ridges. These ridges strengthened the plates in general and may only 712 PALAEONTOLOGY, VOLUME 11 table 1 . Spacing of Dichopores in Pectinirhombs. The values for spacing are determined by measuring 10 dichopores and 10 inter-dichopore spaces and expressing the average as 1 slit per O.xxx mm. In pectinirhombs with fewer than 10 dichopores the average was based on the maximum number of dichopores measurable. The averages for each type of pectinirhomb are based on the total number of measurements (No.) Spacing No. of Species Average Range measurements A. Conjunct pectinirhombs with discrete dichopores Cheirocrinus sp. 0-316 0-217-0-415 2 C. giganteus Leuchtenberg 0-273 1 C. granulatus (Jaekel) 0-253 1 C. cf. atavus (Jaekel) 0 106 1 C. languedocianus Thoral 0-370 0 350-0390 4 B. Conjunct pectinirhombs with confluent dichopores Homocystites alter Barrande 0-215 0205-0-225 2 H. constrictus (Bather) 0-2025 0-201-0-204 2 Plearocystites elegans Billings. 0196 1 P. filitextus Billings 0-241 1 P. cf. rugeri Salter 0-20 approx. 1 P. squamosus Billings 0-171 1 C. Disjunct pectinirhombs with discrete dichopores Cheirocrinus sp. 0174 1 C. anatiformis (Hall) 0-257 0-25-0-27 4 C. forbesi Billings 0-273 0-25-0-29 6 Echinoencrinites angulosus Pander 0-27 0-21-0-31 3 E. senckenbergii von Meyer 0-33 0-32-0-34 3 E. reticulatus Jaekel 0-427 0-41-0-44 3 Erinocystis sp. 0-38 1 D. Glyptocystites type Glyptocystites ehlersi Kesling 0-197 0-19-0-20 3 G. multiporus Billings 0-193 0-18-0-20 10 G. regnelli Sinclair 0-206 0-20-0-22 5 G. batheri Sinclair 0-225 0-205-0-245 7 E. Disjunct pectinirhombs with confluent Sphaerocystites multifasciatus Hall dichopores 0-223 0-212-0-225 4 Jaekelocystis hartleyi Schuchert 0-168 0-150-0 180 4 Lepadocystis moorei (Meek) 0 169 0-165-0-175 6 Apiocystites pentrematoides Forbes 0-208 0-200-0-214 3 Strobilocystites calvini White 0-276 0-265-0-285 6 Lovenicystis angelini (Haeckel) 0-218 0-201-0-236 2 Lipsanocystis traversensis 0-152 0 151-0 154 14 Ehlers and Leighley Tetracystis oblongus (Forbes) 0-156 0 153-0 159 3 Lepocrinites gebhardii Conrad 0-250 2 Pseudocrinites gordoni Schuchert 0-167 0-160-0 174 6 P. bifasciatus Pearce 0-177 0 168-0 187 3 P. pyriformis Paul 0-257 0-233-0-272 3 Staurocystis quadrifasciatus (Pearce) 0-193 0-190-0-201 6 Glansicystis baccata (Forbes) 0205 0 190-0-220 6 Over -all averages Range No. Type A 0-306 0-106-0-415 9 B 0-205 0-171-0-241 8 C 0-300 0-174-0-440 21 D 0-205 0-193-0-245 25 E 0-192 0-151-0-285 68 PAUL: DICHOPORITE PORE-STRUCTURES IN CYSTOIDS 713 incidentally have strengthened the pectinirhombs. In Homocystites constrictus (Bather) (PI. 134, fig. 5) and Pleurocystites spp. (PI. 134, figs. 7, 8) rims occur which are indepen- dent of the plate ridges and whose strengthening effect was confined to the area immedi- ately adjacent to the pectinirhombs. In fully developed disjunct pectinirhombs the intra-rhomb areas between the slits would resist forces acting on the thecal surface. In disjunct pectinirhombs with discrete dichopores the intra-rhomb areas tend to be large (PI. 135, figs. 1, 9) and this partly restores the strength of the rhomb-bearing plates. In cryptorhombs (PI. 138,fig. 4; PI. 139, fig. 1), with only rows of pores, this restoration is virtually complete. In disjunct pectini- rhombs with confluent dichopores the intra-rhomb areas tend to be smaller but the text-fig. 26. A conjunct pectinirhomb with ridges which deflect forces acting within the plate and impart rigidity to the pectinirhomb. adsutural portions of the vestibule rims develop within them (PI. 136, fig. 3; PI. 137, fig. 1). During growth of a disjunct pectinirhomb some marginal slits were conjunct and subject to the same weakness as conjunct pectinirhombs. In disjunct pectinirhombs with confluent dichopores the absutural portions of vestibule rims arise early in develop- ment in exactly the position postulated to strengthen conjunct pectinirhombs (PI. 137, fig. 1). No such rims are present in disjunct pectinirhombs with discrete dichopores. Uncalcified dichopores would allow a greater exchange area than calcified dicho- pores since only the soft tissues function as exchange surface. In pectinirhombs distortion or disarticulation would probably occur if the dichopores were uncalcified but in cryptorhombs strength is provided by the covering of thecal plate. It is still essential to have rigid dichopore walls to prevent the dichopores themselves from dis- torting (under pressure differences or due to gravity) since this would interrupt the func- tioning of the cryptorhomb. A maximum movement of only 0-05 mm. by two adjacent dichopore walls would completely close a dichopore or inter-dichopore space. Calcifica- tion of dichopores imparts rigidity at the expense of the available area of exchange. Most pectinirhombs and cryptorhombs have strengthening structures closely associ- ated with them indicating they were areas of weakness. The conclusion that this weak- ness was mainly due to the dichopore walls seems inescapable. Dichopore walls were not 714 PALAEONTOLOGY, VOLUME 11 only thin absolutely but relatively so thin that the limits of their strength were reached and strengthening devices became necessary to counteract this weakness. In this respect dichoporite pore-structures agree closely with the paradigm of an exchange system. Further conclusions are that conjunct pectinirhombs were intrinsically weaker than disjunct pectinirhombs and all pectinirhombs were weaker than cryptorhombs. Streng- thening devices in pectinirhombs with confluent dichopores apparently allowed a weaker structure with a better exchange area to thecal surface area ratio. Disjunct pec- tinirhombs with confluent dichopores were strengthened throughout a longer period of their growth than those with discrete dichopores. 3. Protective devices and devices to prevent recirculation. The currents which flowed within dichopores formed part of an open current system. Therefore devices to prevent choking by extraneous particles and recirculation of deoxygenated sea-water would have been beneficial to exchange. Such devices are also very useful in determining current directions since they indicate which pores were entrances and which exits. Clearly it is more efficient to modify the entrance (rather than the exit) with a protective device since this will prevent entry of harmful particles. Equally, exhausted sea-water in the vicinity of the entrance will be sucked in by the in-current. Recirculation can only be prevented by modifying the exit to direct the out-current away from the entrance and into the ambient sea-water for remixing. Protective devices, which will be considered first, indicate entrances and devices to prevent recirculation indicate exits. A protective device is a structure which reduces the maximum size of particles that may pass through an aperture without interfering with the functioning of the aperture. There are only three classes of structure which fulfil these requirements : meshes, grilles, and narrow slits. Paradigms for all three have been defined by Rudwick (1961, meshes and grilles; 1964a, slits). The maximum particle which can pass through the device is termed the critical particle and is of critical diameter. In an ideal protective device the critical diameter is the same at all points within the aperture. Two types of harmful particle may be recognized. Passive particles are any particles in suspension (animal or plant debris, sediment, etc.) which may choke the dichopores but not attack the cystoid. Active particles are organisms which may attack the cystoid EXPLANATION OF PLATE 134 Stereophotos of conjunct pectinirhombs A. with discrete dichopores. Fig. 1. Cheirocrinus giganteus Leuchtenberg. BMNH E16130, basal rhombs. Figs. 2, 4 Cheirocrinus sp. SM A3 138 (fig. 4), SM A3 139 (fig. 2). External moulds of isolated plates. Fig. 3. Cheirocrinus granulatus (Jaekel). RM Ec5384. Fig. 9. Cheirocrinus languedocianus Thoral. Ubaghs coll, latex impression showing incomplete pec- tinirhombs and demirhombs in radial and lateral plates. B. with confluent dichopores. Fig. 5. Homocystites constrictus (Bather). HM 3541a, latex impression showing rhombs B2:IL2, LI :L2, and R1 :R2. Note the rim on L2 (middle rhomb). Fig. 6. Pleurocystites rugeri Salter. LU 3175, latex impression of L3 :L4. Fig. 7. Pleurocystites elegans Billings. SM A53059. L3:L4. Fig. 8. Pleurocystites filitextus Billings, BMNH E16047. L4:L3. Figs. 1, 8, 9 x 2, figs. 2-7 x 3. All whitened with ammonium chloride sublimate. Palaeontology , Vol. 11 PLATE 134 PAUL, Cystoid pore structures PAUL: DICHOPORITE PORE-STRUCTURES IN CYSTOIDS 715 by boring into it, attaching to it or browsing off cilia, etc. Dichopores may be regarded as long thin tubes whose critical diameter at any given point is the minimum width. Complete protection against passive particles may be achieved if the critical diameter of the entrance is less than that of any other point along the tube. Any spherical particle capable of entering the tube will pass right through without becoming lodged. Complete protection against active particles, which may be able to swim against cur- rents, can only be achieved if the critical diameters of both the entrance and the exit are less than the diameter of the smallest active particle. As nothing is known about table 2. Ratios of slit lengths and areas in disjunct pectinirhombs Ratios Species Rhomb Lengths longer slit shorter slit Areas longer slit shorter slit A. Discrete dichopores Echinoencrinites angulosus BMNH E29102 B2 IL1 1-27 (7) 0-82 (5) B2 IL2 I 21 (11) 0-84(11) E. reticulatus RM Ec5515 B2 IL1 1-25 (6) 0-94 (6) B2 IL2 1-27 (8) 1-31 (8) L4 R3 Ml (9) 0-94 (9) E. senckenbergii BMNH E29095 B2 IL1 1-54 (8) 0-97 (8) B2 IL2 1 43 (6) 1-06(6) L4 R3 116 (10) 0-87 (10) Cheirocrinus anatiformis R2 R3 M4(15) 1-08 (15) B. Confluent dichopores Glyptocystites multiporus LI L2 1 -46 (9) 1-46 (9) Staurocystis quadrifasciatus LI R5 1-37 08) 0-97 (18) Lovenicystis angelini RM Ec5053 LI R5 1-77 (13) 1 -38 (13) RM Ec5066 LI R5 1-30 (17) 0-98 (17) Glansicystis baccata Bu Ho55 LI R5 115 (9) 1 09 (9) L4 R3 1-43 (9) 1 19 (9) GSM 7380 LI R5 2-63 (8) 1 91 (8) Figures in brackets represent the number of measurements upon which each ratio was based. active particles it is impossible to investigate protection against them. Only passive particles are considered below. The dichopores of pectinirhombs open as slits and if protected the entrances should form protective slits. In conjunct pectinirhombs each slit runs the entire length of the dichopore and acts as both entrance and exit. In Cheirocrinus latavus (Jaekel) BMNH E23515 and Pleurocystites filitextus Billings BMNH E7600b (PI. 140, fig. 2) slit widths are slightly less than dichopore widths within the theca. In these cases the slits were protective. In disjunct pectinirhombs the slits of the two half-rhombs were of different lengths. In many cases the longer slits are also narrower. To satisfy the requirements of a protec- tive device the critical diameter of the incurrent slit must be reduced without reducing current flow. If protection is achieved by reducing the width, constant current velocity may be maintained by a corresponding increase in slit length since the velocity is inversely proportional to the cross-sectional area of the slit. In the ideal case the ratio of entrance to exit slit length will be greater than unity but the ratio of the areas will equal unity. Available measurements are shown in Table 2. In cases where the longer 716 PALAEONTOLOGY, VOLUME 11 slits were also narrower they could have been protective. In cases where the slit widths are equal they could still have been protective if both were narrower than the dichopore width within the theca. The pectinirhombs of Lipsanocystis magnus Stumm (PI. 140, fig. 4) show microscopic bars across the slits of both half-rhombs. These bars reduce the critical diameter of both apertures. Preservation is not quite adequate to determine which slits are better pro- tected or if any difference exists. The formation of these microscopic grilles may be unique to L. magnus; perhaps a response to its muddy environment or to active particles since both apertures are modified. Alternatively it may be that the preservation of most glyptocystitids is not adequate for these delicate structures to be detectable. In the latter case the arguments for protective devices in pectinirhombs may need some modification. table 3. Critical diameters in pectinirhombs and cryptorhombs Entrance Exit Rhomb type Pectinirhombs Average Range No. Average Range conjunct discrete 0 115 0 02-0- 19 6 conjunct confluent 0066 005-008 8 disjunct discrete 0 113 005-0 18 10 0-162 0-12-0-20 confluent disjunct 0075 0-07-0-095 5 0-104 0-09-0012 Cryptorhombs 0058 0-04-0-07 13 0-20 010-0-35 No. = Number of rhombs of each type measured. In pectinirhombs with variable slit widths the average width of all slits was calculated. In cryptorhombs the dichopores open as pores, at least one of which is sieve-like, being composed of a cluster of fine pores. If sieve-pores were protective they formed protective meshes. Measurements given in Table 3 show that the critical diameters of sieve-pores are much less than those of simple pores. The areas (Table 4) of simple and sieve-pores are often approximately equal however. Again protection could be provided while interfering to the least extent possible with current flow. Particles prevented from entering dichopores by protective devices would choke the slits and pores themselves unless removed. Clearance could be achieved by surface ciliary currents, the particles being carried in the same direction as the currents within the EXPLANATION OF PLATE 135 Stereophotos of disjunct pectinirhombs A. with discrete dichopores. Fig. 1. Cheirocrinus sp. nov. SM A15966. Latex impression of isolated plate. Figs. 2-5. Cheirocrinus granulatus (Jaekel). RM Ec5384. Fig. 2, internal surface of basal plate; figs. 3-5, views of isolated half-rhomb to show relationship between dichopores and slits. Fig. 6. Cheirocrinus sp. BMNH El 5983. Fig. 7. Echinoencrinites senckenbergii von Meyer. BMNF1 E29102, rhombs B2:IL1 and B2:IL2. Fig. 8. Echinoencrinites reticulatus Jaekel. BMNH E29095. B2:IL1 and B2:1L2. Figs. 9, 12. Cheirocrinus anatiformis (Hall). GSC. B2:IL2 (fig. 9) and radials (fig. 12). Fig. 10. Echinoencrinites reticulatus Jaekel. RM Ec5515. B2 : 1 L 1 . B. with confluent dichopores. Fig. 11. Glyptocystites multiporus Billings. GSC 1387g (Lectotype). B2:IL1, B2:IL2, and R1 : R2. Note the adsutural rim on R2 (large upper rhomb). Figs. 1, 6-9, 11, 12 x2, figs. 2-5, X 3, fig. 10, x5. All whitened with ammonium chloride sublimate. Palaeontology, Vol. 11 PLATE 135 PAUL, Cystoid pore structures PAUL: DICHOPORITE PORE-STRUCTURES IN CYSTOIDS 717 table 4. Areas of simple and sieve-pores in cryptorhombs Area of sieve- Area of simple Specimen pore ( entrance ) pore {exit) Hemicosmites sp. RM Ec5364 0 0707 sq. mm. 0 0471 sq. mm. H. extraneus RM Ec5283 00226 00201 H. sp. nov. RM Ec2280 0 01 54 (radials) 00314—00572 0 0502 (laterals) H. cf. verrusosus BMNH El 5994 0 0491 0 0433 00201- H. malum RM Ec5517 0 0201 (basals) 0 0402 (laterals) 00254 H. cf. pyriformis BMNH E7592 00314 00227 Caryocrinites ornatus SM A50958* 00308 00154 C. roemeri BMNH E29105* 00628 00628 C. septentrionalis RM Ec25493 00226 00314 Thomacystis tuberculata BMNH E16300 00176 00176 Unless otherwise stated the area of sieve-pores was taken from lateral plates: all areas of simple pores are from infra-lateral plates. * These specimens have sieve-pores at both ends of dichopores in complete cryptorhombs. However, there are fewer larger pores per cluster where one would normally expect simple pores. text-fig. 27. Supposed paths of extraneous particles in conjunct pectinirhombs. Solid arrows indicate external currents; dashed arrows internal currents. dichopores. In conjunct pectinirhombs particles strained from the in-current would be passed along the slit until they reached the out-current and would then be carried away from the theca (text-fig. 27). Similar currents could have been present in disjunct pec- tinirhombs and cryptorhombs but the slit, vestibule, and pore rims would render them less effective (text-figs. 28, 29). We may conclude that the narrower slits of disjunct pectinirhombs and the sieve- pores of cryptorhombs reduced the critical diameters of these apertures, which were 718 PALAEONTOLOGY, VOLUME 11 text-figs. 28, 29. Possible paths of extraneous particles in disjunct pectinirhombs. Note that the slit and vestibule rims make it more difficult for the excurrent to remove the particles. PAUL: DICHOPORITE PORE-STRUCTURES IN CYSTOIDS 719 therefore entrances. Absolute critical diameters shown in Table 3 indicate that crypto- rhombs were better protected than pectinirhombs. However the latter, particularly con- junct pectinirhombs, were better able to prevent the protective devices themselves from becoming choked. This may account for the occurrence of conjunct pectinirhombs in all but one rhomb-bearing member of the Pleurocystitidae. Members of this family were probably vagrant with the theca resting on the substrate (Paul 1967c) and may have encountered many more particles than more typically pelmatozoan cystoids. For maximum efficiency devices to prevent recirculation of exhausted sea-water must do so while interfering with current flow to the least possible extent. The most funda- mental requirement is to have two apertures and to keep them apart. How far apart will depend on the volume, velocity, and direction of the out-current. Even if the entrances were capable of detecting oxygen-depleted water they could not prevent it from entering without arresting or reversing the currents. Recirculation can only be prevented effec- tively by modifying the exit to direct the out-current away from the entrance. Recirculation may occur within a single rhomb or from one rhomb into another. In a theca with many rhombs recirculation of the second type may be prevented by group- ing exits together. In general where two or more apertures are found within a single plate they are either exits or entrances (text-figs. 30, 31). Recirculation within a single rhomb may be prevented by mechanical barriers or by devices to project the out-current through the boundary layer into the ambient sea-water for remixing. The boundary layer is the layer of fluid immediately adjacent to the surface of an immersed solid, within which viscosity is the most important single force controlling flow. Cilia beat within the boundary layer and without them water in contact with the external surface of an echinoderm would remain unchanged in all but the most turbulent seas. Water emerging from the dichopores would be recirculated by surface cilia unless forced through this layer. Mechanical barriers between the entrance and exit were therefore probably less efficient than were devices to project the out-current through the boundary layer. The best method to determine the effects of such structures is to construct models and test them under simulated conditions as close as possible to those under which they operated in life (cf. Rudwick 1961, Jefferies and Minton 1965). Time has not been avail- able to do this and the conclusions outlined below are tentative. The conjunct pectinirhombs of Homocystites constrictus (Bather) (PI. 134, fig. 5) have distinct rims on one half-rhomb only. Water passing along the dichopores towards these rims would be deflected away from the theca. Where two half-rhombs are developed in a single plate they are either rimmed or unrimmed. Current directions shown in text- fig. 30 are derived from the presence of these rims. Sinclair (1948) suggested that the adsutural ridges on one half-rhomb of each pectini- rhomb in Glyptocystites spp. were mechanical barriers to recirculation. This suggestion seems likely although as yet without practical confirmation. Again where two or more half-rhombs occur in one plate they are either with or without ridges. In text-fig. 31 the arrows point to the plates with ridges. Disjunct pectinirhombs with confluent dichopores have a closed vestibule rim on one half-rhomb and an open rim on the other. The absutural portion of the closed rim would have directed the out-current away from the theca. The adsutural portion would have acted as a mechanical barrier to recirculation as Sinclair suggested for the pectinirhombs of Glyptocystites. In addition the closed vestibule rims united all the currents of the 720 PALAEONTOLOGY, VOLUME 11 constituent dichopores and probably reinforced them. If out-currents they would have passed through the boundary layer more readily. The pectinirhombs of Jaekelocystis hartleyi Schuchert (PI. 137, figs. 7, 9, 10) have closed vestibule rims which form a pore 0-5 mm. in diameter. This pore would considerably increase the velocity of currents passing through it. In disjunct pectinirhombs with a constant slit width in both half- rhombs, the shorter slits would have had faster currents. Fast in-currents would be more likely to suck in extraneous particles but fast out-currents would be more likely to pass through the boundary layer and disperse. Shorter slits and closed vestibule rims therefore probably indicate exits. text-figs. 30, 31. Current patterns in Cheirocrinus and G/yptocystites. 30. Diagram to show the arrangement of thecal plates and pectinirhombs in Cheirocrinus anatiformis (Hall). 31. Diagram to show the arrangement of thecal plates and pectinirhombs in Glyptocystites ehlersi Kesling. Arrows indicate supposed current directions in the pectinirhombs. Note where more than one half-rhomb is developed within a single plate they are usually of the same type (i.e. either entrance or exit). Plate outlines based on Kesling 1962 (text-fig. 30) and 1961 (text-fig. 31). Many cryptorhombs have simple pores developed beside or within ridges on the plate surface. These ridges could have directed currents away from the theca. In Caryocrinites septentrionalis Regnell (PI. 139, fig. 5) the ridges beside simple pores are more strongly developed than those associated with sieve-pores. In Hemicosmites spp. (PI. 138, figs. 1, 5) simple pores are developed into chimney-like tubercles and may be interpreted as chimneys. The further away from the thecal surface the out-current emerges the less likely it is to be recirculated by surface ciliary currents. These were therefore probably exits. EXPLANATION OF PLATE 136 Stereophotos of disjunct pectinirhombs with confluent dichopores Figs. 1, 7. Pseudocrinites gordoni Schuchert. BMNH E23122. LI :R5 (fig. 1) and B2:IL2(fig. 7). Note the vestibule rims typical of this type of pectinirhomb. Fig. 2. Cal/ocystites jewetti Hall. GSC 14686. L4:R3. Fig. 3. Pseudocrinites pyriformis Paul. BU HolO. LI :R5. Figs. 4, 5. Lovenicystis angelini (Jaekel). RM Ec5066. L4:R3 (fig. 4) and LI :R5 (fig. 5). Fig. 6. Strobilocystites calvini White. UI. L4:R3. Fig. 8. Sphaerocystites multifasciata Hall. BMNH 22960. B2:IL2. Fig. 9. Pseudocrinites perdewi Schuchert. USNM 35072. L4:R3. Figs. 1, 3-5, 7, 8 X 3, figs. 2, 6, 9 X 2. All whitened with ammonium chloride sublimate. Palaeontology, Vol. 11 PLATE 136 PAUL, Cystoid pore structures PAUL: DICHOPORITE PORE-STRUCTURES IN CYSTOIDS 721 To conclude, devices which prevented recirculation are apparently present in most cryptorhombs and disjunct pectinirhombs but are generally absent from conjunct pectinirhombs. Suggested entrances were described under protective devices and exits under devices to prevent recirculation. In all cases these interpretations should be, and in fact are, complementary. Thus in disjunct pectinirhombs longer slits were interpreted as entrances and shorter slits as exits. In cryptorhombs sieve-pores were interpreted as entrances and simple pores as exits. Closed vestibule rims in pectinirhombs were inter- preted as being associated with exits and they always occur on the half-rhombs with shorter slits. All these interpretations confirm each other and also confirm the presence and direction of donor currents within dichoporite pore-structures. 4. The maintenance of a concentration gradient. A counter-current system is the most efficient method of maintaining a concentration gradient. Ciliary currents within the dichopores would be expected on a priori grounds as suggested by Sinclair (1948). The viscous effect of the boundary layer would be even more marked in a restricted space such as a dichopore. No matter how turbulent the seas, water within the dichopores would not be disturbed without cilia. text-fig. 32. Graph to show relationship between slit length and dichopore length in pectinirhomb LI :R5 of Stauroeystis quadrifasciatus (Pearce), BU Ho 1. Lower scale reversed about the rhomb axis for clarity. Note that the slit lengths are proportional to the dichopore lengths and the slit lengths in LI are constantly less than those of R5. If currents were present, sea-water flowed within the dichopores forming the donor current which was part of an open current system. The best evidence for the presence and direction of the donor current has been outlined in the preceding section. Additional evidence is found in the relationship between slit and dichopore lengths in pectini- rhombs and between pore counts and dichopore length in cryptorhombs. Adaxial dichopores are longer and deeper than marginal dichopores. If the current velocity were constant in all dichopores more water would pass through the larger dichopores per unit time. The capacity of the dichopore would be proportional to its cross-sectional area and, since the width is constant, capacity would be proportional to dichopore 722 PALAEONTOLOGY, VOLUME 11 depth. For constant velocities of in-current and out-current across a rhomb, slit length must be proportional to dichopore capacity. Dichopore depth is not readily measurable but in some instances e.g. Staurocystis quadrifasciatus (Pearce) (PI. 137, fig. 1, text-fig. 32) slit length is almost exactly proportional to dichopore length. In the cryptorhombs of Caryocrinites ornatus Say ( PI . 1 38, fig. 3) the number of pores per sieve-pore cluster increases adaxially as does the total area of the entrance. The morphology of these rhombs would allow constant current velocities across the rhombs and within all dichopores, if currents were present. Ciliary currents are likely to have been of similar velocities in all dichopores. It is more difficult to confirm the presence and direction of internal acceptor currents as they formed a closed current system, but there is some indirect evidence. In rhombs with closely spaced dichopores the inter-dichopore spaces were entirely within the boundary layer. Ciliary currents between dichopores would be expected on the same a priori grounds used to postulate their presence within dichopores. The width of dicho- pores was probably controlled by the combined thicknesses of ciliated epithelium, and lengths of cilia, lining them. If the dichopore width were much greater, dead water would accumulate within the dichopore. Alternatively cilia on opposite sides of the dichopore would interfere with each other if the dichopores were thinner. Where inter-dichopore widths equal dichopore widths both were probably controlled by the presence of cilia. This suggests that internal ciliated currents were present. The most efficient internal current is a counter-current. Internal currents within a rhomb could only act efficiently if the external currents in all the dichopores were counter to them. With very rare exceptions in some incomplete cryptorhombs, current directions in all dichopores of one rhomb were apparently the same. Currents in all dichopores could have been counter to an internal current. EXPLANATION OF PLATE 137 Stereophotos of disjunct pectinirhombs with confluent dichopores Fig. 1. Staurocystis quadrifasciatus (Pearce). BU Hoi. LI :R5. Figs. 2, 4. G/ausicystis baccata (Forbes), typical form. GSM 7380. B2:IL2 (fig. 2) and LI :R5 (fig. 4). Fig. 3. Lipsanocystis traversensis Ehlers and Leighley. UMMP 56222. L4:R3. Fig. 5. Schizocystis armata (Forbes). GSM 7381. B2:IL2. Figs. 6, 8. Glansicystis baccata (Forbes), common form. BU Ho55. L4:R3 (fig. 6) and LI :R5 (fig. 8). Figs. 7, 9, 10. Jaekelocystis hartleyi Schuchert. BMNH E23121. L4:R3 (fig. 7), L1:R5 (fig. 9), and B2:IL2 (fig. 10). Note that the vestibules in IL2, LI, and L4 are small pores. Figs. 1, 2, 4-10 X 3, fig. 3x2. All whitened with ammonium chloride sublimate. EXPLANATION OF PLATE 138 Stereophotos of cryptorhombs Fig. 1. Hemicosmites sp. RM Ec5364. Infra-lateral with simple pores. Fig. 2. Hemicosmites extraneus Eichwald. RM Ec5283. Laterals with simple pores and radials with sieve-pores. Figs. 3, 6. Caryocrinites ornatus Say. BMNH E29091. Internal (fig. 6) and external (fig. 3) views of isolated lateral plate. Note that the number of individual pores in a cluster increases towards centre of plate (fig. 3). Fig. 5. Hemicosmites cf. verrucosus Eichwald. BMNH El 5994. Infra-laterals and laterals. Note chimney- like simple pores. Figs. 4, 7. Caryocrinites ornatus Say. BMNH E29084. Internal (fig. 7) and external (fig. 4) views of isolated infra-lateral plate. Note that only simple pores occur in the infra-lateral. Figs. 1,2 x 3, figs. 3-7 x 2. All whitened with ammonium chloride sublimate. Palaeontology, Vol. 11 PLATE 137 PAUL, Cystoid pore structures Palaeontology, Vol. 11 PLATE 138 PAUL, Cystoid pore structures PAUL: DICHOPORITE PORE-STRUCTURES IN CYSTOIDS 723 In any cystoid relying on diffusion alone to distribute oxygenated fluids internally it would be advantageous to have rhombs over the entire external surface as all internal organs would be equidistant from an oxygen source. An approximation to this distribu- tion of rhombs is found in Cheirocrinus and Glyptocystites (pectinirhombs) and in most cryptorhomb bearing genera. As the latter grew, relatively large areas without crypto- rhombs were developed adjacent to the intra-circlet sutures of the infra-lateral and lateral text-figs. 33-6. Current directions in cystoids with three pectinirhombs. 33. Staurocystis quadrifasciatus (Pearce). 35. Pleurocystites sp. 34. Glansicystis baccatus (Forbes). 36. Erinocystis sp. Note the current directions in all four have the same basic pattern. Plate outlines after Paul 1967a (text-figs. 33-4), Breimer 1963 (text-fig. 35), and Jaekel 1899 (text-fig. 36). circlets. It is within just these areas that incomplete cryptorhombs develop relatively late in ontogeny. The distribution and development of cryptorhombs, and of pectini- rhombs in Cheirocrinus and Glyptocystites , suggests that internal circulations were inefficient or lacking. Cystoids with only a few rhombs probably required some sort of internal circulation. Independent evidence for this is found in the Callocystitinae (Paul 19676). In all cystoids with few rhombs the same external current pattern is found (text-figs. 33-6). This could reflect a fixed internal circulation but only if there were a constant relationship between donor and acceptor current directions. The most efficient relationship is a counter-current system. 3 B C 6055 724 PALAEONTOLOGY, VOLUME 11 As regards maintenance of a concentration gradient, there is considerable direct evidence for the presence of external currents within dichopores and some indirect evidence for internal counter-currents between dichopores. Again dichoporite pore- structures agree closely with the paradigm of an exchange system in this respect. In general most details of dichoporite pore-structures agree remarkably well with the paradigm of an exchange system and would have been beneficial to exchange. Oxygen, and corresponding carbon dioxide, diffusion is the most likely form of this exchange and hence it seems that pectinirhombs and cryptorhombs were most probably respiratory organs. D. EVOLUTIONARY AND TAXONOMIC IMPLICATIONS The distribution of different types of pectinirhomb within the Glyptocystitida is summarized in text-fig. 37. Throughout the detailed functional analysis it was suggested that conjunct pectinirhombs were inherently less efficient than disjunct pectinirhombs (except in clearing the slits of extraneous particles) and that discrete dichopores were less efficient than confluent dichopores. With the possible exception of Scoliocystis Jaekel, conjunct pectinirhombs are con- fined to the Cheirocrinidae and Pleurocystitidae. Their presence in the early (Arenig, L. Ordovician) cheirocrinids is probably primary. The appearance of conjunct pectini- rhombs with confluent dichopores in later cheirocrinids like Homocystites constrictus EXPLANATION OF PLATE 139 Stereophotos of cryptorhombs Fig. 1. Hemicosmites malum Pander. RM Ec5517. Note the ‘hidden rhombs’. Fig. 2. Hemicosmites sp. RM Ec5702. Internal surface of isolated infra-lateral. Figs. 3, 6. Thomacystis tuberculata Paul. BMNH E16300. Complete theca (fig. 3) and external mould of isolated basal plate (fig. 6). Latter shows cast of two dichopores which bifurcate near external surface of plate to form sieve-pore (cf. PI. 140, fig. 5). Fig. 4. Caryocrinites roemeri Jaekel. BMNH E29105. Infra-lateral and lateral plates; both show sieve-pores. Fig. 5. Caryocrinites septentrionalis Regnell. RM Ec25493. Note rims beside simple pores are more strongly developed than those by sieve-pores. Fig. 7. Caryocrinites ornatus Say. SM A50958. Fig. 8. Hemicosmites cf. pyriformis von Buch. BMNH E7592. Figs. 1, 4, 8 x2, figs. 2, 3, 5-7 x 3. All whitened with ammonium chloride sublimate. EXPLANATION OF PLATE 140 Fig. 1. Echinoencrinites angulosus Pander. RM Ec5490, section 9. Vertical section through one dichopore in B2:IL2. Note structure of thecal plate. x35. Fig. 2. Pleurocystites filitextus Billings. BMNH E7600b. Vertical section through B2:IL2 X 35. Fig. 3. Cheirocrinus jamesii (M‘Coy). RSM 1870.12.879. External mould of isolated plate showing multi-disjunct pectinirhomb. X 6. Fig. 4. Lipsanocystis magnus Stumm. UI 32001. Detail of slits in L4:R3 to show microscopic bars across slits. x25. Figs. 5, 6. Hemicosmites sp. nov. Sections through thecal plates to show sieve-pore (fig. 5) and simple pore (fig. 6). X 30 approx. Fig. 7. Strobilocystites calvini White. UI. Sutural view of isolated plate to show dichopores. Photo- graphed under xylol, x 25. Fig. 8. Echinoencrinites angulosus Pander. RM Ec5490, section 3. Tangential section through B2:IL1 x 3. Palaeontology, Vol. 11 PLATE 139 PAUL, Cystoid pore structures 2SS Palaeontology, Vol. 11 PLATE 140 PAUL, Cystoid pore structures PAUL: DICHOPOR1TE PORE-STRUCTURES IN CYSTOIDS 725 pectinirhombs and of confluent and discrete dichopores. 726 PALAEONTOLOGY, VOLUME 11 (Bather) (Ashgill, U. Ordovician), may be due to the mode of life. These pectinirhombs are more highly developed than those of earlier species. The retention of conjunct pectinirhombs in the Pleurocystitidae is probably due to the mode of life as suggested by Paul (1967c). Confluent dichopores are unknown before the Llandeilo (M. Ordovician) while discrete dichopores are unknown after the Ashgill or possibly even the Caradoc (Late Middle Ordovician). Between these time limits there was a change from discrete to confluent dichopores independently in all evolutionary lines. The change may have been gradual in some cases: the pectinirhombs of G/yptocystites Billings are somewhat inter- mediate in character. All Silurian and Devonian glyptocystitids (with the exception of two species of Pleurocystitidae) have disjunct pectinirhombs with confluent dichopores, the most efficient type. Apparently throughout the evolution of the Glyptocystitida there was a progression from less to more efficient types of pectinirhomb. There is no evidence for a gradual change from conjunct to disjunct pectinirhombs but this is not surprising since it is difficult to see, on geometrical grounds, how a transition could exist. At present no estimate of the functional efficiency of multi-disjunct pectinirhombs can be made. It is also much more difficult to compare the relative exchange efficiencies of pectinirhombs and cryptorhombs. The latter show much less variation and as yet no evolutionary trends have been recognized. Within the Glyptocystitida (but not within the Rhombifera as a whole) there is evi- dence of a progressive decrease in the number of pectinirhombs in one theca. Bather (1913) explained this reduction in the Pleurocystitidae as a result of the anus partly taking over respiration. However, this explanation cannot be applied to the Callocysti- tidae, for example, which lack the extensive periproctal membrane supposedly used in respiration per anum. The reduction in the number of pectinirhombs could be due to the development of an internal circulation to distribute more efficiently oxygen gained from the pectinirhombs. Some independent evidence for such a circulation system has been described by Paul (19676). The Glyptocystitida alone show this trend. Rudwick (1961, 1964a) has emphasized that there may be a limited number of different structures which can perform a given function efficiently. In these cases it is inherently likely that the same or similar structures may have evolved independently several times. This was undoubtedly the case with confluent dichopores. It is also undoubtedly the case with rhombs as a whole. Fistulipores and dichopores represent basically different structures and there is no feasible way to evolve one from the other. The fistuliporite and dichoporite rhombs of the Rhombifera were independently acquired and the pos- session of rhombs is a polyphyletic character within the class. The Glyptocystitida and Hemicosmitida (dichoporite) can be separated from the Caryocystitida (fistuliporite) on several characters, e.g. plating arrangement, mode of growth, presence of true stem, besides the type of thecal canal present. The dichoporite superfamilies form natural units but it is still possible that they acquired their rhombs independently. It is easy to envisage that cryptorhombs evolved from disjunct pectinirhombs but there is no evidence at all for such a transition and both types of rhomb appear simultaneously in the Arenig. Furthermore, a rhomb-less ancestor for the Glyptocystitida is known ( Macro - cystella ) which has an identical plate arrangement to later Glyptocystitida but which is quite distinct from all Hemicosmitida. The cystoidea as currently understood (Kesling 1963, 1968) form a polyphyletic group PAUL: DICHOPORITE PORE-STRUCTURES IN CYSTOIDS 727 despite the recognition of the Eocrinoidea, Paracrinoidea, etc. as distinct classes. It is suggested that the Diploporita and Rhombifera should also be regarded as separate classes. Within the Rhombifera two orders, the Dichoporita (Jaekel 1899 emend.) and Fistuliporita nov., may be recognized. The former corresponds to the Pectinirhombi- fera (Delpey 1942, p. 207) but as the distinctive character is the possession of dichopores not pectinirhombs, an emended use of Jaekel’s Dichoporita is preferred. Since its first inception as a class to include all non-crinoid, non-blastoid Pelmatozoa, the Cystoidea has been subject to the removal of constituent groups as their distinctive- ness was realized. This process is here carried to the limit and as neither Diploporita nor Rhombifera can be considered the more typical cystoids, the formal name Cystoidea must fall into disuse. The concept of a ‘cystoid’ is still useful in a general sense similar to that now reserved for ‘Dinosaur’. Cystoid may be used to embrace the Eocrinoidea, Paracrinoidea, Rhombifera, Diploporita, Parablastoidea, Edrioblastoidea, and possibly the Blastoidea s.s. These classes share many common characteristics which separate them from ‘carpoids’ and ‘edrioasteroids’, two other groups once included within the Cystoidea s.I. APPENDIX List of species with dichoporite pore-structures examined. Class Rhombifera Muller Order Dichoporita Jaekel emend. Rhomb type Superfamily Glyptocystitida Bather Pectinirhombs Family Macrocystellidae Bather emend. Jaekel slits dichopores Macrocystel/a mariae Callaway — ■ — M. bohemicus (Barrande) — M. azaisei (Thoral) — ■ Family Cheirocrinidae Jaekel Cheirocrinus sp. (Ngwetung, Burma) C. atavus (Jaekel) C. giganteus (Leuchtenberg) C. languedocicmus Thoral conjunct discrete 55 55 C. gramilatus (Jaekel) conjunct and ,, disjunct C. anatiformis (Hall) C. sp. nov. (Tramore, Eire) C. sp. (Katlino, U.S.S.R.) disjunct 55 ” C. interruptus (Jaekel) C. jamesi (M‘Coy) multi-disjunct unknown Homocystites alter Barrande H. const rictus (Bather) conjunct confluent Family Echinoencrinitidae Bather Subfamily Echinoencrinitinae Paul Echinoencrinites angulosus Pander disjunct discrete E. senckenbergii von Meyer „ „ E. reticulatas Jaekel ,, ,, Erinocystis sp. „ „ Subfamily Scoliocystinae Jaekel emend. Paul Schizocystis armata (Forbes) „ confluent Glansicystis baccata (Forbes) „ „ Osculocystis monobrachiolata Paul „ ,, 728 PALAEONTOLOGY VOLUME 11 Family Rhombiferidae Kesling Rhombifera bohemica Family Pleurocystitidae Amecystis laevis (Raymond) A. sp. nov. (Neebish Channel, Michigan) Praepleurocystis ivatkinsi (Strimple) Pleurocystites elegans Billings P. filitextus Billings P. squamosus Billings P. rugeri Salter P. quadratus (Bather) P. gibbus (Bather) Regulaecystis pleurocystoides Dehm Family Glyptocystitidae Bather Glyptocystites ehlersi Kesling G. multiporus Billings G. regnelli Sinclair G. batheri Sinclair Family Callocystitidae Bernard Cal/ocystites jewetti Flail C. canadensis Billings C. snbglobosus (Hall) Sphaerocystites nmltifascialns Hall S. bloomfieldensis Schuchert Hallicystis imago (Hall) H. elongata Jaekel H. attenuata Paul Lepadocystis moorei (Meek) Brockocystis tecumseth (Billings) Jaekelocystis hartleyi Schuchert Lovenicystis angelini (Haeckel) Apiocystites pentrematoides Forbes Lepocrinites gebhardii Conrad Prunocystites fletcheri Forbes Lipsanocystis traversensis Ehlers & Leighley L. magnns Stumm Tetracystis fene stratus Troost T. chrysalis Schuchert T. oblongns (Forbes) T. elegans (Hall) Strobilocystites calvini White Staurocystis quadrifasciatus (Pearce) Pseudocrinites bifasciatns Pearce P. pyriformis Paul P. gordoni Schuchert P. perdeivi Schuchert P. stellatus Schuchert P. clarki Schuchert Superfamily Hemicosmitida Jaekel Family Hemicosmitidae Hemicosmites pyriformis von Buch H. extraneus Eichwald slits disjunct dichopores discrete disjunct confluent conjunct ,, disjunct ?confluent 99 disjunct confluent cryptorhombs throughout PAUL: DICHOPORITE PORE-STRUCTURES IN CYSTOIDS 729 H. sp. nov. (Kullsberg, Sweden) cryptorhombs throughout H. malum Pander H. cf. verrucosus Eichwald H. sp. (Karrast, Estonia) H. sp. (Uxnorm, Estonia) Family Caryocrinitidae Bernard Caryocrinites ornatus Say C. roemeri Jaekel C. septentrionalis Regnell C. elongatus (Rowley) Family Thomacystidae Paul Thomacystis tuberculata Paul REFERENCES barrande, j. 1887. Systeme Silurien du centre de !a Boheme. I'' par tie: Recherches paleontologiques. 7. Classe ges Echinodermes. Ordre des Cystides. xvii+233 pp., 39 pi., Leipzig and Prague. bather, f. a. 1913. Caradocian Cystidea from Girvan. Trans. R. Soc. Edin. 49, 359-530, 6 pi. breimer, a. 1963. On a specimen of Pleurocystites (Cystoidea) from an erratic boulder near Westerhaar (Netherlands). Proc. K. ned. Akad. Wet. (B) 66, 296-302, 2 pi. delpey, g. 1942. Organes speciaux d’Echinodermes primitifs: les pectinirhombes. Bull. Soc. geol. Fr. [5] 11, 207-17, 5 text-figs. forbes, e. 1848. On the Cystideae of the Silurian rocks of the British Isles. Mem. geol. Surv. U.K. 2, 483-538, pi. 11-23. Hudson, G. H. 1911. Studies of some early Siluric Pelmatozoa. Bull. N.Y. St. Mus. 149, 195-258, 7 pi. 1915. Some fundamental types of hydrospires with notes on Porocrinus smithi Grant. Ibid. 177, 163-73, 2 pi. huxley, T. h. (translator) 1854. On the structure of the echinoderms by Prof. J. Muller. Ann. Mag. nat. Hist. [2] 13, 1-24, 112-23, 241-56. (See Muller 1854.) hyman, l.h. 1955. The invertebrates, 4. Echinodermata. The coelomate Bilateria. vi+763 pp. New York, Toronto, and London. jaekel, o. 1899. Stammesgeschichte der Pelmatozoen. 1. Thecoidea and Cystoidea. x+442pp., 18 pi., 88 text-figs. Berlin. jefferies, R. p. s. and minton, p. 1965. The mode of life of two Jurassic species of ‘ Posidonia'. Palaeon- tology, 8, 156-85, pi. 19. kesling, r. v. 1961. A new Glyptocystites from the Middle Ordovician strata of Michigan. Contr. Mus. Paleont. Univ. Mich. 17, 59-76, 3 pi. 1 962. Morphology and taxonomy of the cystoid Cheirocrinus anatiformis (Hall). Ibid. 18, 1-21 , 4 pi. 1963. Key for the classification of cystoids. Ibid. 101-16. 1968. Cystoids, in R. C. Moore (ed.), Treatise on invertebrate paleontology. Part S. Echinodermata 1, S85-S267. muller, j. 1854. liber den Bau der Echinodermen. Abh. preuss. Akad. Wiss. (for 1853), 123-219, 9 pi. (See Huxley 1854.) paul, c. r. c. 1967o. The British Silurian cystoids. Bull. Br. Mus. nat. Hist. (Geol.) 13, 297-356, 10 pi. 19676. Hallicystis attenuata, a new callocystitid cystoid from the Racine Dolomite. Contr. Mus. Paleont. Univ. Mich. 21, 231-53, 4 pi. ■ 1967c. The functional morphology and mode of life of the cystoid Pleurocystites, E. Billings, 1854. Symp. zool. Soc. Loud. 20, 105-21, 22 figs. 1968. Macrocystella Callaway, the earliest glyptocystitid cystoid. Palaeontology, 11, 580-600, pi. 111-13. regnell, g. 1945. Non-crinoid Pelmatozoa from the Paleozoic of Sweden. A taxonomic study. Meddn. Lunds geol. -miner. Instn 108, 255 pp., 15 pi. RUDWiCK, m. J. s. 1960. The feeding mechanisms of spire-bearing fossil brachiopods. Geol. Mag. 97, 369-83, 8 text-figs. 730 PALAEONTOLOGY, VOLUME 11 rudwick, m. j. s. 1961. The feeding mechanism of the Permian brachiopod Prorichthofenia. Palaeontology, 3, 450-71, pi. 72-4. 1964n. The function of zigzag deflections in the commissures of fossil brachiopods. Palaeontology, 7, 135-71, pi. 21-9. 19646. The inference of function from structure in fossils. Br. J. Phil. Sci. 15, no. 57, 27-40. Sinclair, G. w. 1948. Three notes on Ordovician cystoids. J. Paleont. 22, 301-14, pi. 42-4. stainbrook, m. a. 1941. Last of great phylum of the cystids. Pan-Am. Geol. 76, 83-98, pi. 7. C. R. C. PAUL Geology Department Indiana University Northwest Gary, Indiana 46408 U.S.A. Typescript received 12 February 1968 AUSTRALIRHYNCHIA, A NEW LOWER DEVONIAN RHYNCHONELLOID BRACHIOPOD FROM NEW SOUTH WALES by N. M. SAVAGE Abstract. Australirhynchia cudalensis gen. et sp. nov., family Rhynchotrematidae, is described and figured from the early Siegenian Mandagery Park Formation, New South Wales. Study of a Lower Devonian brachiopod fauna from the Mandagery Park Formation, near Manildra, New South Wales, has led to the recognition of a new rhynchonelloid genus, Australirhynchia. The Mandagery Park Formation consists of interbedded lime- stones and tuffaceous sandstones of probable early Siegenian age (Savage 1967, 1968). A silicified horizon within the basal limestone of the formation, 3 miles south of Manil- dra, at locality 1 of Savage (1968), contains a rich brachiopod fauna from which the material described herein has been selected. SYSTEMATIC PALAEONTOLOGY Superfamily rhynchonellacea Gray 1 848 Family rhynchotrematidae Schuchert 1913 Subfamily orthorhynchulinae Cooper 1956 Genus australirhynchia nov. Type species. Australirhynchia cudalensis sp. nov. Diagnosis. A plicate rhynchonelloid with strongly angular costae which multiply by bifurcation on the dorsal fold and intercalation in the ventral sulcus. The dorsal interior has a small boss-like cardinal process between ponderous inner socket ridges. No median septum is present. The ventral interior is without distinct dental lamellae and the long heavy teeth, which are fused to the valve walls, bear deep longitudinal grooves. Comparison. The ponderous inner socket ridges and small cardinal process, together with the massive, longitudinally grooved teeth and absence of distinct dental lamellae, suggest a relationship to the genera grouped as the Orthorhynchulinae by Schmidt (1965). Of genera possessing a cardinal process but without a septalium the nearest is Macha- eraria Cooper 1955, for this genus also lacks a dorsal median septum and has similarly long curved crura, crescentic in cross-section; moreover, there is a suggestion of bifurca- tion on the dorsal umbones of the specimens figured by Cooper, though this feature is much less prominent and more variable than in Australirhynchia. There are, however, several major differences; the outline of Machaeraria is more triangular with the greatest width anterior, and not posterior, of mid-length, the shell contours are more rounded, and the costae are more numerous. Internally the hinge plate of Machaeraria is deeply [Palaeontology, Vol. 11, Part 5, 1968, pp. 731-5, pi. 141.] 732 PALAEONTOLOGY, VOLUME 11 cleft, the inner socket ridges are much smaller, and the teeth are less massive with no deep longitudinal grooves. Another closely related genus is Sicorhyncha Havlicek 1961, from the Lower Devonian of Bohemia. This resembles Australirhynchia in lacking a septalium and a dorsal median septum. However, it differs greatly in general shell outline, being bluntly rounded posteri- orly and very broad anteriorly with concave lateral margins. Furthermore, it does not possess the strongly inflated inner socket ridges and massive teeth of Australirhynchia. Latonotoechia Havlicek 1961, also from the Lower Devonian of Bohemia, is another genus without a septalium and a dorsal median septum. In addition, it has massive teeth and inflated inner socket ridges (Havlicek 1961, p. 25, text-fig. 1). It is easily dis- tinguished from the Manildra genus, however, for it has rounded contours quite unlike those of Australirhynchia and lacks the bifurcating and intercalating costae. Of the genera referred by Schmidt (1965) to the Rhynchotrematinae, the Lower Silurian genus Stegerhynchus Foerste 1909, and the upper Silurian genus Stegorhynchella Rzhonsnitskaya 1959, both show a general resemblance to Australirhynchia. The Manildra form is distinct from these genera because of its unusual ornamentation and the absence of a median septum or callosity supporting the hinge plates. Moreover, both Stegerhynchus and Stegorhynchella have distinct umbonal cavities in their ventral valves, a feature completely absent from Australirhynchia where dental lamellae cannot be distinguished in any of the specimens examined. Australirhynchia cudalensis sp. nov. Plate 141 Material. Of a total of 167 silicified specimens 146 are complete shells with the valves conjoined. The remaining few specimens consist of 9 dorsal valves, 8 ventral valves, and 4 posterior fragments showing EXPLANATION OF PLATE 141 Australirhynchia cudalensis gen. et sp. nov. (All figures x 6) Figs. 1-4. Dorsal, ventral, posterior, and lateral views of SU 19579, a large specimen with a distinctly pentagonal outline, an emarginate anterior margin, and an anacline ventral interarea. Figs. 5-9. Ventral, dorsal, lateral, anterior, and posterior views of SU 19578 (holotype). Fig. 10. Ventro-anterior view of broken dorsal valve SU 19584 showing the small, rounded, posterior adductor scars and the larger, ovate, anterior adductor scars. Figs. 11, 12. Dorso-anterior and dorsal views of specimen SU 19581, a ventral valve with the socket plates and cardinalia of the dorsal valve still attached to show the articulation. The small, rounded, adductor scar is prominent within the long, faintly impressed, diductor field. Fig. 13. Dorsal view of broken ventral valve SU 19585 showing the deep delthyrial cavity bounded by long, thick, deeply grooved teeth from which the distal parts are broken. Figs. 14-18. Dorsal, lateral, ventral, anterior, and posterior views of young specimen SU 19577. Here the maximum thickness is more posteriorly placed than in older specimens and the ventral interarea is distinctly apsacline. Fig. 19. Ventral view of dorsal valve SU 19580 showing the small cardinal process between large inflated inner socket ridges, and the elongate adductor muscle field. Fig. 20. Posterior view of specimen SU 19580 showing further the inflated inner socket ridges. Figs. 21-5. Posterior, lateral, anterior, dorsal, and ventral views of specimen SU 19576, a young stage in which the bifurcation of the two central costae of the dorsal valve and the intercalation of two new costae in the ventral valve have just occurred. The fold and sulcus are poorly developed at this stage. Palaeontology, Vol. 11 PLATE 141 18 19 25 24 23 SAVAGE, Australirhynchia N. M. SAVAGE: AUSTRALIRHYNCHIA GEN. NOV. 733 the valve articulation. The specimen numbers used are those of the Palaeontology Collection, Depart- ment of Geology and Geophysics, University of Sydney. Holotype. Complete shell SU 19578. Description. Exterior. In outline the shell is transversely ovate to subpentagonal with the greatest width at the hinge line. The cardinal margins are sharply rounded, the lateral margins are gently rounded, and the anterior margin is straight or emarginate. In lateral profile the shell is subequally biconvex with the greatest thickness between mid-length and the anterior margin (PI. 141, figs. 4, 7). The ventral valve is moderately convex with a strongly curved umbo and a prominent, suberect beak (PI. 141, fig. 4). An apsacline to anacline interarea has a width about half the shell width and an apical angle of about 1 10°. The delthyrium includes an angle of text-fig. 1. Australirhynchia cudalensis gen. et sp. nov. A. Interior of the dorsal valve showing the small cardinal process between the large inflated inner socket ridges. Note also the elongate adductor muscle field, b. Interior of the ventral valve showing the long, deeply grooved teeth and the small, rounded adductor field set within the larger diductor muscle field. about 80° and is partly closed by small triangular deltidial plates. These roof over the dorsal umbo and almost meet to leave an oval foramen extending to the posterior extremity of the valve (PI. 141, figs. 1, 6). The dorsal valve is strongly convex with a broad umbo and no interarea. A broad ventral sulcus extends most of the valve length to the rectangularly uniplicate anterior commissure where it is bordered by very steep lateral walls (PI. 141, figs. 5, 8). The dorsal fold is less pronounced posteriorly but very distinct anteriorly (PI. 141, fig. 8). About 10-12 strong angular costae are present on the dorsal valve, with 4 on the fold, and about 9-11 costae are present on the ventral valve, with 3 in the sulcus. The costae on the fold invariably bifurcate to increase from 2 to 4 (PI. 141, figs. 1,6, 14, 24), whilst the single initial costa in the sulcus increases to 3 by an intercalation on either side (PI. 141, figs. 2, 5, 16, 25). In both valves the increase occurs at about the 1-5 mm. growth stage. Numerous faint concentric growth lines cross the costae. Interior of ventral valve. The ventral interior has a deep, narrow, delthyrial cavity bounded by long thick teeth, each bearing a prominent longitudinal groove with which the corresponding dorsal inner socket ridge articulates (PI. 141, fig. 1 1). The teeth diverge anteriorly at about 45°. A long and anteriorly expanding diductor muscle field extends almost half the valve length (text-fig. 1b). The more strongly impressed adductor field 734 PALAEONTOLOGY, VOLUME 11 is very small and rounded, with a length about one-quarter that of the diductor field in which it is almost centrally placed (PI. 141, fig. 12). Interior of dorsal valve. The dorsal interior has a small boss-like cardinal process placed between large inflated inner socket ridges (PI. 141, fig. 19). The sockets, which are small, triangular, and widely divergent, are bounded postero-laterally by the low overhanging valve margin. Thin, ventrally curved crura, which arise from the inside of the socket ridges, are slightly crescentic in cross-section with the convex side directed outwards (PI. 141, fig. 19). No median septum is present. The adductor muscle scars are clearly -o 5- £ i 5- 0< , , , , , , ■ 0 5 10 Length 04 . 1 . . . . . . . . 0 5 10 Length A B text-fig. 2. Australirhynchia cudalensis gen. et sp. nov. Scatter diagrams of the dimensions of 60 speci- mens plotted in mm. a. Plot of width to length, b. Plot of thickness to length. impressed and extend about half the distance to the anterior margin (text-fig. 1a). Rounded, widely spaced, posterior adductors are separated from larger, longitudinally ovate anterior adductors by narrow, transverse ridges (PI. 141, fig. 10). Measurements (in mm.) Length Width Thickness SU 19577 Complete shell 3-5 4-1 2-6 SU 19578 Complete shell 4-5 5-4 40 SU 19579 Complete shell 5-3 7-3 4-6 SU 19580 Dorsal valve 6-5 8-6 — SU 19581 Ventral valve 60 7-6 — Variation. Shells display little variation in external form apart from differences in thick- ness and width related principally to ontogeny (text-fig. 2). The multiplication of the costae by bifurcation on the fold and intercalation in the sulcus is invariably present and shows remarkably little variation. The costation on the flanks also shows a low vari- ability so that in most shells the dorsal valve possesses 10 costae and the ventral valve 11 costae. Internally even the younger specimens show the massive teeth and socket ridges. Ontogeny. A good range of growth stages has been collected and the later part of the ontogeny of the species can be followed in detail. In the youngest stage available the N. M. SAVAGE: AUSTRALIRHYNCHIA GEN. NOV. 735 maximum width occurs just anterior to mid-length, no bifurcation or intercalation of costae has occurred, and the ventral interarea is apsacline. At a slightly larger stage (PI. 141, figs. 24, 25) the bifurcation of the two central costae of the dorsal valve and the intercalation of costae on either side of the central costa of the ventral valve is already apparent. As the shell approaches maturity it assumes a more pentagonal outline with the maximum width posterior to mid-length (PI. 141, fig. 6). In a mature specimen the ventral interarea is distinctly concave and more anacline than apsacline. The maximum thickness of a mature specimen is well towards the anterior of the shell, largely because of the strong development of the dorsal fold and ventral sulcus (PI. 141, figs. 7, 8). Internally, large inflated inner socket ridges and deeply grooved teeth are present in all but the youngest stages. In mature specimens the adductor muscle scars in both valves are deeply impressed. Phylogeny. Although Australirhynchia is distinct from Stegerhynchus and Machaeraria in several important respects, these two genera appear to be the most closely related forms. It seems possible that both Australirhynchia and Machaeraria evolved indepen- dently from Stegerhynchus , although there is little evidence for this apart from the general morphological similarities mentioned above and the absence of possible alternative precursors during the Silurian. It is unlikely that Australirhynchia was derived from Machaeraria as it retains such features as inflated inner socket ridges and a relatively angular shell shape; features characteristic of Stegerhynchus but not present in Macha- eraria. Acknowledgements. The author would like to acknowledge the assistance of Professor C. E. Marshall and Dr. G. H. Packham of the Department of Geology and Geophysics, University of Sydney, where this work was commenced as part of a larger faunal study during the tenure of a Teaching Fellowship. It is also a pleasure to thank Professor F. H. T. Rhodes for his encouragement and for the use of facilities in the Department of Geology, University College of Swansea; also Dr. A. J. Boucot and Dr. J. G. Johnson of the California Institute of Technology for their helpful comments. REFERENCES cooper, G. a. 1955. New genera of Middle Paleozoic brachiopods. /. Paleont. 29, 45-63, pi. 11-14. havlIcek, v. 1961. Rhynchonelloidea des bohmischen alteren Paleozoikums (Brachiopoda). Rozpr. ustred. Ust. geol. 27, 1-211, 27 pi. savage, n. m. 1967. Studies in the Silurian and Devonian of the Manildra District, New South Wales. Unpublished Ph.D. thesis, Univ. Sydney. — — 1968. The Geology of the Manildra District, New South Wales. J. Proc. R. Soc. N.S. W. (in press). schmidt, h. 1965. Family Rhynchotrematidae. In R. C. Moore (ed.), Treatise on invertebrate paleon- tology, Part H. Brachiopoda, H572-6, Fawrence, Kansas. N. M. SAVAGE Department of Geology University of Natal Durban S. Africa Typescript received from author 1 April 1968 THE LLANDOVERY TRANSGRESSION OF THE WELSH BORDERLAND by A. M. ZIEGLER, L. R. M. COCKS, and W. S. McKERROW Abstract. The study of the evolution of several brachiopod genera has enabled the early Silurian shelf sequences of the Welsh Borderland to be correlated with the type area of Llandovery, and, to some extent, with the grap- tolite zonal sequence. The transgression across the Borderland started at the beginning of Llandovery time, when areas in Montgomeryshire became inundated after a short break in deposition at the end of the Ashgill. The sea reached Shropshire, and possibly as far east as the Malverns and May Hill, by Middle Llandovery times. By late Upper Llandovery times, much of the English Midlands were covered. Fossil communities indicate the relative depths in which the Llandovery sediments were deposited. With the advance of the sea, most sequences show a progressive increase of depth with time, although minor reversals are known. The many gaps in the local sequences are probably due to submarine erosion or non-deposition, rather than to uplift and sub-aerial erosion, because they are characteristically followed by progressively deeper-water communities. The community distribution indicates that a continuous gradient was maintained from the coast to the shelf margin, once the topographic relief of the original surface had been filled in. During the Lower Silurian there was a spread of the shelf sea from Central Wales over the Welsh Borderland (text-fig. 1). Brachiopods are the dominant element in the shelf faunas of this region. Now that the evolution of several brachiopod genera is known in detail, it is possible to correlate the shelly deposits with the standard section at Llandovery. Animal communities have been defined for this period, and they have been shown to reflect the depth of the sea. These communities are diachronous, and migrated as the transgression advanced. The twin concepts, of evolving lineages as a basis for correlation and of animal communities as a basis for interpreting the environment, provide the foundation of this work. BASIS OF CORRELATION The type section of the Llandovery is to be seen south-east of the town of this name in Wales (Jones 1925). Brachiopods are the dominant faunal element in the area, but correlation of other shelly areas with the type section has had to await the study of the commoner brachiopod lineages. This approach contrasts with the tendency for workers in the shelf areas to have correlated Silurian beds by the assemblages of species present. Evolutionary sequences are now known for the stricklandiids (Williams 1951, St. Joseph 1935), the pentamerids (St. Joseph 1938), the atrypide Eocoelia (Ziegler 19666) and the strophomenide Leptostroplria (Cocks 1967#). Some brachiopod lineages and their relationship with the type Llandovery are shown in text-fig. 2. The only brachiopods on this list which do not occur at Llandovery itself are E. sulcata, C. lirata typica, and Pentameroicles; their assignment to C6 is based on the fact that they are known to occur between C5 and basal Wenlock beds in adjacent areas of the Welsh Borderland. Most of the Lower Silurian of the British Isles carries a fauna of graptolites. Lapworth (1878) and Elies and Wood (1901-18) established a sequence of graptolite zones which [Palaeontology, Vol. 11, Part 5, 1968, pp. 736-82.] ZIEGLER, COCKS, AND McKERROW: LLANDOVERY TRANSGRESSION 737 have provided an accepted basis for correlation. A few graptolites of zonal value have been found at Llandovery (text-fig. 2) but are only recorded from three horizons. Jones (1925, p. 360) found graptolites referable to his cicinaces Zone in A4 (this is now equated with the lower part of the cyphus Zone — Toghill 1968): Jones also (1949, p. 52) cor- text-fig. 1. Location map of the Llandovery rocks of Wales and the Welsh Borderland, showing traverse lines and position of local sections. relates B3 with the C. cometa Band, which is in the upper part of the convolutus Zone; and he further records graptolites of the sedgwickii Zone (1925, p. 370) from beds assigned to Cx. This last fauna has been confirmed by our recent collecting from the other known locality of this restricted facies near an old ford across the River Sefin (Grid Ref. SN/7419 2811). 738 PALAEONTOLOGY, VOLUME 11 In addition to these finds from Llandovery itself, other links between the brachiopod and graptolite sequences have been determined in the upper half of the Llandovery of Shropshire (Cocks and Rickards, in press); but, as yet, firm ties are not proved for the whole Llandovery succession. Brachiopod Graptolite Lineages Zones Wenlock 1 Eocoelia t Cyrtograptus murchisoni c6 sulcata* Costistncklandia 1 irata . typica * * l ' Prntnn Monoclimacis crenulata Monoclimacis griestoniensis Monograptus crispus Monograptus turriculatus Monograptus sedgwickii Monograptus convolutus Pristiograptus gregarius Pristiograptus cyphus Cystograptus vesiculosus Akidograptus acuminatus Gl yptograptus persculptus | ^ rGnTQr Eocoelia Costistncklandia curtisi lirata, alpha 1 4. >er n -p* "1 ^ 1 Eocoelia Stricklcndia u. ~> ^3 intermedia lens ultima 1 1 C2 Pentar Eocoelia Stricklandia nerus C, ■s. hemisphaerica lens progressa * * O B3 * ° *o R T3 T3 D, c — Strickla lens nd ia ntermedia A4 Strickla ndia 0) A3 > . lens typica l Lov > ro Stricklandia lens prima A, Ashgill Dicellograptus anceps text-fig. 2. The correlation of brachiopod lineages and graptolite zones with the standard sequence at Llandovery. The asterisks against some brachiopods indicate that they have not been found at Llandovery; the tie-arrows indicate graptolite finds at Llandovery. The limits of the Llandovery Series at Llandovery have not yet been properly defined, nor have they been firmly correlated with the sequence of graptolites. However, in this paper, we take the base of the series as equivalent to the base of the Glyptograptus persculptus Zone, and the top of the series at the base of the Wenlock; or as equivalent ZIEGLER, COCKS, AND McKERROW: LLANDOVERY TRANSGRESSION 739 to the base of the C. murchisoni Zone, in the sense of Elies and Wood (1901-18). This last zone is sometimes now divided into a lower Cyrtograptus centrifugus Zone and an upper, restricted, C. murchisoni Zone. DISTRIBUTION OF ANIMAL COMMUNITIES By means of large collections from single beds (many with over a thousand specimens), it is possible to quantify the proportions of species present at each locality. Five main benthic animal communities have been defined by this means from the Llandovery (Ziegler 1965, Cocks 19676, Ziegler, Cocks, and Bambach 1968). Each community is characterized by a particular faunal assemblage, consisting chiefly of brachiopods, and each has been named after a prominent brachiopod genus in the assemblage. They are the Lingula Community, the Eocoelia Community, the Penta- merus Community, the Strieklandia Community, and the Clorinda Community. At any one time, they occurred in this sequence, from the inferred coastline to the outer margin of the shelf. In beds of C5 and C6 age, Pentamerus and Strieklandia have evolved into Pentameroides and Costistricklandia respectively, and the community names change correspondingly. So far the communities have been described only from the Upper Llandovery, but, apart from Eocoelia, the characteristic genera also occur in the Middle Llandovery, and the Strieklandia and Clorinda Communities are present in Lower Llandovery beds. A Cryptothyrella Community occurs in the Lower and Middle Llandovery with associ- ates of the Eocoelia Community, and it appears to be an early equivalent of the latter community; it is abundant in basal beds at several areas, and occurs also in North America (Ziegler and Boucot, in press). In addition, during parts of the Middle and Upper Llandovery, the position of the Eocoelia Community in Shropshire is taken by the parallel ? Rostricel/ula Community (Cocks and Rickards, in press). Also, in places near the edge of the shelf, the Clorinda Community dwindles to a sparser 'Marginal Clorinda' Community, with many typical members absent. Apart from the burrowing Lingula, all the Llandovery brachiopods were epifaunal. Thus the majority of Silurian level bottom communities contrast with equivalent modern communities which are mainly infaunal. These Silurian communities were therefore less dependent on the type of substrate than their modern counterparts. Rocky bottom communities are known at the base of some local sequences, although in all cases where these have been preserved they are accompanied by some soft-bottom elements. The distribution of the main communities at three horizons in the Llandovery is shown in three maps at the end of the paper (text-figs. 12-14). While the actual depths repre- sented by the communities are uncertain, the lines bounding them are presumed to parallel the bathymetric contours. Thus we have a picture of the changing configuration of the shelf sea during the transgression, and of the migration of the communities with time. A distinction is drawn between ‘basin’ and ‘graptolitic’ facies. The former is restricted to areas (for example text-fig. 3, cols. 1 and 10) where rapid sediment supply, great C 6055 3 c 740 PALAEONTOLOGY, VOLUME 11 water depths, and considerable subsidence has resulted in large thicknesses of sediment. The term ‘graptolitic’ facies is used for rocks in which the dominant fossils are grapto- lites, and in which shelly forms, apart from a few orthocones and epiplanktonic bivalves, are generally absent. This facies represents one or other of two environments, firstly deposition under water which was too deep for benthic colonization at that time, or, secondly, water of any depth which had a foul bottom, lower temperature, or other physical conditions inimical to a shelly fauna. When a graptolitic facies occurs to sea- ward of a Clorinda Community, we feel justified in selecting the first of these alternatives, and interpreting the occurrence as being too deep for a shelly fauna ; this situation applies to all of the Llandovery in the Welsh Borderland. However the second alternative certainly applied at other times and places during the Lower Palaeozoic, for example this is how we would interpret the Wenlock Shale to the south of the Long Mynd, Shropshire. Acknowledgements. We are indebted to the late Professor W. F. Whittard for his advice, fossil collec- tions, and loan of field maps. Dr. M. L. K. Curtis kindly made his thesis and collections available, and conducted us to localities in the Tortworth Inlier. Drs. W. B. N. Berry, R. B. Rickards, and P. Toghill identified the graptolites we have found, and discussed older records. The theses by Ziegler (1963) and Cocks (1965), both at Oxford, were supported in part by the Burdett-Coutts Fund and the Department of Scientific and Industrial Research respectively. We are grateful to Mr. J. M. Edmonds and Mr. H. P. Powell for housing our thesis collections in the Oxford University Museum (Appendix 1), and for providing funds for their labelling. Collections made subsequently are deposited in the British Museum (Natural History) (Appendix 2) and the U.S. National Museum (Appendix 3). These last were prepared by Ziegler in the laboratories of Dr. A. J. Boucot at the California Institute of Technology from 1964-6. More recently, Ziegler has been supported by Grant 910-G from the Petroleum Research Fund, administered by the American Chemical Society; and by Grant GB-6592 from the National Science Foundation of America. DESCRIPTIONS OF THE OUTCROPS Three traverses across part of Wales and the Borderland (text-fig. 1) are represented by the composite sections in text-fig. 3. In the following account, the areas studied in detail are treated first; i.e. the sections from Minsterley to Tortworth (text-fig. 3, cols. 5-6, 11-19, and 25-9). These constitute the classical Welsh Borderland area. Secondly, we discuss the small inkers to the east of the Borderland (text-fig. 3, cols. 7-9 and 20). We have examined collections from these areas, and include them for the sake of completeness. Finally we contrast these shelf sequences with areas to the west of the Borderland, which in Llandovery times formed the shelf margin and the basin. These areas are mostly in Wales (text-fig. 3, cols. 1-4, 10, and 21-3); we have collected in all of them, with the exception of the purely grapto- litic sequences. text-fig. 3. Three traverses across the Welsh Borderland (see text-fig. 1) showing local sections with communities, correlation, lithology, and stratigraphical relationships. True thicknesses are shown. The distances between columns are not the straight-line distances, but the N. 60° W. component of these distances. 742 PALAEONTOLOGY, VOLUME 11 MINSTERLEY AND BOG Text-fig. 3, cols. 5 and 11 Exposures of Llandovery sediments to the north and west of the Shelve Inlier, Shrop- shire (text-fig. 4), occur in a sinuous strip running north-east from north of Chirbury through Hope Valley and Venusbank to near Minsterley. Various outliers, including text-fig. 4. Locality map of Shropshire (after many authors, with modifications). those at Bog, rest unconformably on the Ordovician of the Shelve Inlier. The area has been re-mapped, but the only major re-interpretation is based on new exposures indicating that the Venusbank ‘outlier’ is connected northwards with the outcrop in Hope Valley. Whittard (1932) used the same stratigraphical names, i.e. Pentamerus Beds and Purple Shale, throughout Shropshire, but the Llandovery rocks of the Minsterley and Bog area differ in lithology and age from the beds in the Wenlock Edge outcrop (see below). Accordingly we propose new names: the Bog Quartzite, Venusbank Lormation, and Minsterley Lormation. (a) Venusbank Formation and Bog Quartzite. The type section of the Venusbank Lorma- tion is in Hope Brook, running eastwards from near Hope Quarry. The base is exposed in both brook (SJ/3554 0207) and quarry (SJ/3550 0207), where a sandstone with occasional shale fragments rests on Hope Shales of Ordovician (Llanvirn) age. The top of the Lormation in Hope Brook is taken above the highest of the dominant sandstones (SJ/3572 0215). ZIEGLER, COCKS, AND McKERROW: LLANDOVERY TRANSGRESSION 743 The base of the formation may also be seen in an outlier (SJ/3458 0146) two-thirds of a mile south-west of Hope Quarry (Tyler 1925). Here a basal conglomerate is present, in which pebbles (up to 5 cm.) of Hope Shales, Stapeley Volcanics (Llanvirn) and Stiperstones Quartzite (Arenig) occur. This outlier occurs at 900 ft. O.D. (Ordnance Datum — feet above sea level) but another small outlier is present 1 50 yards to the east, on the side of the valley and between 650 and 700 ft. O.D. ; Hope Quarry lies at 600 ft. O.D., and, to the south-east, the Llandovery at Venusbank (SJ/3535 0125) is at 800 ft. O.D. Thus the conclusions reached by Whittard (1932, pp. 893-5), of an irregular base to the Llandovery of the area, are confirmed; steep dips are absent and there is no detectable subsequent faulting. It appears that the present-day Hope Valley has been re-excavated along the line of a pre-Middle Llandovery depression, and that the Llan- dovery sea spread over a surface relief of 200-300 ft. in this area. The thickness of the Venusbank Formation varies between 0 and 200 ft. (0-65 m.). The formation is well exposed at Hope Quarry (SJ/3550 0207), where a thickness of 30 ft. (10 m.) is visible. Sandstone bands up to 3 ft. (1 m.) thick occur there, sometimes including shells or mudflakes at or near their base, with each unit fining up into unfossili- ferous silts. Some of the sandstones have parallel laminae throughout, but cross-bedding is absent, apart from very occasional ripple cross-lamination on the tops of one or two sandstones. Unit bases are sharp, with no visible bottom structures apart from worm traces. Low-angle channelling occurs with occasional silty partings at the base of the channelling sandstone. These sediments may represent proximal turbidites filling a previously scoured channel of Pentamerus- and Stricklcmclia- Community depth. The Llandovery also rests unconformably on the Ordovician in some poorly exposed outliers near Bog (SO/355 979) and at Round Hill (SO/348 993). Many large loose blocks of coarse quartzite are present; often containing pebbles and shale chips. These beds are termed the Bog Quartzite. Fossil communities. The Hope Valley sections of the Venusbank Formation have mainly yielded fossils of the Stricklandia Community (Locs. 68-70, 72, 73) ; the exceptions are the presence of Pentamerus Communities at some places near the base of the formation (Loc. 74 and at SJ/3555 0107). This suggests that the irregular landscape was covered by a moderately deep shelf sea ( Pentamerus Community) that soon became slightly deeper ( Stricklandia Community). However, there may have been some fluctuation in depth, as Whittard (1932, p. 876) reports Pentamerus above Stricklandia- bearing beds in Hope Quarry. In Ox Wood Dingle (Loc. 67), 4 miles west of Hope Valley, the Stricklandia Community is also present. However, early shallow-water conditions may be indicated at The Stubbs (2 miles north-west of Hope Quarry — SJ/324 033) where Whittard records (1932, p. 874) Lingula in growth position near the base of the Venusbank Formation. The fauna of the Bog Quartzite (Locs. 65, 66) is a mixture of a near-shore Crypto- thyrella Community with some rocky-bottom elements (Ziegler, Cocks, and Bambach 1968). Correlation. All the collections from the Venusbank Formation and Bog Quartzite have yielded Stricklandia ; the subspecies present at Bog is S. lens intermedia indicating a 744 PALAEONTOLOGY, VOLUME 11 Middle Llandovery age; the subspecies from the Venusbank Formation is at the bound- ary between S. lens intermedia and progressa and is B3 to Q in age. A collection from near the base of Hope Quarry (Loc. 70) includes the graptolite Climacograptus aff. rectangularis (identified by R. B. Rickards), which indicates a Middle Llandovery age for the lower part of the formation. Whittard (1932, facing p. 896) records Monograptus runcinatus pertinax from the ‘Pentamerus Beds of the Wilmington-Minsterley’ area; this suggests that at least some of the Venusbank Formation is of early Upper Llandovery age. ( b ) Minsterley Formation. The type section goes diagonally across the strike of the formation and is in Hope Brook. The base is taken above the top sandstone of the Venus- bank Formation (SJ/3572 0215), and the top above the last purplish bed in the exposure to the south of Wagbeach (SJ/3642 0270). The thickness of the Minsterley Formation is about 400 ft. The lowest beds seen are calcareous siltstones and mudstones, with occasional coarser beds; these represent a distinct sedimentary environment unknown elsewhere in the Welsh Borderland. They pass up into maroon, green, or blue mudstones, the top 150 ft. (49 m.) being of the same dominantly purple colour as the Hughley (Purple) Shales on the main Wenlock Edge outcrop. The mudstones contain some interbedded siltstone and bands of fragmented shells. Locally, the Minsterley Formation overlaps the Venusbank Formation, e.g. at Estell (Whittard 1932, p. 877) and east of Minsterley; in these areas the formation tends to be coarser. The Minsterley Formation is followed by several hundred feet of unfossiliferous turbidites, well exposed in Flope Valley, which may be either late Llandovery or early Wenlock in age. Fossil communities. Collections in Hope Brook (Loc. 71) and Minsterley-Habberley Lane (Loc. 75) are both in the Clorinda Community; this indicates that the Minsterley Formation was deposited in deeper water than the underlying Venusbank Formation. Correlation. The brachiopods collected from the Minsterley Formation provide no precise evidence of age, but Whittard’s (1932, p. 877) record of graptolites from Locality 75 includes Monograptus halli, M. becki, and M. cf. proteus. This assemblage indicates a turriculatus Zone age (Cocks and Rickards, in press). A lack of fossils in the upper part of the formation, and in the overlying turbidites, makes it uncertain whether it ter- minates, even approximately, at the end of Llandovery time. NORBURY AND CHURCH STRETTON Text-fig. 3, cols. 12 and 13 A long winding outcrop of Llandovery rocks borders the south side of the Cambrian and Ordovician rocks of the Shelve Inlier, and continues all around the southern part of the Pre-Cambrian rocks of the Long Mynd (text-fig. 4). The formation names which we use for this area are the same as those used by Greig et al. (1968); the Pentamerus Beds and the Hughley (or Purple) Shales. The area, sometimes termed the southern Long Mynd-Shelve Outcrop, was mapped by Whittard (1932, pi. 59, 60) and more recently by the Geological Survey (Church Stretton Sheet, New Series, no. 166). Six ZIEGLER, COCKS, AND McKERROW: LLANDOVERY TRANSGRESSION 745 boreholes were drilled to assist the Survey in their mapping, and four of these passed through beds of Llandovery age (Cocks and Rickards, in press). These are called the Eaton Farm (text-fig. 3, col. 12) and Robury Ring Boreholes, both of which are to the west of the southern Long Mynd spur, and the Hamperley (text-fig. 3, col. 13) and Springbank Farm Boreholes, both of which are to the east of the Long Mynd spur and in the Church Stretton Valley. A fifth borehole, at Botvyle (also shown on text-fig. 4), did not penetrate to the local base of the Wenlock. (a) Pentamerus Beds. Between Plowden (SO/382 874) and Little Stretton (SO/445 921), up to 60 ft. (20 m.) of conglomerates are intermittently present below mudstones and other sediments with Pentamerus. These beds, when fossiliferous, contain a Lingula Community, and, in parts, the formation name Kenley Grit (used in the adjacent Wen- lock Edge outcrop — see below) might be employed. However, as the conglomerates fill quite small topographic pockets, and locally vary both in lithology and possible age, we follow Whittard (1932) and the Geological Survey (Greig et al. 1968) in treating these beds as a local coarse base to the Pentamerus Beds. The Hamperley and Springbank Farm Boreholes both revealed complete sections through the Pentamerus Beds, which are 480 (157 m.) and 170 ft. (56 m.) thick in the respective cores (although these thick- nesses may be affected by minor faulting). Whittard (1932) interpreted the basal deposits to the south of the Long Mynd as a beach with sea stacks. However, from faunal and sedimentological considerations, it seems more probable that most of the sedimentation occurred sub-tidally, and certainly a fairly deep-water Pentamerus Community was established at Hillend Farm (Loc. 62), only 20 ft. (7 m.) above the unconformity with the Pre-Cambrian. The Pentamerus Beds are absent in the large embayment to the west of the southern Long Mynd spur (text-fig. 3, col. 12) and also south of The Roveries (SO/323 915); at both places Hughley Shale rests directly upon the unconformity. However, between these two places, at Norbury (Locs. 63 and 64), a local lens of calcareous sandstone, often crowded with Pentamerus, is present between the unconformity and the Hughley Shale, and this lens is termed ‘Pentamerus Beds’. Further west, near Snead, Whittard (1932, p. 871) recorded 280 ft. of Pentamerus Beds, but his description shows that the lowest Silurian rocks contain a limited shelly fauna, perhaps representing an environ- ment deeper than the Clorinda Community. Fossil Communities. In the Hamperley Borehole, the Lingula Community is present between 65 and 115 ft. (21-38 m.) above the base (Cocks and Rickards, in press). This is followed by a IRostricellula Community, which is a parallel community to that of Eocoelia, and occupies a similar position, between Lingula and Pentamerus Communi- ties. In the borehole, a Pentamerus Community succeeds the IRostricellula Community, and is itself succeeded by a Stricklandia Community. After the Stricklandia Community, there was a reversion to the Pentamerus Community immediately below the Hughley Shales. In the Springbank Farm Borehole, the Pentamerus Beds carry a Pentamerus Community throughout, although at approximately two-thirds up the formation, the proportion of Stricklandia increases enough to be termed a Pentamerus! Stricklandia mixture for that part of the core. This mixture was probably contemporary with the Stricklandia Community in the Hamperley Borehole. 746 PALAEONTOLOGY, VOLUME 11 In the surface exposures, the Lingula Community is developed in various parts of the basal coarse facies. At Marshbrook (Loc. 61) we have collected a mixed Lingula /? Ros- tricellula Community. IRostricellula is a close homeomorph of Eocoelia and was listed as the latter by Whittard (1932, p. 864). Pentamerus Communities occur at Hillend Farm (Loc. 62) and at Norbury (Locs. 63 and 64). Further to the west, deeper water was pre- sent near Snead; above a basal Pentamerus Community, Whittard (1932, p. 872) records a fauna which we interpret as a ‘Marginal’ Clorinda Community. Correlation. Graptolites from the Hamperley Borehole indicate the convolutus Zone for the middle part of the formation, with the sedgwickii Zone above. The sedgwickii Zone is also proved for the middle of the Springbank Farm core. The brachiopod collections from Hillend Farm and Norbury both contain Eocoelia, but the former locality carries E. hemisphaerica (indicating C4_2 age), while at Norbury E. intermedia is present (Ziegler 1966 b), and indicates a C3_4 age for beds near the local base of the formation. Whittard (1932, p. 872) records Climacograptus sp., Monograptus halli and M. dextrorsus from Snead; these indicate the turriculatus Zone. This evidence, together with Whittard’s statement (1932, p. 873) that ‘there is a gradation along strike from arena- ceous to argillaceous sediments’, points to a fairly rapid increase in depth westwards from the Long Mynd during C3_4 times. It also emphasizes that both base and top of the Pentamerus Beds are diachronous. ( b ) Hughley (or Purple) Shales. These beds are not well exposed south of the Longmynd- Shelve Inlier. The Hamperley Borehole shows a thickness of 1 50 ft. (49 m.) and Whittard (1932, p. 866) estimated a thickness of 200-30 ft. of Hughley Shales to the south of the Long Mynd. To the west of the Long Mynd spur there is a greater thickness; 560 ft. (184 m.) is seen in the Eaton Farm borehole, although small faults may have affected this figure. Still further west, near Snead, Whittard (1932, p. 871) estimated that a mini- mum of 250 ft. were present. The formation is distinguished from the underlying Penta- merus Beds and overlying Wenlock Shale mainly by its colour, which is a characteristic maroon or purple, although green and brown bands also occur. The Pentamerus Beds are usually blue-hearted mudstones, and the Wenlock Shale grey and slightly siltier. The contacts between the formations are usually sharp. Fossil Communities. Whittard (1932, p. 869) reports a Clorinda Community fauna between Plowden and Minton, and this community is also present in the Eaton Farm, Robury Ring, Hamperley, and Springbank Farm Boreholes. Above the Clorinda Com- munity in the two boreholes to the west of the southern Long Mynd spur, a ‘Marginal’ Clorinda Community occurs, which probably indicates (Cocks and Rickards, in press) a continued deepening of the sea. Correlation. Graptolites from the boreholes show a sequence from within the turri- culatus Zone, through the crispus Zone and into the griestoniensis Zone. The top 75 ft. (25 m.) of the Hughley Shale in the Eaton Farm Borehole are above the griestoniensis Zone, but carry no diagnostic graptolites. Thus the presence or absence of the highest Llandovery crenulata Zone is not yet confirmed. ZIEGLER, COCKS, AND McKERROW: LLANDOVERY TRANSGRESSION 747 WENLOCK EDGE Text-fig. 3, cols. 6 and 14 Llandovery rocks crop out in a strip twenty miles long, which parallels, and directly underlies, the rocks of Wenlock Edge (text-fig. 4). Whittard (1928) described this ‘Main’ outcrop in detail; the three formation names which he used, i.e. Arenaceous Beds, Pentamerus Beds, and Purple Shales, have been subsequently named by the Geological Survey Kenley Grit, Pentamerus Beds, and Hughley Shales. The northern part of this area was covered by the Shrewsbury Memoir (Pocock et al. 1938), and the southern part by the Church Stretton Memoir (Grieg et al. 1968). A recent revision of the northern- most part of the area was prompted by a very large temporary exposure at Devil’s Dingle (Cocks and Walton 1968). (a) Kenley Grit. A maximum thickness of 200 ft. (65 m.) is reached near Kenley itself (SJ/558 004). Four-and-a-half miles to the south-west, the Kenley Grit dies out near Gretton (SO/519 946). To the north-east, it thins more gradually, and is 35 ft. (11 m.) thick in Harper’s Dingle (SJ/634 072), near the point where the Llandovery is covered by the Carboniferous, 6 miles from Kenley. The formation consists of conglomerates, sandstones, arkoses, and some finer beds, but to the north-east there is no conglomerate except at the base; these basal pebbles are mainly derived from Uriconian rocks, but occasional Cambrian and Ordovician pebbles are present (Whittard 1928, p. 740). The formation rests with strong uncon- formity upon Ordovician and older rocks, and is distinguished from the overlying Pentamerus Beds by the abundance of coarse elastics and the absence of calcareous beds. Fossil communities. The fossils recorded from the Grit (Whittard 1928, p. 741) all belong to the Lingula Community. Lingula itself occurs near the top of the formation at Mor- rellswood (near Loc. 46) and Sheinton Brook (near Loc. 50); it is also present in an old quarry near Kenley (Pocock et al. 1938, p. 106). An assemblage with abundant rhyn- chonellides, referable to the Lingula Community, but without Lingula itself, occurs at Cressage Park, two miles east of Kenley (Whittard 1928, p. 741). Correlation. No fossils found in the Grit are of zonal significance; however, at Mor- rellswood (Loc. 46), beds of C2_3 age occur immediately above the Grit, which must therefore be earlier. ( b ) Pentamerus Beds. The Kenley Grit passes up gradationally into the Pentamerus Beds, but to the south-west, where the Grit is absent, the Pentamerus Beds rest uncon- formably upon Uriconian, Tremadoc, and Ordovician rocks. The Pentamerus Beds reach a maximum thickness of between 400 and 500 ft. (about 150 m.) in the north- eastern half of the area, and thin southwards to nothing in the vicinity of Wistanstow (SO/431 857). A Pentamerus-beanng conglomerate 1 in. thick has been reported (Whit- tard 1928, p. 749) in the bed of the Onny River (Loc. 60), but it is absent higher up on the river bank, where Hughley Shales rest directly upon the Ordovician. The lithology of the Pentamerus Beds consists of blue mudstones, with sporadic shelly limestones, and less fossiliferous siltstones and concretionary limestones. At the extreme northern end of the outcrop the ’Dingle Conglomerate’ is seen near the base of the Pentamerus Beds in Harper’s Dingle (SJ/6328 0704); the bed is confined to this one locality, and consists 748 PALAEONTOLOGY, VOLUME 11 of pebbles of Uriconian rock and Pentamerus shells set in a matrix of quartz sand and calcite (Whittard 1928, p. 745, Pocock et al. 1938, p. 107). Apart from this distinctive conglomerate, the Pentamerus Beds may be differentiated from the Kenley Grit by their palaeontology, finer grain size, and more calcareous sediments, but the contact between the two formations is gradational. The top of the Pentamerus Beds is usually sharp, with the colour change to the Hughley Shales, but the contact is sometimes confused by the presence of small turbidite bands (e.g. at Sheinton Brook — Loc. 51). Fossil communities. A collection from near Gilberries (Loc. 57), in the lower part of the Pentamerus Beds, contains abundant Hyattidina. This genus does not fit precisely into any of the main Llandovery communities as at present defined (Ziegler et al. 1968), but abundant Hyattidina are associated with the Eocoelia Community in New York State and also at Presteigne (Locs. 90 and 93); we conclude that the Gilberries collection represents an intermediate environment between the Lingula Community of the Kenley Grit, and the deeper-water communities of the remainder of the Pentamerus Beds. Collections from north-west of Merrishaw (Locs. 52 and 53) and near Ticklerton (Loc. 59) are typical of the Pentamerus Community. At about the centre of the outcrop, the deeper-water Stricklandia Community is present (Loc. 54) and in Sheinton Brook (Loc. 50), although this last collection only contains 1% of Stricklandia. At Morrells- wood (Loc. 46) an assemblage dominated by IRostricelluIa is present, this assemblage also occurs in the Church Stretton area (see above), where it is interpreted as a parallel community to that of Eocoelia. Correlation. Eocoelia hemisphaerica indicates a C4_2 age for the Pentamerus Communi- ties at Ticklerton (Loc. 59) and Merrishaw (Loc. 52). The Stricklandia Community at Sheinton (Loc. 50) and the collection from Merrishaw (Loc. 53) contain S. lens ultima of C3_4 age. Our evidence is thus consistent with a progressive increase of depth from C4 to C4 times during the deposition of the Pentamerus Beds of the Wenlock Edge Outcrop. Graptolites from various parts of the formation (Whittard 1928, Pocock et al. 1938) are all attributable to the turriculatus Zone, which is equivalent to C2-3 times. (c) Hughley (or Purple) Shales. This formation has a thickness varying from 0 to 500 ft. (0-175 m.). The lithology is dominantly purplish mudstone, but occasional coarser turbidite bands also occur locally. The basal contact with the Pentamerus Beds is sharp and there is overlap on to the Ordovician to the south-east. North of Much Wenlock there is an apparently conformable upper contact with the Buildwas Beds; Cocks and Rickards (in press) conclude that these basal Wenlock beds may include equivalents of all the Lower Wenlock graptolite zones. However, as one proceeds south-west down the outcrop, the Middle Wenlock overlaps both the Lower Wenlock and the Hughley Shales and comes to rest unconformably on the Ordovician four miles south of the Long Mynd. Fossil Communities. Most of the collections from the Hughley Shale of the Main Outcrop are typical of the Clorinda Community (Locs. 47-9, 51, 55, 56, 58, and 60), which reflects a deeper-water environment than that of the Pentamerus Beds. However, near the top of the formation to the north of the River Severn, Cocks and Walton (1968) describe a Clorinda! Stricklandia mixture, indicating some slight local shallowing in latest Llandovery times. ZIEGLER, COCKS, AND McKERROW: LLANDOVERY TRANSGRESSION 749 Correlation. The oldest brachiopod fauna collected from the Hughley Shale is from the very base of the formation in Sheinton Brook (Loc. 51), where Eocoelia intermedia indicates a C3_4 age. E. curtisi, of C5 age occurs in collections from Boathouse Coppice (Loc. 49) and Wall-under-Heywood (Loc. 58). Costistrieklandia lirata and Eocoelia sulcata indicates a C6 age for collections from Boathouse Coppice (Loc. 47), Domas (Loc. 55), and Hughley (Loc. 56), all very near the top of the formation. Thus the brachio- pod evidence indicates that the formation spanned the period C3 to C6. Graptolites from the Onny River (Loc. 60) indicate the turriculatus Zone (probably equivalent to C2_3); at several localities in the north-east part of the outcrop the griestoniensis Zone has now been proved. This last occurs in conjunction with C5 indicators such as Eocoelia curtisi, and the probable presence of equivalents of the highest Llandovery crenulata Zone may be inferred by the presence of the C6 brachiopods. PRESTEIGNE Text-fig. 3, cols. 24 and 25 To the south of Presteigne, Radnorshire, the Upper Llandovery crops out along the B4362 road between Nash Scar and the railway bridge near Corton (text-fig. 5); we propose the name Folly Sandstone for these beds. They are folded in an anticline with 750 PALAEONTOLOGY, VOLUME 11 its axis extending west-south-west through The Folly (SO/316 633). These Llandovery beds are overlain by a thin (15-20 ft. — about 6 m.) development of the Nash Scar Lime- stone to the north; to the south their contact is obscure, and the Folly Sandstone may either dip below or be faulted against the 200 ft. (65 m.) of limestone seen in Nash Scar Quarry. The outcrop is covered by drift east of The Folly, and is cut off by an extension of the Church Stretton Fault to the west of Nash Wood. The only published map of the area is the Geological Survey, Old Series, no. 56, NE. and SE., of more than a century ago; but it shows the Llandovery correctly, except for a possible fault north of Nash Scar quarry. The most recent description of these rocks was published in abstract form by Kirk (1951). The Folly Sandstone consists of dark sandstones in beds ranging from a few inches to a few feet thick. There are many thin conglomerates with rounded pebbles up to 1| in. (4 cm.) in diameter. The thickness of the Folly Sandstone is in excess of 100 ft., but its base is not exposed. At Old Radnor, 4 miles to the south-west, this unit is missing alto- gether and Precambrian conglomerates and mudstones occur unconformably beneath the Dolyhir Limestone, the Old Radnor equivalent of the Nash Scar Limestone (Kirk, 1951, p. 56; Garwood and Goodyear, 1918, pi. vii). On the north side of the Presteigne inlier (SO/3152 6347), the contact with the Nash Scar limestone appears to be a dis- conformity; over an exposed distance of about 20 ft. the limestone rests on various sandstone beds indicating a relief of about 4 ft. Fossil communities and correlation. The stricklandiids from this area have recently been described by Ziegler (1966a). Collections from above Nash Scar quarry (Loc. 93) and from south of The Folly (Loc. 90) yielded Eocoelia hemisphaerica', the latter locality also yielded Stricklandia lens aff. progressa. These collections are both representative of the Eocoelia Community and are bothCx-C2 in age. Pentamerus Community collections have been made in the Folly area (Locs. 91 and 92) from higher beds than Locality 90, but the presence of E. hemisphaerica shows they are also of Q-Q age. The Pentamerus Community is again present in Corton House Quarry (Loc. 89) and in Folly road a few feet under the base of the Nash Scar Limestone (Loc. 88); both these collections contain S. lens aff. progressa which shows that the top of the Folly Sandstone is still Q-C2 in age. The evidence from the Folly Sandstone shows that the Nash Scar Limestone could be as old as C3 but the presence of Rhipidium (in the Garwood Collection, Geological Survey Museum) proves that part, at least, is Wenlock. An upper age limit is provided by the occurrence of Cyrtograptus symmetricus of Wenlock age in the overlying shales (Kirk 1951, p. 56). ANKERDINE HILL AND OLD STORRIDGE COMMON Text-fig. 3, cols. 15 and 16 Llandovery beds are exposed in a faulted block at Ankerdine Hill, 6| miles north of West Malvern, and again in the core of a north-plunging anticline at Old Storridge Common, 3 miles north of West Malvern (text-fig. 6). The base of the Llandovery suc- cession is not exposed in either area. Groom (1910, p. 704) applied the terms, Cowleigh Park Beds, Wych Beds, and Woolhope Shales to the Llandovery rocks of the contiguous Malvern Hills area. However, the term Woolhope Shales is confusing and in any case this relatively thin unit cannot be distinguished with certainty from the Wych Beds. ZIEGLER, COCKS, AND McKERROW: LLANDOVERY TRANSGRESSION 751 text-fig. 6. Locality map of Ankerdine Hill and Old Storridge Common, Worcestershire. This map is continuous with text-fig. 7. 752 PALAEONTOLOGY, VOLUME 11 We therefore reject the term Woolhope Shales for these Llandovery beds and regard the Wych Beds as extending up to the first appearance of the dominantly calcareous beds of the Woolhope Limestone. The latter is a gradational contact and could, of course, be diachronous. (a) Cowleigh Park Beds. Fine green sandstones and siltstones make up the lowest beds exposed on Ankerdine Hill, while at Old Storridge Common the sediments are coarser and thicker bedded (up to 2 ft.) sandstones. Lack of exposures and the presence of faults make thickness estimates very uncertain; minimum exposed thicknesses are about 200 ft. (65 m.) at Ankerdine Hill and 400 ft. (130 m.) at Old Storridge Common. Fossil communities. In the Old Storridge Common area, fossils diagnostic of the Lingula Community have been recorded by Lamont and Gilbert (1945, p. 642) near the axis of the anticline, and therefore probably low in the Cowleigh Park Beds. The Eocoelia Community occurs within a few feet of the top of the formation in Leigh Brook (Loc. 36) and in a north-flowing tributary (Loc. 35). It also occurs in lower beds south of Leigh Brook (Locs. 34 and 87). At Ankerdine Hill, the Eocoelia Community is represented in all three collections obtained (Locs. 42-4), but the relative stratigraphic positions of these are not known. Correlation. All the Eocoelia obtained from the Cowleigh Park Beds at Ankerdine Hill and Old Storridge Common are E. hemisphaerica of Q-Ca age; the type of this species comes from Ankerdine Hill (Ziegler 19666). The lower beds with the Lingula Community may either be of a very early Upper Llandovery or a late Middle Landovery age. ( b ) Wych Beds. The Wych Beds consist of shales with some siltstones (less than 10% of the formation). The base is exposed in Leigh Brook (Loc. 37) near the disused Gun- wich Mill (SO/7430 5152) west of Old Storridge Common. The contact is sharp, but, although there is palaeontological evidence of a non-sequence, there is no sign of erosion of the underlying Cowleigh Park Beds. The formation appears to be about 400 ft. (130 m.) thick in the section west of Gunwich Mill. Fossil communities. The basal few feet of Wych Beds at Gunwich Mill (Loc. 37) contain a Pentameroides Community. The non-sequence at this locality is thus followed by deeper-water conditions than in the Eocoelia Community of the Cowleigh Park Beds. The evidence suggests a progressive increase in water depth throughout Upper Llan- dovery times with a long pause in sedimentation at this locality before the deposition of the finer grained Wych Beds. A Pentameroides Community has also been obtained from loose blocks on Ankerdine Hill (Loc. 45) and may represent a similar low horizon in the Wych Beds. The Pentameroides Community is followed by the Costistricklandia Community some 10 or 20 ft. (about 5 m.) above the base of the Wych Beds; collections low in the forma- tion were obtained from a track in Coneygore Coppice (Locs. 38, 39) and in a north- flowing tributary of Leigh Brook (Loc. 40) to the south and south-west of Old Storridge Common. A continued increase in the depth of water is indicated by the occurrence of the Clorinda Community near the top of the Wych Beds to the north-east of Mousehole Bridge (Loc. 41). ZIEGLER, COCKS, AND McKERROW: LLANDOVERY TRANSGRESSION 753 Correlation. The basal Wych Beds at Gunwich Mill contain Eocoelia curtisi, showing that these beds are of C5 age and that the non-sequence covers the whole of C3 and C4 time. The collection from Ankerdine Hill also contains E. curtisi. The highest Wych Beds near Mousehole Bridge yield E. sulcata of C6 age. No fossils of C6 age have been collected on Ankerdine Hill, and it is probable that the higher parts of the Wych Beds have been cut out by faulting. MALVERN HILLS Text-fig. 3, cols. 17, 18, and 19 Llandovery rocks crop out continuously from Old Storridge Common south through West Malvern to Herefordshire Beacon (text-fig. 7). They appear again south of Here- fordshire Beacon in the area to the east of Eastnor (text-fig. 8). As in the Old Storridge Common area, we recognize two formations, the Wych Beds (named from The Wyche 1J miles south of West Malvern) and the Cowleigh Park Beds (Cowleigh Park, 1 mile north of West Malvern contains many small exposures of this lower formation). This area has been mapped by Groom (1899, 1900) and by Phipps and Reeve (1967). We differ from these authors in our interpretation of the Llandovery in two main respects: the upper limit of the Wych Beds is taken immediately below the Woolhope Limestone (see above) ; and we have no doubt that an unconformity is present below the Silurian along the west side of the Malverns (Reading and Poole 1961, 1962, Ziegler 1964). The Cowleigh Park Beds rest on Cambrian shales east of Eastnor and probably again east of Cowleigh Park; to the south-east of Cowleigh Park they are inferred to rest on Malver- nian. Between West Malvern and Ragged Stone Hill (at the south end of the Malvern Hills — SO/759365) the Cowleigh Park Beds are overlapped by the Wych Beds which rest directly on the Malvernian. (a) Cowleigh Park Beds. Exposures of this formation are scattered, but our mapping suggests that there are at least four distinct members : 1. A basal member of up to 15 ft. (5 m.) of red micaceous mudstone interbedded with several thin (4 in. (10 cm.) or less) green sandstones composed of round quartz grains about 1 mm. in diameter. These were exposed during a Geologist’s Association field trip in 1963, leader Mr. N. E. Butcher, with the aid of a mechanical digger, near the base of the Silurian to the south-west of Bronsil Castle. 2. About 100 ft. of light brown siltstones and sandstones containing fossils of the Lingula Community and occasional pebbles of Malvernian rock. This lithology crops out in both Cowleigh Park and in the Eastnor Obelisk — Howler’s Heath area (Loc. 32, 23, and 24). 3. A coarse pink sandstone about 100 ft. (33 m.) thick which forms a ridge extending south-south-east from Rough Hill (SO/758 482) into Cowleigh Park. A similar topo- graphic feature extends from Eastnor Obelisk to the eastern margin of Howler’s Heath, and, though there are no outcrops there, loose blocks of a similar pink sandstone occur, containing chips of Cambrian shale. 4. Up to 200 ft. (65 m.) of unsorted purple conglomerates with angular pebbles up to 3 cm. in diameter, interbedded with coarse green sandstones up to a foot thick. This unusual lithology is known only from Cowleigh Park (SO/7616 4723) and may be restricted to this area which is immediately adjacent to the Malvernian ridge source area. PALAEONTOLOGY, VOLUME 11 storridge! COWLEIGH i- '\park\| j NORTH • H I L L . \wESTj: : MALVERN GRE AT\ MALVERN ■WORCESTER- V .BEACON. ■; / WYCHET, CUTTING MALVERN^ VWELLSs STRATA OF UPPER SILURIAN AND DEVONIAN AGE L ' 1 V/A J WYCH BEDS EZ3 COWLEIGH PARK BEDS j'lllil CAMBRIAN (UNDIFF.) . j PRE-CAMBRIAN ROCK ^23^__F0SSIL LOCALITY T FOLD AXIS FAULT ... — UNCONFORMITY text-fig. 7. Locality map of the northern Malvern Hills, West Malvern, and Cowleigh Park areas. This map is continuous with text-figs. 6 and 8. ZIEGLER, COCKS, AND McKERROW: LLANDOVERY TRANSGRESSION 755 To the south-west of the Malverns, Holl (1865, fig. 4) and Groom (1899, p. 167) have described the unconformable relationship at the base of the Cowleigh Park Beds where the Llandovery rocks rest on the Tremadocian Bronsil Shales between the Obelisk and Howler’s Heath. Although Cambrian shales have never been mapped to the north-west of the Malverns, it is probable that a similar unconformity exists in the Cowleigh Park area; Groom (1900, p. 157) states that Cambrian shale was exposed in a well 1 mile north of West Malvern, near the foot of North Hill. Our interpretation of the geology is given in text-fig. 7. The pebbles in the conglomerates show that a landmass composed of Malvernian and Cambrian rocks was being eroded — this source area may have extended eastwards from the present-day Malvern Hills. The thickness of the Cowleigh Park Beds is perhaps 250 ft. (80 m.) at the Eastnor Obelisk. At Cowleigh Park there is probably between 350 and 400 ft. (100-30 m.), but lack of exposures and the possible presence of faults make this estimate uncertain. Fossil communities. The Lingula Community is the only fossil community represented in the Cowleigh Park Beds of the Malverns; the fossils are confined to the second (siltstone and sandstone) member. They have been collected from Cowleigh Park (Loc. 32), south of the Obelisk (Loc. 23), and north of Howler’s Heath (Loc. 24). This com- munity indicates very shallow marine conditions, and it is possible that some of the Cowleigh Park Beds are beach deposits. The coarseness of the sediments and the absence of the Eocoelia Community indicate environments nearer to the shore-line than those in correlated beds to the north-north-west near Old Storridge Common and Ankerdine Hill, and to the south-south-west at May Hill. Correlation. No fossils of value in correlation occur in the Cowleigh Park Beds of the Malverns. Comparison with Old Storridge Common to the north and with May Hill to the south suggests an early Upper Llandovery or late Middle Llandovery age. ( b ) Wych Beds. The Wych Beds are seen resting on the fourth (conglomerate) member of the Cowleigh Park Beds at the Cowleigh Park football field (SO/7616 4723); the basal 6 ft. (2 m.) are coarse sandstones (they appear to be re-worked Cowleigh Park Beds) with occasional broken Fentameroides, which probably indicate a deeper-water environment than in the Cowleigh Park Beds; these sandstones are followed by shales and green siltstones typical of the main mass of the Wych Beds. The only other two exposures of the base of the Wych Beds are on the western margin of the Malvern Hills at West Malvern (SO/7646 4554) commonly known as the Sycamore tree quarry ( Robert- son 1926, p. 168), and the Gullet Quarry (SO/7612 3811) (Reading and Poole 1961); at both these places a basal conglomerate is present, up to 2 or 3 ft. (1 m.) thick. The conglomerate consists of rounded boulders of various Malvernian rocks, ranging from 4 mm. to 80 cm. in diameter, set in a matrix of green siltstone and mudstone — typical of the Wych Beds — containing a rich fauna of brachiopods and corals (see below). The Wych Beds are poorly exposed and their thickness is determined with difficulty; estimates of about 400 ft. (130 m.) have been obtained from sections near Eastnor Obelisk and in Cowleigh Park. This is about the same thickness as at Old Storridge Common. It includes all the beds up to the base of the Woolhope Limestone, that is, where calcareous nodular beds become abundant. 3 D C 6055 756 PALAEONTOLOGY, VOLUME 11 WYNDS, ==p’OIN T>; HEREFORDSHIRI /.“.'BEACON}3.' S WIN YARD JhillI iULLE i ■ -a MIDSUMMER . • ’ ‘ H I L L - . • EASTNOR 'HOCtYRUSH* RA G G E D. 'STONE . =HILL ' HOWLERS HEATH ;HASE; '•END HILL* ^ NEW RED SANDSTONE V / / A STRATA OF UPPER SIL V//\ AND DEVONIAN AGE WYCH BEDS a COWLEIGH PARK BEDS BLACK SHALES \ 1 "-I HOLLYBUSH SANDSTONE/ 2 PRE-CAMBRIAN ROCK ( 25) FOSSIL LOCALITY FOLD AXIS NORMAL FAULT • THRUST FAULT UNCONFORMITY SCALE IN MILES text-fig. 8. Locality map of the southern Malvern Hills. This map is continuous with text-fig. 7. ZIEGLER, COCKS, AND McKERROW: LLANDOVERY TRANSGRESSION 757 Fossil communities. The fossils in the matrix of the conglomerates at West Malvern (Loc. 30) and the Gullet Quarry (Loc. 26) contain many species that are absent elsewhere in the Upper Llandovery of the Welsh Borderland, apart from Bog, Shrop- shire, and are interpreted (Ziegler, Cocks, and Bambach 1968), as a mixture of rocky bottom and soft bottom animals. Rocky bottom communities are usually in the areas of erosion, and do not frequently get preserved, but in this case the beds immediately above the conglomerate at the Gullet Quarry contain a Pentameroides Community (Loc. 27), and the fauna on the rounded boulders probably lived well below low-tide level and uncovered by sediment for a long period. At Cowleigh Park (SO/7616 4723) the basal 6 ft. (2 m.) of the Wych Beds are coarse sandstones containing broken Pentameroides, and at the same exposure (Loc. 33) the Costistricklandia Community was found in typical Wych Beds lithology a few feet above the highest sandstone. Higher horizons in the Wych Beds also yield the Costistricklandia Community at Gullet Quarry (Locs. 28, 29) and on the south side of Howler’s Heath (Loc. 25). The Clorinda Community has been found at West Malvern (Loc. 31 ) indicating, as at Old Storridge Common, still deeper water in the upper part of the Wych Beds. Correlation. At Gullet Quarry Eocoelia curtisi is present in the basal Wych Beds (Locs. 26, 27). This species is of C5 age, as is Costistricklandia Ur at a alpha which occurs in higher beds at Gullet Quarry (Locs. 28, 29). The Cowleigh Park (Loc. 33), and West Malvern (Loc. 31) collections contain C. lirata typica, showing that, in the Malverns, the greater part of the Wych Beds is of CG age. WOOLHOPE Text-fig. 3, col. 26 Squirrel and Tucker (1960) proposed the names Lower Haugh Wood Beds and Upper Haugh Wood Beds for the Llandovery rocks exposed in the Woolhope dome, east- south-east of Hereford. The former are at least 280 ft. (90 m.) thick, their base unexposed, and are lithologically similar to the Wych Beds of the Malverns. The Upper Haugh Wood Beds are 30 ft. (10 m.) thick; the basal 20 ft. (7 m.) are purple and green muddy siltstones, but more calcareous beds are present in the top 10 ft. (3 m.). Fossil communities. In the Lower Haugh Wood Beds the Costistricklandia Community is represented by collections from Kidley Coppice (Loc. 85) and Rudge Wood (Loc. 86). Judging from the faunal lists of Squirrel and Tucker this community is also present in the Upper Haugh Wood Beds. Correlation. C. lirata typica occurs in both the Lower and Upper Haugh Wood Beds, indicating a C6 age. The Petalocrinus Limestone is present in the highest 10 ft. (3 m.) of the Upper Haugh Wood Beds; it was found by Pocock (1930, pp. 60-1) at equivalent horizons in the May Hill and Malvern areas, but these exposures are now obscure. It is very close to the lithological boundary at the base of the Woolhope Limestone, and we conclude that this limestone is of late C6 or early Wenlock age. MAY HILL Text-fig. 3, col. 27 Llandovery Beds are exposed in the core of an anticlinal structure at May Hill, 8 miles west of Gloucester (text-fig. 9). Two formation names, Huntley Hill Beds and 758 PALAEONTOLOGY, VOLUME 11 text-fig. 9. Locality map of the May Hill Inlier (after Lawson 1955, with minor changes). Yartleton Beds, were proposed by Gardiner (1920) who also proposed the term Huntley Quarry Beds for ‘Fine and somewhat coarse grits, containing lapilli’ for beds in the vicinity of the quarry. At the present time, volcanic flows with interbedded red-beds are seen in Huntley Quarry, and these apparently underlie coarse sandstones with angular fragments of the volcanic rocks which grade up into fossiliferous Llandovery sediments. ZIEGLER, COCKS, AND McKERROW: LLANDOVERY TRANSGRESSION 759 The contact is obscure and faulted but it is probable that the Llandovery lies unconform- ably on these volcanics, which might possibly be correlated with similar volcanics in the Malvern Hills assigned to the Pre-Cambrian Warren House Series (Groom 1910). We regard the coarse grits in the vicinity, which have sometimes been included in the Huntley Quarry Beds, as the basal part of the Huntley Hill Beds, and propose the term Huntley Quarry Volcanics for the volcanic rocks and their associated red beds. We also differ from the current literature concerning the relationship of the Yartleton Beds and the Woolhope Limestone. It has long been known (Phillips 1848, pp. 184-5) that a series of sandstones occur within the Woolhope Limestone sequence at Little London (SO/700 184) in the southern part of the inlier. Our mapping shows that the sandstone at Old Oaks Farm Quarry (Loc. 21) and the stream section near Hill Farm (Loc. 22) in the north-west and central parts of the inlier are also within the Woolhope sequence; in these two areas Lawson (1955) mistakenly referred these Woolhope sand- stones to the Llandovery. Except for these minor points, we agree with Lawson’s map of the Llandovery beds. (a) Huntley Hill Beds. Lawson’s ( 1 955) minimum thickness of 600 ft. (200 m.) is probably an underestimate. The Little London stream section (SO/7085 1832 to SO/6997 1834) where the beds, with the exception of one outcrop, dip continuously to the west, sug- gests a thickness of about 850 ft. (280 m.). The oldest beds exposed in this section are on strike with the basal beds at Huntley Quarry, and are therefore near the base of the formation. The Huntley Hill Beds are mostly grey coarse sandstones and conglomerates; they weather to a buff colour and are often reddened in areas close to the Triassic cover. The sandstone beds are normally one to 2 ft. thick; they show lateral changes in thick- ness, and they have little internal structure except towards the top of the formation where thinner bedded sandstones are present with fine parallel laminations. Shales are rare, but a few tens of feet of shale are exposed in the Little London stream (SO/7057 1837) at about 300 ft. above the base of the formation. The pebbles in the conglomerates include the distinctive quartz-orthoclase plutonic rocks of the Malvern Hills. Volcanic pebbles are also present and could have been derived from other Malvern rocks or, in the case of the angular andesite prophyry pebbles, directly from the Huntley Quarry Volcanics. There is no doubt that adjacent pre- Cambrian rocks were being eroded at this time. Fossil communities. The lowest 350 ft. (115 m.) of the Huntley Hill Beds are largely barren of fossils, but Lingula pseudoparallela has been reported from the basal sediments of Huntley Quarry (Symonds 1872, pp. 147-8). Eocoelia hemisphaerica and other repre- sentatives of the Eocoelia Community have been collected in the Little London stream section at about 350 ft. (Loc. 6) and 410 ft. (Loc. 7) above the supposed base of the formation ; closely similar assemblages also occur near Dursley Cross (Loc. 5) and on the south-east side of May Hill (Loc. 4). At higher stratigraphical levels there is a reversion to the Lingula Community; about 440 ft. (140 m.) above the base in the Little London stream (Loc. 8), and in old quarries on Brights Hill (Locs. 9, 10). The highest Huntley Hill Beds are poorly exposed; a single fossil locality near old quarries at the top of Nottswood Hill (Loc. 11) has yielded the Eocoelia Community, with E. intermedia at about 610 ft. above the base of the formation. 760 PALAEONTOLOGY, VOLUME 11 Correlation. The lowest, largely unfossiliferous 350 ft. (115 m.) may be pre-Upper Llandovery in age. The beds between 350 and 410 ft. (115-35 m.) from the base of the formation in the Little London stream section are known to be of C4 or C2 age from the presence of E. hemisphaerica, as are the beds that have been correlated with these at Dursley Cross and May Hill. The Nottswood Hill quarries at around 610 ft. (200 m.) from the base contain E. intermedia of C3 or C4 age. ( b ) Yartleton Beds. The lower contact of the Yartleton Beds is not exposed, but there appears to be a gradational transition from the coarse sandstones of the Huntley Hill Beds to the fine laminated sandstones (typically 3-4 in. thick), siltstones and shales of the Yartleton Beds. The formation is about 500 ft. (160 m.) thick. Fossil communities. Fossil localities are abundant in the Yartleton Beds, but there are no well-exposed sections showing the stratigraphic relationships between exposures more than a few feet apart, as most of the outcrops occur along streams in the vicinity of faults. The fossil collections however can be arranged stratigraphically by comparing the Eoco- e/ia, stricklandiids and pentamerinids with contemporaneous sections at Llandovery and in other parts of the Welsh Borderland. Table 1 summarizes the data for eleven localities at May Hill. The method used for calculating percentages is described in Ziegler, Cocks, and Bambach (1968, p. 4). Except for the two collections (Locs. 18, 19) within 100 ft. (33 m.) of the Woolhope Limestone, all the Yartleton fossils belong either to the Pentameroides or Costistrick- landia Communities or to mixtures of the two. Some changes in community composition apparently took place in a short period of time, for example, Locality 16, with 34% Pentameroides and 44% Costistrieklandia contains elements of both communities but occurs a few inches above Locality 15 which consists of Costistrieklandia with 8% Clorinda. Locality 16 may have a fauna from the boundary between the Costistrieklandia and Pentameroides Communities, or it may be due to mixing of faunas after death. Locality 18 is interpreted as a death assemblage; it contains fairly abundant Eocoelia together with some deeper- water species. The highest locality in the Yartleton Beds, 19, is assigned to the Clorinda Community and represents deeper water than the earlier faunas. Correlation. The majority of the Yartleton Beds are of C5 and C6 age (Table 1). It is possible that the basal parts are C4; Locality 12 has no Eocoelia nor pentamerinids, and the stricklandiids are fragmentary and cannot definitely be assigned to Stricklandia or Costistrieklandia, so it could represent an earlier horizon than the remaining collections. At May Hill there is a transition between the Llandovery and Wenlock Beds. Phillips (1848, pp. 74-5, 184-5) arbitrarily drew the contact at the first appearance of limestone. Unfortunately the Woolhope Limestone and the overlying Wenlock Shales in the southern part of the Welsh Borderland have not yet yielded any graptolites that allow correlation with the graptolite zones and Phillips’ arbitrary criterion must continue for the time being. TORTWORTH Text-fig. 3, cols. 28 and 29 Curtis (1955) has assigned the Llandovery rocks of this area (text-fig. 10) to two for- mations: the Damery Beds and the Tortworth Beds. These are separated by the ‘Upper ZIEGLER, COCKS AND McKERROW: LLANDOVERY TRANSGRESSION 761 Trap’, an andesite flow with a fossiliferous ashy tuff bed on its upper surface (Reed and Reynolds 1908, p. 515, Reynolds 1924). The basaltic ‘Lower Trap’, considered intru- sive by Reynolds (1924) but extrusive by Curtis, occurs at the base of the Damery Beds, resting unconformably on Tremadocian shales. Both basalts thin towards the north-west, and appear to be absent north of Stone. ( a ) Damery Beds. This unit is comprised of about 500 ft. (160 m.) of alternating fine- grained sandstone and shale beds (Curtis, 1955), with the former being dominant in the lower part of the unit. Table 1. Critical community and dating elements of the collections from the Yartleton Beds (E, Eocoelia ; P, Pentameroides', S, Costistricklandia ; and C, Clorinda) Community elements Com- Dating Stratigraphical Section Loc. E P 5 C munity elements Date notes Stream 450 yd. 19 7% 0% 3% 10% C E. sulcata C6 Within 100 ft. of W. of Hay Farm C. lirata typica Woolhope Lst. (NE. side of 18 18 0 1 1 E. sulcata G Along strike from May Hill) C. lirata typica Loc. 19. 84 3 0 8 1 s E. sulcata c6 10 ft. above Loc. C. lirata typica 83. 83 12 18 26 0 P/S E. sulcata Pentameroides sp. C. lirata typica c6 (See above.) 17 0 0 42 0 s C. lirata alpha G Stream S. of Pit- 20 4 0 4 0 s E. sulcata ? G man’s Farm (W. C. lirata typica side of May Hill) 16 0 34 44 0 P/S Pentameroides sp. G A few inches C. lirata alpha above Loc. 15. 15 0 0 29 8 s C. lirata alpha C5 (See above.) Stream NW. of 14 3 37 7 0 p E. curtisi ? G Hill Farm (SE. side of May Hill) Pentameroides sp. Stream S. of Holly- 13 3 34 19 0 p E. curtisi ? G bush Farm, Dursley Cross 12 0 0 34 0 s Pentameroides sp. ?C5 Fossil Communities. The Eocoelia Community is present in the basal 130 ft. (43 m.) of the Damery Beds. Collections have been made within a few feet of the top of the Lower Trap in Damery Quarry (Loc. 82), and in higher beds at Charfield Green (Loc. 1) and near Damery Bridge (Loc. 77). The Pentameroides Community has been recognized in two collections, 10 ft. (3 m.) stratigraphically apart in the middle of the formation at Charfield Green (Locs. 78, 79). Still deeper water is indicated by the presence of the Costistricklandia Community in the top two-thirds of the Damery Beds; collections have been made on the road south of Damery Bridge at about 165 and 176 ft. (54- 7 m.) above the base of the formation (Locs. 80, 81), and in the railway cutting at Charfield Green (Loc. 2) within 25 ft. (8 m.) of the top of the formation. Correlation. Eocoelia curtisi occurs in all the collections from the lowest third of the Damery Beds, and Costistricklandia lirata alpha is present in the collections from the upper two-thirds of the formation. These show that all the Damery Beds are of C5 age. The stream south of Charfield Green (Loc. 2) is the type locality of E. curtisi (Ziegler 19666, p. 537). 762 PALAEONTOLOGY, VOLUME 11 text-fig. 10. Locality map of the Tortworth Inlier (after Curtis 1955). (b) Tortworth Beds. The lithology of these beds is similar to the shales and fine-grained sandstones of the Damery Beds. Exposures are rare except near the base; Curtis (1955) estimates the thickness at 200 ft. (65 m.), but the upper contact appears to be gradational with an impersistent limestone forming the base of the Wenlock in the western parts of the Tortworth inlier. Fossil communities. The only collections from the Tortworth Beds are from near the base. In Daniel’s Wood (Loc. 3), the Eocoelia Community is present within a few feet ZIEGLER, COCKS, AND McKERROW: LLANDOVERY TRANSGRESSION 763 of the Upper Trap which is 150 ft. (50 m.) thick here; but north-east of Stone (Loc. 76) the Costistricklandia Community occurs in the basal foot of the Toriworth Beds where the basalt is only about 15 ft. (5 m.) thick. The different thicknesses of the lava appear to have controlled the local depth of water; Costistricklandia Community depths are present before and after the flow where it is thin, but where the flow is 150 ft. (50 m.) thick the deeper-water Costistricklandia Community at the top of the Damery Beds is replaced by the shallower-water Eocoelia Community above the andesite (Ziegler 1965). Correlation. Eocoelia sulcata is present in both collections from the base of the Tort- worth Beds; Costistricklandia lirata alpha is also present north-east of Stone (Loc. 76). The presence of C. lirata alpha and E. sulcata indicates that the horizon must be near the C5-C6 boundary. WALSALL AND GREAT BARR Text-fig. 3, col. 6 Butler (1937) has described Llandovery rocks in the Walsall Borehole, and Jukes (1853) and Eastwood et al. (1925, p. 11), mention a very small area of the Llandovery rocks at Great Barr near the Eastern Boundary Fault of the South Staffordshire Coalfield. The Walsall Borehole penetrated Wenlock Limestone, Wenlock Shale, Ban- Limestone, and below this, some 4577 ft. (150 m.) of shales, mudstones, siltstones, and sandstones before hitting Cambrian quartzite with a dip differing by 8° to 15° from the Silurian. Some of the beds beneath the Barr Limestone are Wenlock in age as Cyrto- graptus murchisoni has been recorded 36 ft. (12 m.) below this unit (depth 8317 ft.), but 747 ft. (24 m.) below the Barr Limestone (depth 870 ft.) are purple and green shales, similar to the Hughley Shales of Shropshire, and we prefer to draw the Wenlock-Llan- dovery boundary at this point. Butler records 24 bentonitic clays above depth 870 ft., and 27 bentonites below this depth; one bed is 7 in. (18 cm.) thick, the remainder range from 2 in. down to ^ in. (5 cm.-L5 mm.). Fossil communities. The beds from 28 to 303 ft. (9-5-100 m.) above the base of the Silurian (depths 1225 to 950 ft.) contain Costistricklandia ; in this sequence, Pentameroides (previously recorded as Pentamerus oblongus ) is known only over a short thickness (233-48 ft. (76-81 m.) above base, or depths 1020 to 1005 ft.), so it appears probable that the Llandovery beds accumulated mainly in the moderate depths characteristic of the Costistricklandia Community. This interpretation agrees well with Butler’s (1937, p. 251) picture of the depositional environment. The Costistricklandia Community also occurs at Great Barr where a Llandovery quartzite is exposed south-south-east of the house formerly known as the Australian Arms (SP/038957). Correlation. Costistricklandia lirata alpha occurs between 28 and 93 ft. (9-5-31 m.) above the base of the sequence (depths 1225 to 1160 ft.) (Butler, 1937, p. 247), showing that the oldest Llandovery in this area is of C5 age. C. lirata typica has been recorded between 123 and 283 ft. (41-93 m.) above the base (depths 1130 to 970 ft.), which proves a C6 age for these beds. Graptolites occur at several horizons; those that Miss Elies examined from below 120 ft. (39 m.) from the base (depth 1133 ft.) are ‘suggestive of the Upper Valentian’, and those below 27 ft. (9 m.) from the base (depth 1226 ft.) suggest a horizon at least as low as the M. griestoniensis Zone (Butler 1937, p. 246). At the Great Barr exposure, we have collected C. lirata typica. 764 PALAEONTOLOGY, VOLUME 11 RUBERY Text-fig. 3, col. 8 Llandovery rocks rest with slight angular unconformity on the Lickey Quartzite (Cambrian) to the north and south-east of Rubery (Wills et al. 1925, 1938, Eastwood et al. 1925, p. 12); the latter authors infer that it occurs in hollows in the quartzite. Wills et al. (1925, p. 68 and 1938, p. 177) describes the Rubery Shales Series as consisting of a minimum of 40-50 ft. (13-16 m.) of shales and thin sandstones with some calcareous beds; these rest on 100 ft. (33 m.) of Rubery Sandstone Series which are coarse and fine sandstones, some shale and a basal conglomerate. Fossil communities. Costistricklandia is present in both the Rubery Sandstone series and the Rubery Shale Series; the faunal list of Wills et al. (1925, pp. 72-3) suggests that the Costistricklandia Community is present throughout the whole Llandovery sequence in this area, although Pentameroides does occur in some beds. Correlation. Costistricklandia lirata alpha is abundant in the Rubery Sandstone Series (St. Joseph 1935, p. 419) and Pentameroides also occurs (previously recorded as Pentamerus oblongus). The higher Rubery Shale Series contains C. lirata alpha , Mono- graptus marri and M. nudus. It is concluded that the section described by Wills et al. (1925, p. 68) is all of C5 age. The fossils listed (Eastwood et al. 1925, pp. 14-15) as coming from Wenlock beds near Rubery Hill Asylum (SO/992778) include undoubted Llandovery fossils, but we have not examined this locality. COVENTRY Text-fig. 3, col. 9 Shotton (1927) records Upper Llandovery sandstone pebbles from the Corley Con- glomerate of Upper Carboniferous age, which is exposed from Coventry to Corley and Hollyberry End (SP/262842) 6 miles to the north-west. Shotton has shown that, approaching Coventry from the north-west, the proportion of Llandovery sandstone pebbles increases from 46% at Hollyberry End, to 70% at Corley (SP/304850) and 85% at King Street, Coventry (SP/333797). This traverse along the present outcrop of the Corley Conglomerate runs obliquely to a south-south-eastward extension of the Nuneaton Ridge (composed of Cambrian and Precambrian rocks), and Shotton suggests (1927, p. 616) that this ridge was capped by a strip of Upper Llandovery rocks (as at Walsall and Rubery) which provided the pebbles. It is probable that the source area extended north-north-west from around Binley, 3 miles east of Coventry. Fossil communities. The faunal list (Shotton 1927, pp. 609-10) contains 55 species, which, with one or two possible exceptions, are typical of, or known in, the Eocoelia Community. Stricklandiids, which are dominant to the west at Rubery and Walsall, are absent, as are Pentameroides and other elements of the offshore communities. Correlation. Examination of Eocoelia in pebbles (some kindly lent from the Birmingham University collections by Professor Shotton, others in the Sedgwick Museum, Cam- bridge) show the present of E. intermedia , E. curtisi, and E. sulcata , which indicates ZIEGLER, COCKS, AND McKERROW: LLANDOVERY TRANSGRESSION 765 ages from C3_4 through C5 to C6. This is the only record of Silurian beds as old as C3_4 to the east of the Malvern line. LOWER LEMINGTON Text-fig. 3, col. 20 Silurian rocks are recorded below Coal Measures in Strahan’s (1913) description of a borehole 350 yd. W. 22° N. of Lower Lemington Church, 1 mile north-east of Moreton-in-Marsh. Grey shales with sandstone and limestone occurred from a depth of 1546 ft. (510 m.) to the completion of the hole at 1700 ft. (558 m.); allowing for a dip of 22°, the strata proved are about 140 ft. (46 m.) thick. Fossil communities. Examination of the fossils in the Geological Survey collections (JP3143-61) from between 1590 and 1682 ft. show an Eocoelia Community present throughout this thickness. Correlation. The Eocoelia present are all E. sulcata , indicating a C6 age for all the fossiliferous Silurian beds that were penetrated. EASTERN MENDIPS Curtis (1955) summarizes the earlier work of Reynolds (1907, 1912, 1929) on the stratigraphy of the Silurian inlier which extends 3^ miles westwards from Downhead, Somerset : 3. Wenlock mudstones with some micaceous sandstones (120 ft. — 40 m.). 2. Andesites and tuffs (at least 400 ft. — 130 m.). 1. Tuffs (110 ft. (36 m.) seen). These oldest beds have, in the past been assigned to the Upper Llandovery. Fossil communities. A collection from a field 1 mile west of Downhead (Loc. 94) yielded a dominance of Salopina and Sphaerirhynchia, which suggest a near-shore environment for beds which our interpretation of the literature suggests are beneath the andesites. Correlation. It is probable that there are no Llandovery beds exposed in the inlier. Collections from tuffs (ST6752 4584) suggest an environment similar to the Eocoelia Community of the Llandovery, but no Eocoelia or other distinctive Llandovery forms have been found by us or recorded in the literature. We have examined the specimens (GSM DEW 3833-935) upon which Green (1962) based his report of beds attributable to the Llandovery, and conclude that they are also of Wenlock or Ludlow age. We thus tentatively consider these beds to be of Wenlock age. This conclusion is supported by the presence of Acaste downingiae (Reynolds 1907, p. 227, Curtis 1955, p. 7) which we believe to be absent in the Llandovery. RUMNEY The Silurian inlier at Rumney (Rhymney), 2\ miles north-east of the centre of Cardiff', has been described by Sollas (1879). The Silurian beds in this inlier have always been assigned to the Wenlock and Ludlow, although Sollas recorded Pentamerus and Stricklandia. The problem has been investigated by Dr. M. G. Bassett, who kindly writes: ‘I am certain that all the beds are Wenlock-Ludlow. The best exposure at present is at 766 PALAEONTOLOGY, VOLUME 11 Penylan Quarry, where virtually the oldest beds brought to the surface in the inlier are exposed. These beds are certainly below the Rumney Grit, and hence older than the horizons from which Sollas quoted his diagnostic Llandovery brachiopods. The beds at Penylan contain abundant Meristina obtusa, which I consider to be a good indicator of late Wenlock age. This is supported by the occurrence of other fossils which could only be Wenlock or Ludlow in age.’ In addition there are many Wenlock brachiopods from the inlier at the National Museum of Wales which bear old labels mis-identifying them as ‘P. oblong us' and ‘ Stricklandia' . BRE1DDEN HILLS Text-fig. 3, col. 4 Watts (1885, pp. 536-7) gave the first description of the Llandovery beds of the Breid- dens, and Wade (1911, p. 419) subdivided these beds into a lower unit, the Cefn Beds, and an upper unit, the Buttington Shales. The use of local names is preferable to the Shropshire names employed by Whittard (1932, pp. 880-2) in his description of this area. These beds are exposed sporadically along a 5-mile strip of country north-east of Buttington Station; the largest exposures are in the Brick Works (SJ/265 100) a quarter of a mile east of the station, and adjacent road sections. (a) Cefn Beds. Whittard (1932) states that these consist of about 300 ft. (100 m.) of sandy mudstones with calcareous sandstones and occasional conglomeratic beds, but in the south-west the calcareous beds are more common. Ripples exposed at the top of the formation, over a large face in Buttington Brick Works, give a direction of movement from east to west. Fossil communities. Whittard (1932, p. 881) mentions two fossil localities: north of Middletown in the centre of the outcrop (SJ/3008 1276) within 50 ft. (16 m.) of the base, and at the south-west end, south-east of Buttington Station (SJ/2633 0997) within 100 ft. (33 m.) of the base of the formation. Our collections show that the Stricklandia Community is present in both places. Correlation. The presence of Stricklandia lens cf. intermedia suggests a Middle Llan- dovery age for the lower part of this formation. ( b ) Buttington Shales. Whittard (1932, p. 881) described maroon, green, and blue mudstones with some calcareous beds up to 6 in. (15 cm.) thick, becoming uniformly maroon towards the top; he estimated their thickness to be about 350 ft. (115 m.). The upper contact is defined by an abrupt termination of the maroon coloured beds; above this Cocks and Rickards (in press) record the presence of a complete succession of Lower Wenlock graptolite zones at Buttington Brick Works. Fossil communities. The Buttington Shales have few fossils, but we have a collection, from near the top of the sequence in the Brick Works, in the Clorinda Community. Whittard (1932, p. 882) also found fossils in loose blocks at the north-east end of the outcrop; his faunal list is also typical of the Clorinda Community. Correlation. It is not possible to state what horizon within the Llandovery is represented by these two collections from the Buttington Shales, but they are overlain with apparent conformity by a basal Wenlock centrifugus Zone fauna. ZIEGLER, COCKS, AND McKERROW: LLANDOVERY TRANSGRESSION 767 WELSHPOOL Text-fig. 3, col. 3 Wade (1911) described two formations of Llandovery age in the Welshpool area, the Powis Castle Beds and the higher Cloddiau Beds. The poorly exposed outcrop extends from Belan (2 miles south-west of Welshpool) to the Llansantffraid ym Mechain area, north-east of Meifod (see below). (a) Powis Castle Beds. This unit consists of 100 ft. (33 m.) of massive bedded con- glomerates; sandstones, shales and limestones are also present (Wade 1911, p. 431). It rests on Ashgill beds {Phillip sinella parabola age) in the Gwern y Brain (SJ/219 128) (Cave 1965, p. 282) and at the Laundry (SJ/197 103) (ibid., p. 295). Towards Welshpool the uppermost Ordovician beds are overstepped and the Powis Castle Beds rest on Lower Caradoc rocks (Harnagian Stage). In the Gwern y Brain, the Powis Castle Beds contain angular pebbles of these Ashgill and Caradoc units. Fossil communities. A collection from the Gwern y Brain (SJ/2186 1283) within 10 or 15 ft. (3-5 m.) of the base of the Powis Castle Beds has yielded nothing but derived fossils from the underlying units, but the angularity of the pebbles and boulders is very sug- gestive of deposition in non-marine conditions. We have found the Cryptothyrella Com- munity, a probable early Llandovery equivalent of the Eocoelia Community, in a loose block in the southern part of the area, 600 yd. west of Belan Locks (SJ/2100 0518). On Cherry Tree Bank (SJ/2226 0831) on the north edge of Welshpool, we have collected Clorinda from the higher beds of the unit, and this would suggest that some of the Powis Castle Beds accumulated in deeper water; this conclusion is strengthened by the presence of Stricklandia lens in the conglomerates in the north of Powis Castle Park. In fact, the full depth range of communities may be present in the formation. Correlation. The fact that the overlying Cloddiau Beds contain a Lower Llandovery graptolite fauna (vesiculosus Zone) suggests that the Powis Castle Beds are of Lower Llandovery age, and suggests a short pause in sedimentation between the Ashgill beds in the Gwern y Brain and the earliest Silurian. The presence of Stricklandia lens indicates a Llandovery as opposed to a late Ordovician age for the unit. ( b ) Cloddiau Beds. Around Welshpool, upper Silurian beds rest directly on the Powis Castle Beds, but from Cloddiau northwards, some 250 ft. (80 m.) of mudstones and siltstones are seen above the conglomerates. Fossil communities. A collection from the Cloddiau Beds at Y Frochas (SJ/1955 0824) yielded the Clorinda Community. Correlation. Wade (1911, p. 434) reported Climacograptus innotatus, C. cf. rectangularis and C. medius from the lower part of the formation near Cloddiau. These indicate the vesiculosus Zone, but as they were found only 20 ft. (7 m.) above the base of the forma- tion, it is possible that the Middle and Upper Llandovery is also present — the Cloddiau Beds extent upwards into Wenlock strata with no change in dip. 768 PALAEONTOLOGY, VOLUME 11 MEIFOD Text-fig. 3, col. 2 The Llandovery outcrop extends northwards from Welshpool for 7 miles to the Llansantffraid ym Mechain area (Whittington, 1938), and thence changes direction south-westwards to Meifod (King, 1928). King recognized four formations which he labelled Vx, V2, V3, and VS and Whittington employed this terminology at Llansant- ffraid ym Mechain. (a) Craig-wen Sandstone (F2). This basal unit is a 30-ft. (10 m.) sandstone with some large calcareous concretions near the base. The sandstone fills hollows cut in Ashgill mudstone south-west and west of Meifod at Graig-wen and Glan-yr-afon-isaf (King 1928, pp. 682-3). These hollows are a foot or more in depth and contain angular fragments of Ashgill mudstone. Fossil communities. King (1928, pp. 686-7, 699) records two communities from the Graig-wen Sandstone. In the calcareous beds fragmentary corals and brachiopods are present, while the sandstone contains 70% orthids, 20% Cryptothyrella (the Meristina crassa of earlier reports), and about 10% rhynchonelloids. Our collection from the basal 1| ft. (50 cm.) of the unit at Graig-wen Quarry (SJ/0991 0919) has yielded the Crypto- thyrella Community, the probable early Llandovery equivalent of the Eocoelia Com- munity. Correlation. The fossils of this unit are not diagnostic as to age, but the underlying Ash- gill unit has yielded graptolites normally associated with the Dicellograptus anceps zone (King, 1928, p. 698) and the overlying formation has yielded late Lower Llandovery graptolites, so the Graig-wen Sandstone is probably of early Lower Llandovery age. ( b ) Blue silty mudstones (TV)- Blue-black micaceous shales and hard siltstones (V2a) come at the base of this formation west of Meifod; higher beds (V2u and V2c) in this 630-ft. (205 m.) formation are blue mudstones with silty siliceous bands and calcareous nodules (King 1928, pp. 687-8). Five miles north-east of Meifod, Whittington (1938, pp. 438—42) describes blue mudstones and limestones near Tre-wylan House and Gelli Farm (which he correlates with King’s V2b beds) resting on Ashgill. Similar beds occur at Godor and it would seem likely that the outcrop is continuous with that of Meifod. Beds like King’s V2c are recorded at Godor and Sarnau (Whittington 1938, pp. 440-1). Fossil communities. King (1928, p. 699) lists Stricklandia and Clorinda [ Barrandella ] as common in V2b; and Clorinda as common in V2o. This suggests deeper water than in the Graig-wen Sandstone. Correlation. In the lower part of this unit, Climacograptus normalis, ‘and what Dr. G. L. Elies thinks may be a specimen of Monograptus acinaees ’ have been found (King 1928, p. 688). These species would suggest the zone of M. cyphus (late Lower Llandovery), but the unit may extend into the Upper Llandovery as the lowest graptolite fauna in the overlying formation is in the turriculatus Zone. (c) Pale grey and purple-red shales ( V3). King’s (1928, p. 690) V3 beds start with an abrupt change of colour from grey-blue mudstones to pale dove-grey fine mudstones; they consist of 330 ft. (110 m.) of pale grey and red shales. ZIEGLER, COCKS, AND McKERROW: LLANDOVERY TRANSGRESSION 769 Fossil communities. Apart from some fragmentary trilobites and brachiopods near the base, graptolites are the only fossils present. Correlation. The zones present range from turriculatus to crenulata (King 1928, p. 690) and represent the upper two-thirds of the Upper Llandovery. The red shales occur in the crispus and griestoniensis Zones; they are the oldest dated red beds known in the Welsh Llandovery succession. (d) Passage Beds {VS). King (1928, p. 691) records 60 ft. (20 m.) of blue-grey mudstones above V3 beds. Fossil communities. Benthic forms are absent from this unit. Correlation. King found Monoclimacis crenulata in these beds to the north-east of Meifod, so they are assigned to the topmost Upper Llandovery. They are followed by lowest Wenlock beds with a similar lithology. TRANNON AND LLANIDLOES Text-fig. 3, cols. 1 and 10 These areas are in the middle of the graptolitic-greywacke facies of central Wales and have been included here because they contrast with the dominantly shelly facies of the Welsh Borderland areas already described. Wood (1906) and Jones (1945) have described the stratigraphy of the areas. At Trannon (called Tarannon by Wood) at least 3650 ft. (1200 m.) of Llandovery sediments (we use Llandovery to include all the beds up to the base of Wood’s murchisoni Zone) are present, and 9670 ft. (3175 m.) at Llanidloes. In both areas the whole succession consists of graptolitic shales with grey- wackes at some horizons. The base of the Silurian is not quite reached near Trannon, but at Llanidloes a sharp but comfortable base is present. Greywackes are especially abundant in the griestoniensis Zone. Bassett (1963, pp. 57- 9) interprets the Talerddig Grits of Trannon and the contemporary Moelfre Group of Llanidloes as turbidites and comments on the problems in estimating the depth of water in which they were deposited. We would consider that these turbidites could have been laid down over a very large depth range, provided that the depth was greater than that of the Clorinda Community, which is completely absent around Trannon. The highest Llandovery consists of purple and green mudstones which contain a fauna characteristic of the crenulata Zone. There is a gradational passage upwards into the Wenlock beds. BUILTH WELLS Text-fig. 3, col. 23 Upper Llandovery rocks occur in a very narrow discontinuous belt running north-east from Park Wells (SO/027 519 — 1 mile west-north-west of Builth Wells) for a distance of 4 miles (Jones 1947). The beds consist of about 50 ft. (16 m.) of coarse sandstone and mudstones, resting unconformably on Ordovician beds, and overlain unconformably by beds of Lower Wenlock age. The Wenlock overstep is considered to be responsible for the discontinuous outcrops (Jones 1947, p. 35). Fossil communities. A collection from Trecoed (SO/0527 5516) has yielded an abundance 770 PALAEONTOLOGY, VOLUME 11 of Pentamerus and Stricklandia , suggesting a mixture of these two communities. These came from a coarse sandstone about 14 ft. (5 m.) above the base of the Llandovery. Correlation. S. lens ultima is present, indicating a C3_4 age for this horizon. GARTH Text-fig. 3, cols. 21 and 22 The Garth area is situated some 1 5 miles north-east of Llandovery and is part of the same belt, which, during Llandovery times, formed the junction of the stable shelf to the south and east and the subsiding basin to the north and west. The two columns representing this area were selected to illustrate some of the many lateral changes in thickness and lithology in the Llandovery rocks of the Garth area. Andrew (1925) divided the Lower Llandovery into three formations, Aa, Ab, and Ac; designated the Middle Llandovery B; and divided the Upper Llandovery into Ca, Cb, Cc, and Cd. (а) Lower Llandovery. The basal Llandovery sandstones and conglomerates rest on Bala mudstones with a sharp boundary; these are followed by mudstones with occasional sandy beds, all totalling 1880 ft. (600 m.) in the centre of the area, but thinning to 1200 ft. (400 m.) in the south-west (Andrew 1925, pp. 392, 402). Fossil communities. No fossils are recorded in Aa. In the southern part of the area, the Stricklandia Community is present in Ab, and both Stricklandia and Clorinda occur in Ac, but the shelly faunas become sparse or absent in the north (Andrew 1925, pp. 396-7), that is, towards the basin. Correlation. Andrew (1925, p. 403) considered ‘on faunal and lithological grounds’: that his units ‘Aa, Ab, and Ac appear accurately to represent A4, A,, and A3 of the southern part of the Llandovery area’. It is of importance to note that the top 20 ft. of the ‘Bala’ beds of Garth contain ‘ Glyptograptus cf. persculptus and Mesograptus cf. modestus var. parvulus, which . . . are usually associated with the basal zone of the Valentian’ (Andrew 1925, p. 392). This fauna indicates the persculptus Zone, the base of which is now conventionally taken (e.g. Toghill 1968) as the base of the Llandovery in the graptolite sequence. If Andrew was correct in equating his unit Aa with the Al sandstone at Llandovery itself, then the base of the formation is probably diachronous. The latter conclusion is supported by recent unpublished sedimentological work by Dr. M. A. Woollands. (б) Middle Llandovery. Only a small thickness of mudstones is present in a restricted area in the south-east of the Garth area (Andrew 1925, p. 397). The few specimens recorded are suggestive of the Clorinda Community. Andrew suggested a correlation with B3 of Llandovery on the basis of lithological resemblance (Andrew 1925, p. 403), but its age is still not definitively known. (c) Upper Llandovery. The lower three Upper Llandovery units of Andrew (Ca, Cb, Cc) comprise a sequence of mudstones, argillaceous sandstones, and rather prominent conglomerates which is very variable in thickness (0-300 ft., 0-100 m.). The upper unit (Cd) consists of 150-1000 ft. (50-330 m.) of olive-green and purple mudstones, similar in colour and lithology to the high Llandovery of many other areas; it overlaps earlier ZIEGLER, COCKS, AND McKERROW: LLANDOVERY TRANSGRESSION 771 beds, coming to rest on Ab at the extreme south-western and north-eastern ends of the area, according to the mapping of Andrew. Ca passes up without a break into beds containing the basal Wenlock zone of Cyrtograptus murchisoni (Andrew 1925, p. 400). Fossil communities. Andrew's faunal lists (1925, pp. 398-9) show that Ca contains a Pentamerus Community, which indicates a shallowing after the Middle Llandovery. Cb has yielded a fauna which includes Pentamerus, Stricklandia, and Clorinda, suggesting a return to deeper water. It is not possible to determine the community present in Cc, but increase in depth is again shown by the graptolites present in Cd. Correlation. The Upper Llandovery of Garth can be tentatively correlated with the beds of Llandovery on gross lithological and faunal grounds. Andrew and Jones (1925, p. 412) suggest it may be equivalent to C4-6. The presence of beds with Monograptus priodon and Monoclimacis crenulata conformably below the basal Wenlock shows that the highest Llandovery is present in this area. CHRONOLOGICAL SUMMARY Text-fig. 1 1 illustrates the chronological sequence of Llandovery deposits in the Welsh Borderland. It brings out the fact that, although the spread of the sea was continuous from Lower Llandovery times, there were several gaps in sedimentation. In general, however, there is evidence of a progressive deepening at any one locality. Lower Llandovery (T4-4) times The shelf was narrow during the Lower Llandovery, and deposits are confined to an area west of Welshpool and Garth (text-fig. 12). In the basin area of Central Wales (text-fig. 11, cols. 1, 10, 21, and 22), basal Llandovery graptolitic shales follow above similar Ashgill sediments without a break. At Trannon and Llanidloes this facies con- tinues throughout the Lower Llandovery, but at Garth Stricklandia and Clorinda Com- munities are present, showing that shallowing took place there in A2_3 times. This was probably followed by a depositional break, as no A4 has been proved at Garth. On the edge of the basin at Meifod and Welshpool (text-fig. 11, cols. 2 and 3), there was uplift of Ashgill sediments prior to the deposition of beds containing A4_2 shallow- water Cryptothyrella Communities. In Shropshire, to the east (text-fig. 11, cols. 5, 6, 1 1-14), the Caradoc and earlier rocks are seen to be considerably folded before the Middle Llandovery. If these movements are correlated with the basal A4 non-sequence at Meifod and Welshpool, then it may be that the acme of this folding occurred at the very beginning of the Silurian. After the initial uplift, the shallow-water A4_2 beds at Meifod and Welshpool are followed by deeper-water A3 Stricklandia and Clorinda Com- munities. The general subsidence characteristic of the Welsh Borderland throughout the Lower Silurian thus started as early as A2_3 times. Middle Llandovery ( fi,-3) times In the basin (text-fig. 1 1 , cols. 1 and 10), the graptolitic facies continues without break; this facies spread over the basin margin at Meifod (text-fig. 11, col. 2) in B2time, when the Clorinda Community was replaced by graptolites. At about the same time, the sea C 6055 3 E 772 PALAEONTOLOGY, VOLUME 11 Early Wenlock C6 v. C5 £.c4 §C3 N °2 S' C, ^ — > > ' V V — 7 — 7 \ ^ \ \ \ \ \ G \ \ \ G G ^ V "V \\ (Buildwas Bed ////////, / / /;? ; ? ;?; X \ \ \ \ \ G \ \ \ G / // / -/ / / / / / */ / / / / ' ' ' y / sypp ; ///////' ' -- 7 • ssss\\sxs f \ \ \ \ \ N G \ \ \ G / / / / ////// C ilSl f \ \ \ \ \ ' G \ \ \ G \ ( / / / / / / / / / \ \ \ \ \N \ \ \ \ \ \ n \ \ j ' , / / / / / / / / P ’/////A' / / / /XA\ o S \ \ \ \ ' u \ \ \ o \ ^ (> p / / / / L/ n ■////.'.'////'A' Q = Graptolitic Facies = Clorinda Community = Strick/andia or Costistricklandia 11 = Pentamerus or __ Pentameroides " = Eocoe/ia or Cryptothyrella " = Lingula " \ \ \ \ \ G \ \ \ \ \| ^ (/ L- / / / / / / AUxy//. r \ \ \ \ \ \ \ \y \ / J/l ■ NUN L pi .,A _ S \ \ \ \ \ G \ \ \ // / i s S \ \ \ \ \ ' \ / C / / ■ C, : p \ \ \ \ \ [ / / > / 'J \ \ \ \ \ ' \ \ \ \ \ ' Xp c, S 7 F AT N VO i / A £ E \ \ \ \ \ ' 1 V SE 1 — o ON 1 L \ \ \ \ \ ' ? text-fig. 1 1. Three sections across the Welsh Borderland (see text-fig. 1 and 3) showing communities plotted against time. ZIEGLER, COCKS, AND McKERROW: LLANDOVERY TRANSGRESSION 773 deepened at Welshpool (text-fig. 11, col. 3) and spread over West Shropshire (text-fig. 11, cols. 4, 5, 11, and 13) and may have just reached the Malverns (text-fig. 11, cols. 15-17) by the end of Middle Llandovery times. It is worth noting here that there are no Middle Llandovery sediments in nearby sections at Eaton Farm and Wistanstow text-fig. 12. Palaeogeography of the area during Lower Llandovery (A2) time. E, Cryptothyrella Community; S, Stricklandia Community; C, Clorinda Community; G, graptolitic facies. Note. The coastline is shown as a smooth curve, but we envisage a coastal topography perhaps similar to Essex at the present time (this applies also to text-fig. 13 and 14). The size of some of the smaller outcrops is enlarged for greater clarity, and the Upper Severn Estuary is omitted. (text-fig. 11, cols. 12 and 14). The earliest deposits at these localities contain C2_3 Clorinda Communities; this shows that the sea was deep before sedimentation com- menced, and that the advance of the sea was not everywhere accompanied by a supply of sediments. Thus the sea probably extended over the Eaton Farm and Wistanstow areas in Middle Llandovery time. 774 PALAEONTOLOGY, VOLUME 11 Early Upper Llandovery (Cx-3) times At Trannon, Meifod, and Llanidloes (text-fig. 11, cols. 1, 2, and 10) the graptolite facies continues, both in the basin and on the edge of the shelf (text-fig. 13). In West Shropshire (text-fig. 11, cols. 4 and 5) the sea is seen to have deepened with the replace- text-fig. 13. Palaeogeography of the area during early Upper Llandovery (Q) time. L, Lingula Community; E, Eocoelia Community; P, Pentamerus Community; S, Stricklandia Community; C, Clorinda Community; G, graptolitic facies. ment of the Stricklandia Community by the Clorinda Community. South of the Long- mynd (text-fig. 11, col. 13) the communities show a minor reversal, but there was an over-all deepening from C1 to C3 times. At neighbouring sites (text-fig. 11, cols. 12 and 14) sedimentation commenced for the first time with the deposition of deep-water Clorinda Communities resting directly on Pre-Cambrian and Ordovician rocks. In the Wenlock Edge outcrop, and in the Malverns and at Presteigne (text-fig. 11, cols. 6, 15-17, and 25), there are undoubted early Upper Llandovery sediments; these contain shallow-water communities showing the progressive deepening so characteristic ZIEGLER, COCKS, AND McKERROW: LLANDOVERY TRANSGRESSION 775 of this transgression. There is a sequence from a basal Lingula Community (not exposed at Presteigne) deepening to Eocoelia or Pentamerus Communities. At May Hill (text-fig. 1 1, col. 27) there is a minor reversal in the progressive deepening as Lingula and Eocoelia Communities alternate. text-fig. 14. Palaeogeography of the area during late Upper Llandovery (C5) time. E, Eocoelia Community; P , Pentaineroides Community; S, Costistricklandia Community; C, Clorinda Community; G, graptolitic facies. At Garth and Builth (text-fig. 11, cols. 21-3) no beds of C4-3 age are present, this is probably to be accounted for by non-deposition in a sea of Pentamerus Community depth, as this community is present above the non-sequence in beds of C4 age. Late Upper Llandovery (C4-6) times During this time, basin facies (the Pale Shales) continue at Garth (text-fig. 11, col. 21), but, in Wenlock times, graptolitic deposits are present at Builth as well (col. 23); an eastwards spread of this deep environment thus presumably occurred in C6 times (text- fig. 14). A similar C6 spread probably took place over the edge of the basin from 776 PALAEONTOLOGY, VOLUME 11 Montgomeryshire into West Shropshire (text-fig. 11, cols. 2-4); this deepening of the sea may also be related to the turbidites at Minsterley (which are of late Upper Llandovery [post C4] or early Wenlock [pr e-riccartonensis Zone] age). On Wenlock Edge (text-fig. 11, col. 6), the progressive deepening, mentioned above, continues through most of C4_6, but in late C6 time there is an indication of slight shallowing in the beds immediately below the Buildwas Beds. From sandstone pebbles in the Upper Carboniferous north-east of Coventry (text-fig. 11, col. 9), Eocoelia has been found, ranging in age from C4 to C6. However, at Walsall and Rubery (text-fig. 11, cols. 7 and 8) the earliest Silurian deposits contain C5 beds with Costistricklandia Communities and rest directly on Cambrian rocks ; here, as in Shrop- shire during C4-3 (see above), we have strong indications that the sea was present at Walsall and Rubery well before sedimentation commenced. The evidence from Lower Lemington (text-fig. 11, col. 20) suggests a shallow sea there in C6 times, but the local base of the Silurian is unseen. In the Malvern areas (text-fig. 11, cols. 15-19), no beds of C3_4 age are known; this non-sequence is followed by C5 beds with Pentameroides and Costistricklandia Com- munities which overlap the C4-2 deposits (which contain Lingula and Eocoelia Com- munities), and rest in places on the Pre-Cambrian. Text-fig. 11 shows that there is a progressive deepening of the sea across this non-sequence; thus there is no suggestion of any marine regression during C3_4 time in the Malvern area. A different type of stratigraphical break is present beneath the Nash Scar Limestone at Presteigne and Old Radnor (text-fig. 11, cols. 24 and 25); a shallow-water Wenlock algal limestone rests unconformably on C2 beds (with a Pentamerus Community) at Presteigne, and on Pre-Cambrian at Old Radnor. There must have been shallowing during this interval, and local uplift may well have occurred here during the late Upper Llandovery. The progressive deepening, seen or inferred for C4_6 beds of most areas, was also present to the south of the Malverns in the May Hill and Tortworth areas (text-fig. 11, cols. 27-9). At May Hill the communities range from Eocoelia through Pentameroides and Costistricklandia to Clorinda, and at Tortworth the Stone succession (col. 28) also shows a progressive sequence. This last sequence contains a 15-ft. (5-m.) thick basalt within the beds containing the C6 Costistricklandia Community. But at the other end of the Tortworth Inlier (at Charfield Green, text-fig. 11, col. 29) the basalt has thickened to 150 ft. (50 m.), and is followed directly by beds containing an Eocoelia Community. We conclude that the sudden shallowing of the shelf sea by 150 ft. is directly responsible for the replacement of a Costistricklandia Community by an Eocoelia Community; this implies that the depth range of the intervening Pentameroides Community is signi- ficantly less than 150 ft. CONCLUSIONS The elucidation of the evolution of several brachiopod genera has enabled wide cor- relation of the 13 subdivisions of the Llandovery established in the type area by O. T. Jones. Recognition of similar assemblages of fossils at different times has led to the establishment of five main animal communities. In the Silurian the marine communities of the shelf were largely epifaunal and relatively independant of substrate; depth was thus the main controlling factor in community distribution. ZIEGLER, COCKS, AND McKERROW: LLANDOVERY TRANSGRESSION 111 In the shelf regions on the Welsh Borderland, the Llandovery always rests uncon- formably on older rocks; on the margin of the basin there is only a small break at the very base of the Lower Llandovery ; and in the basin of Central Wales there is conformity between the Ordovician and Silurian. From Ax to C6 there was a progressive spread of the sea to the east and south. In the Lower Llandovery marine deposits are present at Welshpool and Garth; by Middle Llandovery times West Shropshire had been sub- merged ; and by early Upper Llandovery (Q) times the sea had transgressed as far as the Malverns and May Hill; it did not reach Tortworth until the late Upper Llandovery (C5). The theory (Jones 1938) that there was a sudden submergence of the whole area at the beginning of the Upper Llandovery cannot now be accepted. With few exceptions, each local sequence shows a progressive increase of depth with time. This progressive depth increase can even span periods of non-deposition, for example the C5 beds above the non-sequence in the Malverns were laid down in deeper water than in C2 beds below the break. We have concluded that, where a deeper-water community is present at the base of a local sequence, the sea was present for some time before deposition commenced. It follows that the spread of the sea with time was even more uniform than the sedimentary sequences record. It also follows that there were periods when no sediments were being deposited in certain parts of the sea floor; this may have been due either to lack of sediment supply or to no deposition or erosion in current-swept areas. The Welsh Borderland is on the south-eastern margin of a Lower Palaeozoic geo- syncline which included the Lake District, the Southern Uplands of Scotland and much of central Ireland. Both this margin and the western margin in County Galway show an extension of the sea during Llandovery times, but at the same time north-east New- foundland (and perhaps also parts of the Midland Valley of Scotland, for example, Lesmahagow), show a transition from deep-water basin sediments to shelf or even non- marine conditions. When the Welsh Borderland is viewed as part of this larger region, it seems improbable that the advance of the sea during Llandovery times was due to eustatic deepening. Although there was probably little structural control of sedimentation from place to place within the Welsh Borderland at this time, there were three major tectonic regions affecting the area: (i) The basin region to the west which subsided rapidly, accommodating large thick- nesses of sediment (nearly 10 000 ft. (3300 m.) of Llandovery beds at Llanidloes). (ii) The shelf region, which subsided slowly throughout the whole of Llandovery time, and upon which thinner sediments accumulated. (iii) A region of uplands to the south and east, which provided the sediment for both of the other regions. REFERENCES Andrew, g. 1925. The Llandovery rocks of Garth, Breconshire. Q. Jl geoL Soc. Loud. 81, 389-406, pi. 22. — — and jones, o. t. 1925. The relations between the Llandovery rocks of Llandovery and those of Garth. Ibid. 407-16. bassett, d. a. 1963. The Welsh Palaeozoic geosyncline: a review of recent work on stratigraphy and sedimentation. In The British Caledonides, ed. M. R. W. Johnson and F. H. Stewart, 35-69. Oliver & Boyd, Edinburgh and London. 778 PALAEONTOLOGY, VOLUME 11 butler, a. j. 1937. On Silurian and Cambrian rocks encountered in a Deep Boring at Walsall, South Staffordshire. Geol. Mag. 74, 241-57. cave, r. 1965. The Nod Glas sediments of Caradoc age in North Wales. Geol. Jnl, 4, 279-98, pi. 12. cocks, l. r. m. 1967m Llandovery stropheodontids from the Welsh Borderland. Palaeontology, 10, 245-65, pi. 37-9. 19676. Depth patterns in Silurian Marine Communities. Marine Geol. 5, 379-82. — and rickards, r. b. (in press). Five boreholes from Shropshire and the relationships of shelly and graptolitic facies in the Lower Silurian. Q. Jl geol. Soc. Land. and walton, g. 1968. A large temporary exposure in the Lower Silurian of Shropshire. Geol. Mag. 105, 390-7. curtis, m. l. k. 1955. Lower Palaeozoic. In Bristol and its adjoining counties, ed. C. M. Maclnnes and W. F. Whittard, 3-7. Brit. Assn Adv. Sci. Bristol. Eastwood, t., whitehead, t. h., and robertson, t. 1925. The geology of the country around Birming- ham. Sheet 168. Mem. Geol. Surv. ( U.K. ). 1-152, pi. 1-7. elles, G. l. and wood, e. m. r. 1901-18. A monograph of British graptolites. Palaeontogr. Soc. [Monogr.], 1-539, pi. 1-52. gardiner, c. i. 1920. The Silurian rocks of May Hill. Proc. Cotteswold Nat. Fid Club, 20, 185-218, pi. 1. garwood, e. J. and Goodyear, E. 1918. On the geology of the Old Radnor District, with special refer- ence to an algal development in the Woolhope Limestone. Q. Jl geol. Soc. Lond. 74, 1-30, pi. 1-7. green, G. w. 1962. In Summ. Progr. geol. Surv. (U.K.) (for 1961), p. 29. greig, d. c., wright, J. e., and mitchell, G. h. 1968. Geology of the country around Church Stretton, Craven Arms, Wenlock Edge and Brown Clee. Sheet 166. Mem. geol. Surv. (U.K ). 1-379, pi. 1-13. groom, t. t. 1899. The geological structure of the Southern Malverns, and of the adjacent district to the west. Q. Jl geol. Soc. Lond. 55, 129-69, pi. 13-15. 1900. On the geological structure of portions of the Malvern and Abberley Hills. Ibid. 56, 138-97, pi. 8. — — 1910. The geology of the Malvern and Abberley Hills, and the Ledbury District. Geol. Assoc., Jubilee Vol. 698-738, pi. 23. holl, H. b. 1865. On the geological structure of the Malvern Hills and adjacent districts. Q. Jl geol. Soc. Lond. 21, 72-102. jones, o. t. 1925. Geology of the Llandovery District. Part 1 : the southern area. Ibid. 81, 344-88, pi. 21. 1938. On the evolution of a geosyncline. Ibid. 94, lx-cx. 1947. The geology of the Silurian rocks west and south of the Carneddau Range, Radnorshire. Ibid. 103, 1-36, pi. 1. 1949. Geology of the Llandovery District. Part 2: the northern area. Ibid. 105, 43-64, pi. 3. jones, w. d. v. 1945. The Valentian succession around Llanidloes, Montgomeryshire. Ibid. 100 (for 1944) 309-32, pi. 14. jukes, J. b. 1853. On the occurrence of Caradoc sandstone at Great Barr, South Staffordshire. Ibid. 9, 179-81. king, w. b. r. 1928. The geology of the district around Meifod, Montgomeryshire. Ibid. 84, 671-702, pi. 52. kirk, N. h. 1951. The Upper Llandovery and Lower Wenlock rocks of the area between Dolyhir and Presteigne, Radnorshire. Proc. geol. Soc. Lond. 1471, 56-8. lamont, a. and gilbert, d. f. l. 1945. Upper Llandovery Brachiopoda from Coneygore Coppice and Old Storridge Common, near Alfrick, Worcs. Ann. Mag. nat. Hist. [11] 12, 641-82, pi. 3-7. lapworth, c. 1878. The Moffat series. Q. Jl geol. Soc. Lond. 34, 240-346, pi. 1 1-13. lawson, j. d. 1955. The geology of the May Hill Inlier. Ibid. Ill, 85-116, pi. 6. Phillips, j. 1848. The Malvern Hills, compared with the Palaeozoic Districts of Abberley, Woolhope, May Hill, Tortworth, and Usk. Mem. geol. Serv. (U.K.) 2 (1), 1-330, pi. 1-3. phipps, c. b. and reeve, f. a. e. 1967. Stratigraphy and geological history of the Malvern, Abberley and Ledbury Hills. Geol. Jnl, 5, 339-68. pocock, r. w. 1930. The Petalocrinus Limestone horizon at Woolhope (Herefordshire). O. Jl geol. Soc. Lond. 86, 50-63, pi. 6, 7. whitehead, t. h., wedd, c. b., and robertson, T. 1938. Shrewsbury District, including the Han- wood Coalfield. Sheet 152. Mem. geol. Surv. (U.K.). 1-297, pi. 1-8. ZIEGLER, COCKS, AND McKERROW: LLANDOVERY TRANSGRESSION 779 reading, h. G. and poole, a. b. 1961. A Llandovery shoreline from the Southern Malverns. Geol. Mag. 93, 295-300, pi. 15, 16. 1962. Malvern structures. Ibid. 99, 377-9. reed, f. r. c. and Reynolds, s. h. 1908. On the fossiliferous Silurian rocks of the southern half of the Tortworth Inlier. Q. Jl geol. Soc. Load. 64, 512-45. Reynolds, s. h. 1907. A Silurian inlier in the Eastern Mendips. Ibid. 63, 217-40, pi. 18. 1912. Further work on the Silurian rocks of the Eastern Mendips. Proc. Bristol Nat. Soc. 3, 76-82. 1924. The igneous rocks of the Tortworth Inlier. Q. Jl geol. Soc. Load. 80, 106-12, pi. 7, 8. 1929. The geology of the Bristol District. Proc. Geol. Assoc. 40, 77-103. robertson, t. 1926. The section of the new railway tunnel through the Malvern Hills at Colwall. Summ. Progr. geol. Surv. ( U.K .) (for 1925), 162-75. st. Joseph, j.k. s. 1935. A critical examination of Stricklandia(= Stricklandinia ) lirata (J. deC. Sowerby) 1839 forma typica. Geol. Mag. 72, 401-24, pi. 16, 17. 1938. The Pentameracea of the Oslo region. Norsk geol. Tidsskr. 17, 225-336, pi. 1-8. shotton, F. w. 1927. The conglomerates of the Enville Series of the Warwickshire Coalfield. Q. Jl geol. Soc. Load. 83, 604-21, pi. 47, 48. sollas, w. j. 1879. On the Silurian district of Rhymney and Pen-y-lan, Cardiff. Ibid. 35, 475-507, pi. 24. squirrell, h. c. and tucker, e. v. 1960. The geology of the Woolhope Inlier (Herefordshire). Ibid. 116, 139-85, pi. 15. strahan, A. 1913. Batsford (or Lower Lemington) Boring, near Moreton-in-Marsh. Summ. Progr. geol. Surv. (U.K.) (for 1912), 90-1. symonds, w. s. 1872. Records of the rocks. John Murray, London, 433 pp. toghill, p. 1968. The graptolite associations and zones of the Birkhill Shales (Llandovery) at Dobb’s Linn. Palaeontology , 11, 654-68. tyler, w. h. 1925. Notes on Sheet 48 N.W. (Shropshire). Proc. Geol. Assoc. 36, 377. wade, a. 1911. The Llandovery and associated rocks of north-eastern Montgomeryshire. Q. Jl geol. Soc. Loud. 67, 415-59, pi. 33-6. watts, w. w. 1885. On the Igneous and Associated Rocks of the Breidden Hills in East Montgomery- shire and West Shropshire. Ibid. 41, 532-46. whittard, w. f. 1928. The stratigraphy of the Valentian rocks of Shropshire: the Main Outcrop. Ibid. 83 (for 1927), 737-59, pi. 56, 57. 1932. The stratigraphy of the Valentian rocks of Shropshire. The Longinynd-Shelve and Breidden Outcrops. Ibid. 88, 859-902, pi. 58-62. Whittington, h. B. 1938. The geology of the district around Llansantffraid ym Mechain, Montgomery- shire. Ibid. 94, 423-57, pi. 38, 39. williams, a. 1951. Llandovery brachiopods from Wales with special reference to the Llandovery District. Ibid. 107, 85-136, pi. 3-8. wills, l. j. and laurie, w. h. 1938. Deep sewer trench along the Bristol road from Ash ill Road near the Longbridge Hotel to the City Boundary at Rubery, 1937. Proc. Birmingham nat. Hist. phil. Soc. 16, 175-80, pi. 1. wilkens, l. G. and hubbard, G. H. 1925. The Upper Llandovery Series of Rubery. Ibid. 15, 67-83. wood, e. m. r. 1906. The Tarannon Series of Tarannon. Q. Jl geol. Soc. Lond. 62, 644-701, pi. 47, 48. ziegler, a. m. 1964. The Malvern Line. Geol. Mag. 101, 467-9. 1965. Silurian marine communities and their environmental significance. Nature, Lond. 207, 270-2. 1966m Unusual stricklandiid brachiopods from the Upper Llandovery beds near Presteigne, Radnorshire. Palaeontology, 9, 346-50, pi. 58. 1 9666. The Silurian brachiopod, Eocoelia hemisphaerica (J. de C. Sowerby) and related species. Ibid. 523-43, pi. 83, 84. cocks, l. r. m. and bambach, r. k. 1968. The composition and structure of Lower Silurian marine communities. Lethaia, Oslo, 1, 1-27. and boucot, a. j. 1968. North American Silurian animal communities. In W. B. N. Berry, and A. J. Boucot. Silurian of North America. Mem. geol. Soc. Am. (in press). 780 PALAEONTOLOGY, VOLUME 11 APPENDIX 1 Localities of collections in Oxford University Museum Area Locality Tortworth 1 2 3 May Hill 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 Malvern Hills 23 24 25 26 27 28 29 30 31 32 33 Old Storridge Common 34 35 36 37 38 39 40 41 Ankerdine Hill 42 43 44 45 Field designation Grid reference T-M-A (ST/7268 9212) T-R-A (ST/7233 9234) T-D-A (ST/6962 9390) M-N-A (SO/7014 2104) M-D-A (SO/7055 1987) M L-A (SO/7052 1838) M-L-B (SO/7049 1838) M-L-C (SO/7048 1839) M-Q-A (SO/7066 1994) M-Q-B (SO/7067 1993) M-H-A (SO/7047 1811) M-B-A (SO/7041 1974) M-B-B (SO/7039 1973) M-G-C (SO/7051 2097) M-F-D (SO/6874 2154) M-F-A (SO/6874 2154) M S-A (SO/6918 2261) M-S-B (SO/6932 2270) M-S-C (SO/6936 2271) M-F-C (SO/6872 2155) M-O-A (SO/6869 2244) M-G-B (SO/7055 2103) H-O-A (SO/7533 3764) H-H-A (SO/7487 3595) H-T-A (SO/7494 3492) H-G-A (SO/7612 3811) H-G-D (SO/7612 3811) H-G-E (SO/7612 3811) H-G-F (SO/7608 3811) H-P-A (SO/7646 4594) H-C-A (SO/7672 4442) H-R-A (SO/7613 4784) H-S-A (SO/7616 4723) H-L-A (SO/7467 5115) H-F-C (SO/7444 5124) H-M-A (SO/7430 5152) H-M-B (SO/7430 5152) H-L-C (SO/7464 5108) H-L-B (SO/7464 5111) H-F-B (SO/7438 5114) H-M-C (SO/7405 5167) A-A-A (SO/7352 5625) A-H-A (SO/7376 5696) A-Q-A (SO/7363 5630) A-K-A (SO/7347 5690) Collections from loose blocks. ZIEGLER, COCKS, AND McKERROW: LLANDOVERY TRANSGRESSION 781 Area Locality Field designation Grid reference Wenlock Edge 46 Morrellswood (SJ/6284 0637) 47 Boathouse A (SJ/6210 0379) 48 Boathouse B (SJ/6206 0390) 49 Boathouse C (SJ/6205 0398) 51 Sheinton B (SJ/6116 0310) 52 Merrishaw A1 (SJ/5805 0090) 53 Merrishaw A2 (SJ/5805 0090) 54 Merrishaw B (SJ/5852 0065) 55 Domas (SJ/5936 0062) 56 Hughley (SO/5605 9747) 58 W all-u-Hey wood (SO/5120 9276) 59 Ticklerton (SO/4821 9088) 60 Onny River (SO/4260 8532) Church Stretton 61 Marshbrook (SO/4341 8982) 62 Hillend Farm (SO/3956 8769) Norbury 63 Norbury (SO/3587 9284) Bog 65 Bog A (SO/3510 9815)* 66 Bog B (SO/3510 9815)* Minsterley 67 Ox Wood Dingle (SJ/2909 0123) 68 Venusbank (SJ/3534 0125) 69 Hope Outlier (SJ/3482 0140) 70 Hope Quarry (SJ/3551 0208) 71 Hope Brook (SJ/3578 0212) 72 Josey’s Wood A (SJ/3656 0213) 73 Josey’s Wood B (SJ/3653 0221) APPENDIX 2 Localities of collections in British Museum ( Natural History ) Area Locality Field designation Grid reference Wenlock Edge 50 Sheinton A (SJ/6099 0324) 57 Gilberries (SO/5115 9361) Norbury 64 Linley Wall (SO/3587 9284)* Minsterley 74 Josey’s Wood C (SJ/3641 0216) 75 Minsterley Lane (SJ/3803 0487) APPENDIX 3 Localities of collections in United States National Museum Area Locality U.S.N.M. no. Grid reference Tortworth 76 10215 (ST/6879 9574) 77 10217 (ST/7056 9428) 78 10218 (ST/7267 9206) 79 10219 (ST/7267 9206) 80 10220 (ST/7055 9426) 81 10221 (ST/7055 9426) 82 10223 (ST/7045 9440) * Collections from loose blocks. 782 PALAEONTOLOGY, VOLUME 11 Area Locality U.S.N.M. no. Grid reference May Hill 83 10225 (SO/6927 2263) 84 10226 (SO/6927 2263) Woolhope 85 10230 (SO/5804 3730) 86 10232 (SO/5948 3564) Old Storridge Common 87 10237 (SO/7496 5156) Presteigne 88 10242 (SO/3177 6343) 89 10243 (SO/3197 6334) 90 10244 (SO/3160 6320) 91 10245 (SO/3153 6316) 92 10246 (SO/3155 6317) 93 10247 (SO/3018 6235) Eastern Mendips 94 12936 (ST/6755 4585)* * Collections from loose blocks. A. M. ZIEGLER Department of the Geophysical Sciences University of Chicago Illinois 60637, U.S.A. L. R. M. COCKS Department of Palaeontology British Museum (Natural History) London S.W. 7 W. S. MCKERROW Department of Geology and Mineralogy Parks Road Oxford Typescript received 1 March 1968 THE UNUSUAL BRACHIAL SKELETON OF ATTENUATELLA CONVEXA S P. NOV. (BRACHIOPODA) by JOHN ARMSTRONG Abstract. Excellently preserved specimens of Attenuatella convexa sp. nov. have enabled reconstruction of a unique brachial skeleton for Attenuatella. Apart from this characteristic the genus is closest to members of the Ambocoeliinae (Spiriferida) and its inclusion in this subfamily is still the most plausible supra-generic grouping. Globally, eleven species of Attenuatella are now documented and the known range of the genus is from the lower Artinskian (Aktastinian) to the Kazanian or Wordian Stage. The species Attenuatella convexa sp. nov., Attenua- tella sp. A, and Attenuatella sp. cf. incurvata Waterhouse are here described from Queensland. Attenuatella Stehli (1954) is a small, very distinctive spiriferid which occurs at only a few localities throughout the world. The genus has recently been documented from Australia by Waterhouse (1967) and by Armstrong and Brown (1968) who describe species from New South Wales and Queensland respectively. Several additional occurrences are known from the Permian faunas of Queensland and the species from these are described here. Serial acetate peels of well-preserved specimens of one of these species, Attenuatella convexa sp. nov., have supplemented the hitherto only fragmentary information that is available about the brachial skeleton of the genus. Specimens of Attenuatella from a number of the localities in Queensland were kindly lent by Dr. J. M. Dickins of the Bureau of Mineral Resources, by Dr. J. F. Dear of the Geological Survey of Queensland, and by Dr. B. N. Runnegar of the University of Queensland. Dr. Dickins made available specimens which he collected from three localities in the Bowen Basin, and Dr. Dear and Dr. Runnegar each supplied a specimen which they had collected from the Barfield Formation in the Bowen Basin, and from just below the Upper Limestone near Gympie respectively. Bureau of Mineral Resources localities referred to, occur on the Duaringa and Taroom 1 : 250,000 military maps and are designated by a number prefixed by Du and T respectively. Geological Survey of Queensland, and University of Queens- land Department of Geology and Mineralogy locality numbers are prefixed by GSQL and UQL respectively. All specimens mentioned have individual numbers which follow the initials of the institu- tion in which the specimens are housed; CPC, Commonwealth palaeontological type collection. Bureau of Mineral Resources, Canberra; GSQF, Geological Survey of Queensland; UQF, Department of Geology and Mineralogy, University of Queensland. Order spiriferida Waagen 1883 Subfamily ambocoeliinae George 1931 Remarks. The brachidia of Attenuatella are unusual and are very distinctive. Serial acetate peels prepared from ten specimens of Attenuatella convexa sp. nov. do not show any traces of spiralia but they reveal the structures shown in text-fig. 1. Waterhouse (1964) has also figured serial sections of specimens of a species of Attenuatella. He found that each brachidium in specimens of A. incurvata Waterhouse consisted of an anteriorly directed lamella which at the front of the shell gives rise to a semicircle-like structure parallel to the hinge of the shell (Waterhouse 1964, fig. 48). The brachidia of the [Palaeontology, Vol. 11, Part 5, 1968, pp. 783-92, pi. 142.] 784 PALAEONTOLOGY, VOLUME 11 sectioned specimens of A. convexa, although initially similar to the preserved brachidia of A. incurvata, are more complete and they have enabled reconstruction of brachidia presumably characteristic of the genus (text-fig. 2) though those of the type species are unknown. Each brachidium of A. convexa consists of an anteriorly directed lamella text-fig. 1. Serial acetate peels of four (A, B, C, D) specimens of Attenuatella convexa sp. nov. The interval between successive sections is 025 mm. except where indicated, and in the latter cases the dimensions given on the diagram are in millimetres. which leads into an S-shaped structure (text-fig. 1) from the ventral end of which there is a ventrally placed posteriorly directed lamella. The consistency of the arrangement of these features in the sectioned specimens of A. convexa suggests that they represent the complete brachial skeleton of the species. Spiral lamellae are not present in A. con- vexa but rather the brachial skeleton is basically loop-like with curious anterior modi- fications. The nature of the brachidia of Attenuatella does not necessitate its reclassification. Indeed except for the fact that the spiralia characteristic of several members of the Ambocoeliinae (George 1931, Veevers 1959, Vandercammen 1956) are absent from J. ARMSTRONG: BRACHIAL SKELETON OF ATTENUATELLA 785 Attemiatella, the characteristics of Attenuatella completely support its classification in this subfamily. Attenuatella has an impunctate shell, a shell form which is comparable with other ambocoelins, a spinose micro-ornament, areas on both valves, and a non- striated tuberculate cardinal process. The placement of Attenuatella in the Ambocoe- liinae is therefore maintained. A spiriferid having abbreviated brachidia resembling those of Attenuatella is the Middle Ordovician genus Protozyga. Protozyga possesses a jugum but otherwise each of its brachidia has the same basic form of a single volution parallel to the median plane of the shell as the brachidia of Attenuatella. text-fig. 2. Diagrammatic reconstruction of an antero-ventral view of the brachidia of Attenuatella convexa sp. nov., based on serial acetate peels of ten specimens, four of which are shown in text-fig. 1 . The reason for the brachidial deviation in Attenuatella from the normal type of spiriferid spiral is problematic and it is conjectural whether the lophophore of Attenua- tella had the form of a spiral. The shape of the brachidia of A. convexa does not pre- clude the presence of a spiral-like lophophore, and if such an organ did exist it appears that in general only the first flexured whorl of the lophophore was provided with a cal- careous support. If the lophophore was not spirolophous but rather was confined to the reduced brachidia it is possible that the S-shaped brachidial modifications were developed to compensate the loss of length of the lophophore. In this instance the feeding organ was probably zygolophous. The brachial skeleton of Attenuatella seems to be of a more primitive type than the spiralia of other ambocoelins. Perhaps it is relevant that species of Attenuatella are the youngest known representatives of the Ambocoeliinae and that reversion to this loop-like arrangement in Attenuatella occurred not long before the extinction of the subfamily in the Upper Permian. 786 PALAEONTOLOGY, VOLUME 11 Genus attenuatella Stehli 1954 Attenuatella Stehli; Chernyak 1963 Attenuatella Stehli; Waterhouse 1964 Type species (original designation). Attenuatella texana Stehli 1954, pi. 25, figs. 31-3 from the lower part of the Lower Permian Bone Spring Formation, Texas, text-fig. 3 (1). Diagnosis. Small essentially plano-convex spiriferids; ventral valve elongate, smooth, or sulcate, with elongate relatively narrow umbo and more or less incurved beak; ventral area apsacline; dorsal valve convex, flat, or gently concave with low area; ornament of small spines often arranged in concentric lines, and fine discontinuous radial grooves; each groove runs anteriorly for a short distance from a spine; field of muscular attach- ment in ventral valve an elevated platform; no plates support the teeth; crural plates sessile and cardinal process tuberculate; shell impunctate. Other species and specimens A. acutirostrus (Krotova) 1885, pi. 4, fig. 24 of Artinskian age from the Ural Mountains in Russia, text-fig. 3 (2). A. stringocephaloides (Chernysheva and Likharev) in Likharev and Einor 1939, pi. 13, fig. 5a, b from Abrosimov Bay, Novaya Zemlya, text-fig. 3 (3). According to Likharev and Einor (1939, p. 189) a number of the species occurring with A. stringocephaloides at Abrosimov Bay (Likharev and Einor 1939, p. 185) also occur in the Spirifer Limestone and Brachiopod Chert in Spitzbergen. Likharev and Einor further intimate (p. 188) that all of the faunas of Novaya Zemlya are to be equated with the faunas of the Spirifer Limestone and the Brachiopod Chert for which Stepanov (1957) has suggested either an upper Lower Permian or a lower Upper Permian age or perhaps both (i.e. belonging to his Svalbardian Stage). A. attenuata (Cloud) 1944, pi. 17, figs. 22-5 from the Wordian Waagenoceras Zone in Mexico, text-fig. 3 (4). A. stringocephaloides (Chernysheva and Likharev); Chernyak 1963, pi. 42, figs. 3 and 4 from the lower Baykura sub-horizon of eastern Taimyr, text-fig. 3 (5). Ustritskiy and Chernyak (1963) correlate the lower Baykura sub-horizon with the lower Vorkuta Series in the Pechora Basin and they consider that both of these units correspond to the Ufa Suite on the Russian Platform. Stepanov (1957) intimated that the latter Suite should be included in the Svalbardian Stage which he proposed (1957) for faunas in Spitzbergen intermediate in aspect between the faunas of the Artinskian Stage and the Kazanian Stage. For this time interval Ustritskiy (1960) also proposed a new stage, the Paykhoyian Stage based on the fauna of the Vorkuta Series, and in 1963 Ustritskiy and Chernyak had little hesitation in referring the fauna of the lower Baykura sub-horizon to the upper part of Ustritskiy’s stage. Ustritskiy and Chernyak also suggest, in view of the absence of a complete Permian section in Spitzbergen, that it might be preferable to recognize the Paykhoyian Stage rather than the Svalbardian Stage. A. taimyrica Chernyak 1963, pi. 42, figs. 5-9 from the Byrranga horizon of eastern Taimyr, text-fig. 3 (6). Ustritskiy and Chernyak (1963) correlate the Byrranga horizon with a composite Artinskian- Kungurian Stage in the Ural Mountains. A. incurvata Waterhouse 1964, pi. 20, figs. 1-12, pi. 21, figs. 1-9 from the Arthurton Group in New Zealand, text-fig. 3 (7). Waterhouse (1964, 1967) considers the fauna in this unit to be of Kazanian age. A. sp. Landis and Waterhouse, 1966, pi. 1, figs. 1-5 from the Wesney Siltstone in the Eglinton Volcanics, New Zealand, text-fig. 3 (8). Landis and Waterhouse seem to be uncertain about the age of these specimens. They firstly state (p. 139) that the specimens 'are probably identical with Aktastinian (lower Artinskian) Attenuatella in the Takitimu Group', but later (p. 145) intimate a Kazanian age from their remark regarding the age of the Wesney Siltstone, that ‘a Kazanian age is possible on an objective consideration of the fossils alone’. A. sp. Landis and Waterhouse 1966, p. 144 also mention some specimens of Attenuatella from New Caledonia, and after a preliminary examination of the material they consider that the specimens could be conspecific with A. incurvata, text-fig. 3 (9). J. ARMSTRONG: BRACHIAL SKELETON OF ATTENUATELLA 787 A. multispinosa Waterhouse 1967, pi. 24, figs. 1-7 from the Gilgurry Mudstone in the Boorook Group in northern New South Wales, text-fig. 3 (10). Waterhouse (1967, p. 172) considers that this species is likely to be of late Kungurian or early Kazanian age. Australis Armstrong and Brown (1968) describe a new species from below the Gigoomgan Limestone in the Maryborough Basin, Queensland, text-fig. 3 (11). They suggest a lower Artinskian (Aktastinian) age for the species. A. convexa sp. nov. (see p. 788), text-fig. 3 (12). A. sp. A (see p. 790), text-fig. 3 (13). A. cf. incurvata (see p. 791), text-fig. 3 (14). text-fig. 3. Global distribution of Atteiuiatella. Each solid triangle signifies at least one occurrence of the species nominated by the adjacent number. References to the numbers are in the text. Comparison. Attenuatella is distinguished from Crurithyris, its closest relative by its narrower more elongate ventral valve and umbo, and the elevated platform of muscular attachment in the ventral valve. It is also possible to separate the genera on the nature of their brachidia. The brachial skeleton of A. texana is unknown. However, from the work of Waterhouse (1964) and from information gained from A. convexa it is known that each brachidium of these two species of Attenuatella consists essentially of an anteriorly directed lamella which eventually curves ventrally and then is deflected pos- teriorly to form a single volution parallel to the median plane of the shell. In contrast the brachial skeleton of Crurithyris urii (Fleming), the type species of Crurithyris, con- sists of laterally directed spires of from three to six whorls (George 1931, p. 40). Water- house (1964, p. 108) has further noted that on I. incurvata only one series of spines seems to be present whereas according to George (1931, pp. 51 and 55) there are two series of spines on C. urii. However, these differences may not be significant for on some speci- mens of A. convexa the spines are relatively uniform in size, and on others there are both large and small spines (PI. 142, figs. 5, 12). Another genus superficially similar to Attenuatella is Moumina Fredericks 1919 C 6055 3 F 788 PALAEONTOLOGY, VOLUME 11 (1924), but unfortunately this genus has been little used or described since its institution. Moreover, the only description of M. incertia (Chernysheva), the type species of Mou- mina (OD; Fredericks 1919) is Chernysheva’s original description of the general external appearance of the species. Distribution. In addition to its occurrence in Australia, Attenuatella is found in New Caledonia, Russia, North America, and New Zealand. The global distribution is shown in text-fig. 3. Range. Lower Artinskian (Aktastinian) to Kazanian or Wordian. Attenuatella convex a sp. nov. Plate 142, figs. 1-12, 19 Holotype. UQF53036 from UQL3127 in the lower part of the Tiverton Formation in the north-eastern part of the Bowen Basin, Queensland. (UQL3127, about half a mile east of Homevale Homestead, 20 miles north of Nebo on the Nebo-Collinsville road.) Diagnosis. Ventral valve moderately inflated with a distinct sulcus, and a broad mas- sive umbo that is not incurved over the area; commissure ligate; dorsal valve gently EXPLANATION OF PLATE 142 All figures are x 4 except where indicated Figs. 1-12. Attenuatella convexa sp. nov. All specimens are from the Tiverton Formation at UQL3127, about half a mile east of Homevale Homestead, 20 miles north of Nebo on the Nebo-Collinsville road, Queensland. 1, Internal mould of ventral valve, UQF53029. 2, 3, 4, Internal moulds of dorsal valves (note impressions of areas of adductor muscle attachment in figs. 3 and 4), UQF53030, UQF53031, and UQF53032 respectively. 5, 8, Latex casts from external moulds of ventral valves (note two sizes of micro-ornamental spines on specimen in fig. 5), UQF53033 and UQF53034 respectively. 6, Lateral view of latex cast of external mould of both valves, UQF53035. 7, 9, Latex casts of external moulds of dorsal valve and umbo of ventral valve, UQF53036 (holotype) and UQF53037 respectively. 10, Same specimen as in fig. 9, Natural size. 11, Micro-ornament on a ventral valve x 17 (note the relative uniformity in the size of the spines), UQF53038. 12, Internal mould of the umbonal part of a dorsal valve showing the impression of a tuberculate cardinal process X 12, UQF53039. Fig. 19. Attenuatella sp. cf. convexa. Postero-ventral view of internal mould of ventral valve. UQF47256, from the sandstone unit below the Upper Limestone near Gympie. The specimen is from the first ridge north of Rammut Road, three-quarters of a mile east of Chatsworth, 5 miles north of Gympie, south Queensland. Figs. 13-18. Attenuatella sp. cf. incurvata. Waterhouse. 13, 14, 15, Internal moulds of ventral valves. CPC901 8, CPC9019, and CPC9020 respectively from T1 11,2-7 miles north-north-west of the Cracow- Theodore road crossing of Delusion Creek, 10 miles north-west of Cracow, Queensland. 16, Latex cast from an external mould of a ventral valve. CPC9021 from Till. 17, Latex cast from an external mould of a dorsal valve and the umbo of the ventral valve. CPC9022 from Till. 18, Lateral aspect of a ventral valve. CPC9023 from Till. Figs. 20-3. Attenuatella sp. A. 20, Latex cast from external mould of dorsal valve. CPC9027 from Till. 21, Internal mould of dorsal valve. CPC9026 from Dul51, 1 mile south-west of Bundaleer Homestead, 37-5 miles north of Bluff, Queensland. 22, 23, Internal moulds of ventral valves. CPC9024 and CPC9025 respectively from Dul51. Figs. 24-6. Attenuatella sp. 24, Internal mould of ventral valve. CPC9028 from Du764, 2 miles west- north-west of Bundaleer Homestead 37-5 miles north of Bluff, Queensland. 25, Latex cast of external mould of specimen in fig. 24. 26, Internal mould of ventral valve (note the absence of ridges along the margins of the platform of muscular attachment). GSQF3459 from GSQL D16, in the north- eastern corner of portion 359, Parish Walloon, Queensland. Palaeontology, Vol. 11 10 PLATE 142 ARMSTRONG, Attenuatella J. ARMSTRONG: BRACHIAL SKELETON OF ATTENUATELLA 789 to moderately convex having a wide median flattening bordered by low ridges for its entire length. Description. The shell is biconvex with an elongate, moderately inflated ventral valve and a semicircular, gently to moderately convex dorsal valve. Cardinal extremities are obtusely rounded and the shell is usually widest at the mid-length of the dorsal valve. The ventral umbo is high, and relatively broad, but is not incurved over the area. The ventral area is apsacline and is about three times as high as the anacline dorsal area, both areas being separated from the remainder of their valves by distinct beak ridges. There is usually a narrow sulcus in the ventral valve corresponding to a median flatten- ing or depression on the dorsal valve. Flanking the depression on the dorsal valve there are two low ridges. Otherwise the dorsal valve is gently convex, being slightly more so umbonally than towards the commissure. The commissure is ligate. The valves are externally ornamented with small variably sized spines which lie along vaguely concen- tric lines and are sometimes developed at the edges of growth lamellae. The lines of spines are between OT and 0-3 mm. apart and along them 9 to 12 spines occur in each millimetre. On some specimens there are large spines interspersed with smaller ones (PI. 142, fig. 5) whereas on other specimens the spines are relatively uniformly sized (PI. 142, fig. 12). A small groove extends anteriorly for a short distance from each spine. The internal features of the ventral valve of A. convexa are similar to those of At- tenuateUa sp. nov. of Armstrong and Brown (1968). There is little umbonal cavity thickening, and in the apical part of the delthyrial cavity there is a small delthyrial plate somewhat depressed below the level of the area. The musculature is similar to that of At- tenuatella sp. nov. of Armstrong and Brown except that in A. convexa the ridges border- ing the muscle platform are less prominent, especially posteriorly; the diductor scars are wider; and the platform of muscle attachment is generally less elevated posteriorly. In the dorsal valve the sockets are prominent and they lie along the margins of the notothyrium. They are defined internally by prominent socket ridges, and medio- anteriorly confluent with each ridge and extending anteriorly for a short distance along the floor of the valve are the crural bases. From these arise anteriorly directed crural lamellae which give rise to the brachidia diagrammatically represented in text-fig. 2. The arrangement of the scars of adductor muscle attachment is shown in text-fig. 4. The cardinal process occupies the apical part of the delthyrium and is tuberculate, the tubercles being roughly arranged in rows radiating from the beak (PI. 142, fig. 13). Comparison. A. convexa is readily distinguished by its massive and only gently curved ventral umbo, its broad shell, and its convex dorsal valve. A. taimyrica from Russia is the species morphologically closest to A. convexa. According to Chernyak’s description and figures, A. taimyrica shares with A. convexa a massive umbo, a gently concave ventral area and a narrow but distinct sulcus. A. taimyrica is, however, distinguished by its flat dorsal valve. Distribution. A specimen from the sandstone unit underlying the Upper Limestone near Gympie is a representative of Attenuatella convexa. The specimen is an internal mould of a ventral valve which had a broad massive umbo and a relatively low muscle platform. This is the only known occurrence in addition to the type locality. Age. The fauna occurring with Attenuatella convexa at the type locality is listed by Armstrong et al (1967, p. 89) in their record of the fauna which occurs with Uraloceras lobulatum Armstrong, Dear, and 790 PALAEONTOLOGY, VOLUME 11 Runnegar. Armstong et al. suggest that the most likely age of this fauna is lower Artinskian (Aktas- tinian) and the similarity between A. convexa and A. taimyrica provides a measure of support for this determination. Attenuatella sp. A Plate 142, figs. 20-3 Material. Several deformed internal moulds of ventral and dorsal valves (CPC9024-6) and a fragment of an external mould fromDu!51 (1 mile south-west of Bundaleer Homestead 37-5 miles north of Bluff, central Queensland). A B text-fig. 4. Arrangement of the scars of muscular attachment in two dorsal valves of Attenuatella convexa sp. nov. A, UQF53031, and B, UQF53032 are figured respectively in Plate 142 figs. 3 and 4. cp, cardinal process; cb, crural base; md, median depressian; pA, posterior abductor scar; aA, anterior abductor scar. Description. Contour of ventral valve similar to that of Attenuatella incurvata; exterior of ventral valve unknown; dorsal valve gently convex and slightly elongate; it is less convex and narrower than dorsal valves of A. convexa but is more convex than dorsal valves of A. incurvata ; on the fragment of external mould the spines number about 10 to 12 per mm. along roughly concentric lines; the interior of the ventral valve is similar to interiors of ventral valves of A. incurvata in that the ridges along the edges of the muscle platform are not as prominent as they are in Attenuatella sp. nov. of Armstrong and Brown, or in A. nmltispinosa Waterhouse. Comparison. Attenuatella sp. A is distinguished from A. incurvata and Attenuatella sp. of Landis and Waterhouse by its elongate gently convex dorsal valve. Attenuatella convexa has a broader shell and a lower, more massive ventral umbo. The most closely comparable species are A. stringocephaloides and A. attenuata both of which have virtually non-sulcate ventral valves and flat or gently convex dorsal valves. The lack of internal details for A. attenuata and A. stringocephaloides of Chernysheva and Likharev and the lack of ventral exteriors of Attenuatella sp. A precludes complete comparison of these species. The interiors of the valves of Chernyak’s specimens of A. stringocepha- loides are known however and again in one of the ventral valves figured by Chernyak J. ARMSTRONG: BRACHIAL SKELETON OF ATTENUATELLA 791 there are only indistinct ridges along the edges of the muscle platform (Ustritskiy and Chernyak 1963, pi. 42, fig. 3a). Chernyak’s specimens have flat dorsal valves and although they are thereby distinguishable from Attenuatella sp. A Chernyak’s specimens seem to be morphologically closest to Attenuatella sp. A. Distribution. A dorsal valve (CPC9027) from T1 1 1 (2-7 miles north-north-west of the Cracow-Theodore road crossing of Delusion Creek, 10 miles north-west of Cracow, Queensland) probably belongs to this species. It is elongate and gently convex, and diverging from the beak there are two low ridges separated by a broad median flattening. Spines on this valve number 12 per mm. along vaguely con- centric lines. Age. The specimens of Attenuatella which in one or another respect are morphologically close to Attenuatella sp. A are recorded from lower Upper Permian faunas suggesting possibly a similar age for Attenuatella sp. A. Attenuatella sp. cf. incurvata Waterhouse 1964 Plate 142, figs. 13-18 Material. Several specimens (CPC9018-23) from the Barfield Formation at Till (above). Description. The specimens have elongate almost non-sulcate ventral valves with moderately inrolled beaks. The only dorsal exterior is very gently convex rostrally and slightly concave commissurally, and two ridges similar to those on the dorsal valves of A. convexa occur on its posterior part. Internal moulds of dorsal valves are more flat. There are about 12 spines per mm. on the valves. The interiors of the ventral valves are similar to those of specimens of A. incurvata. Only indistinct ridges occur along the sides of the areas of muscle attachment in the ventral valve, and there is a pronounced groove separating the adductor scars. Discussion. The specimens from T1 1 1 are compared with A. incurvata on the basis of their non-sulcate ventral valves, on the shape and contour of their ventral and dorsal valves, and on the absence of prominent ridges along the edges of the platform of muscle attachment in the ventral valve. Age. Lower Upper Permian? Attenuatella sp. Plate 142, figs. 24-6 Material. Two specimens (GSQF3459 and CPC9028). The first specimen is from the Flat Top Forma- tion at GSQL D16 (in the north-eastern corner of portion 359, Parish Walloon, 3-3 miles north-east of Theodore, Queensland) and the other specimen is from Du764 (2 miles west-north-west of Bundaleer Homestead, 37-5 miles north of Bluff, central Queensland). Description. Both specimens are internal moulds of ventral valves and are characterized by their posterior elongation and their incurved beaks. The external mould of the speci- men from Du764 is non-sulcate. In the interiors of the valves there is an elevated plat- form of muscular attachment, but as in specimens of Attenuatella sp. A and A. sp. cf. incurvata only indistinct ridges occur along the sides of the platform. In both specimens the areas of adductor muscle attachment are separated by a prominent groove. The specimens are probably conspecific with either Attenuatella sp. A or 4. sp. cf. incurvata but the absence of dorsal valves in association with the specimens from D16 and Du764 does not permit their correct assignment. 792 PALAEONTOLOGY, VOLUME 11 Acknowledgements. I would like to thank Professor D. Hill, F.R.S. for reading the manuscript and for making a number of helpful suggestions. The work was carried out at the Department of Geology, University of Queensland, while the author was in receipt of a Commonwealth Post-graduate Award. The photographs are the work of Mr. J. Coker and Mr. E. Hollywood of the Photographic Depart- ment of the University of Queensland. REFERENCES Armstrong, j. d. and brown, c. d. (1968). A new species of Attenuatella from the Permian of Queens- land. Mem. Qd Mus. 15 (2), 56-64. dear, j. f. and runnegar, b. 1967. Permian ammonoids from eastern Australia. J. geol. Soc. Aust. 14, 87-97. chernyak, g. e. 1963. In V. I. Ustritskiy and G. E. Chernyak. Biostratigrafiya i brakhiopody verkhnego paleozoya Taymyra. Trudy nauchno-issled. Inst. Geol. Arkt. 134, 1-139. (Biostratigraphy and brachiopods of the Upper Permian of Taimyr.) cloud, p. e. 1944. Permian Brachiopoda. In R. E. King, C. O. Dunbar, P. E. Cloud, and A. K. Millar. Geology and palaeontology of the Permian area north-west of Las Delicias, south western Coahuila, Mexico. Spec. Pap. geol. Soc. Am. 52, 1-72. dear, J. f. 1966. An ammonoid from the Permian of Queensland. Mem. Qd Mas. 14 (5), 199-203. dickins, j. m. 1964. Correlation and subdivision of the Permian of Western and eastern Australia. Rep. Int. geol. Congr. 22, Abstracts, 115-16. Fredericks, G. 1919. Etudes paleontologiques. 2. Sur les spiriferids du Carbonifere superior de l’Oural. Izv. geol. Korn. 38 (3), 295-324. george, t. n. 1931. Ambocoelia Hall and certain similar British Spiriferidae. Q. Jl geol. Soc. Lond. 87, 30-61. krotova, p. 1885. Artinskiy Yarus. Geologopaleontologicheskaya monografiya Artinskogo pescha- nika. Trudy Obshch. Estest. imp. khar'kov. Univ. 13 (5), 314 pp. (Artinskian Stage. Geological- palaeontological monograph of the Artinsk sandstones.) landis, c. a. and Waterhouse, j. b. 1966. Discovery of Permian fossils in the Eglinton Volcanics, Southland. Trans, roy. Soc. N.Z. Geol. 4 (6), 139-46. likharev, b. k. and einor, o. l. 1939. Materialy k poznaniyu verkhnepaleozoyskikh faun Novaya Zemli, Brachiopoda. Trudy arkt. nauchno-issled. Inst. 127, 1-245. (Contributions to the knowledge of the upper Palaeozoic faunas of Novaya Zemlya, Brachiopoda.) stehli, f. g. 1954. Lower Leonardian Brachiopoda of the Sierra Diablo. Bull. Am. Mus. nat. Hist. 105 (3), 261-358. stepanov, d. L. 1957. O novom yaruse permskoy sistemy Arktiki. Vest, leningr. gos. Univ. 24, ser. geol.- geog. 4, 20-4. (A new stage of the Permian system in the Arctic.) ustritskiy, v. i. 1960. O granitse nizhney i verkhney Permi v Pechorskom basseyne i v Arktike. Trudy nauclmo-issled. Inst. Geol. Arkt. 1 14. (On the boundary between the Lower and Upper Permian in the Peckora Basin and in the Arctic.) and chernyak, g. e. 1963. Biostratigrafiya i Brakiopody verkhnego paleozoya Taymra. Trudy nauchno-issled. Inst. Geol. Arkt. 127, 1-245. (Biostratigraphy and brachiopods of the Upper Permian of Taimyr). vandercammen, a. 1956. Revision des Ambocoeliinae du Devonien de la Belgique. Bull. Inst. r. Sci. nat. Belg. 32 (43), 1-51. veevers, j. 5. 1959. Devonian brachiopods from the Fitzroy Basin, Western Australia. Bull. Bur. Miner. Re sour. Geol. Geophys. Aust. 45, 220 pp. Waterhouse, j. b. 1964. Permian Brachiopoda of New Zealand. Bull. geol. Surv. N.Z. Palaeont. 35, 1-287. 1967. A new species of Attenuatella (Brachiopoda) from Permian beds near Drake, New South Wales. Rec. Aust. Mus. 21 (7), 167-73. john Armstrong Department of Geology University of Queensland St. Lucia Brisbane, Queensland 4067 Typescript received from author 2 May 1968 Australia TAXONOMIC PROBLEMS IN THE STUDY OF COCCOLITHS by MAURICE BLACK The Eleventh Annual Address, delivered 6 March 1968 abstract. From a review of Jurassic, Cretaceous, Tertiary, and Recent coccolithophorid and discoaster material, it is suggested that the abrupt appearance of some complex forms, hitherto difficult to classify, may be explained by mineralization of already existing unmineralized stocks; it also seems likely that some living un-mineralized forms may have fossil mineralized ancestors. Some recent work on uncalcified Haptophyceae reveals interesting parallels with the coccolithophorids. One new family, Stephanolithiaceae, is erected, and four specific names are recorded in new combination. Living in the present oceans there are about 200 species of coccolith-bearing algae. About half of these fall quite naturally into two big families; the rest make up a pecu- liarly heterogeneous collection, with numerous monospecific genera which are very difficult to group into families or higher taxa. Attempts to devise a scheme of classifica- tion for these organisms have not been entirely successful. The simpler schemes tend to force very diverse genera into an inappropriately small number of families, as may be seen in many reports on quantitative plankton-surveys. Some of the more elaborate schemes try to solve the problem by building up a theoretical phylogenetic tree which is used as a basis of classification. For this purpose, the evidence has been drawn exclusively from the comparative morphology of living species, and up to the present palaeonto- logical observations have been almost completely ignored. Twenty years ago, this may have been excusable, but it is not so today, for there is now enough information at hand for us to start establishing phylogenetic series on the evidence of the fossil record. When we attempt to do this, one of the first results is the discovery that the two largest families, the Syracosphaeraceae and Zygosphaeraceae, each with more than 50 living species, have virtually no geological record at all. With the smaller taxa, the situation is quite different; many of the systematically isolated genera are found to have long ancestries, and some are survivors from once prosperous Mesozoic or Tertiary families. Clearly, these survivors are more closely related to their own fossil ancestors than they are to living genera of different parentage, and it is no longer reasonable to lump all the living coccolith-bearing algae into a single family, as has been done in some of the most recently published classifications (Kamptner 1958, pp. 68-71, Deflandre 1966, pp. 5-6). Considerations such as these suggest that the time has now arrived when we must look carefully at the impact of palaeontological research upon the taxonomy of living forms. During the last two decades there has been a strong tendency to deny the possi- bility of recognizing fossil representatives of many living taxa above the rank of species. This is, of course, a strictly logical attitude, because the micropalaeontologist has to work largely with individual coccoliths, and some modern genera are defined in terms that take into account the way in which the coccoliths are assembled to form a complete skeleton. Nevertheless, I think it is an unnecessarily defeatist attitude, because it assumes that each fossil coccolith is to be considered as a completely isolated entity, with no [Palaeontology, Vol. 11, Part 5, 1968, pp. 793-813, pis. 143-54.] 794 PALAEONTOLOGY, VOLUME 11 ancestors and no descendents. This, of course, is not true, and I hope to show that many living taxa can be traced back with reasonable confidence into the Tertiary, and a few can be followed still further into the Mesozoic. In this way, family relationships sort themselves out as phylogenetic lineages, and the theoretical difficulties of working with isolated coccoliths fall into the background. PROBLEMS IN THE CLASSIFICATION OF LIVING COCCOLITHOPHORES From the early days, it has been found satisfactory to identify and classify the living species on the basis of coccolith-structure. No two species are known which have exactly similar coccoliths, and, as a working hypothesis, it is not unreasonable to assume that this also holds good for extinct species. Coccolith-structure has likewise been relied upon extensively for the delimitation of genera, but other characters such as the shape of the cell and the presence or absence of a naked area at the flagellar pole have also been used to some extent in defining genera and higher taxa. Living taxa defined purely in terms of coccolith-morphology can be recognized with certainty when they are found as fossils, but characters involving the arrangement of coccoliths on the cell-surface are not as a rule observable in fossil material. With this difficulty in mind, some micro- palaeontologists have argued that all fossil coccoliths should be assigned to provisional taxa until a complete coccosphere has been discovered. This practice has the serious disadvantage of requiring a dual nomenclature, with one set of names for living material and another for fossils, thus creating an artificial break between modern forms and their fossil ancestors. In order to resolve this problem, we will first of all look more closely at some of the taxonomic difficulties presented by living coccolithophores, and then try to discover just how serious these difficulties really are when we try to work out some sort of phylo- genetic story. Dimorphism of coccoliths Lohmann (1902) found that in some species the coccoliths covering a single individual were not all alike; for example, those round the flagellar pole might bear spines, whereas the ordinary body-coccolith did not. Usually the polar coccoliths are easily recognizable as straightforward modifications of the normal type, and Lohmann did not in general segregate species with this kind of dimorphism into separate genera. The only genus which he distinguished on these grounds was Scyphosphaera , which has an equatorial girdle of very large float-coccoliths. Taxonomic difficulties did not arise until 40 or 50 years later, when Kamptner (1941) and Deflandre (1952) took a much more serious view of coccolith-dimorphism, and removed all species with dimorphic coccoliths into separate genera. The consequent practical difficulty in naming fossil coccoliths was perhaps unnecessarily exaggerated, but the theoretical implications led Deflandre to abandon the use of a single classification for both living and fossil forms, and to create an independent taxonomic system for the fossils. The general adoption of this policy during the decade that followed did much to retard progress in the study of phylogenetic and taxonomic relationships among the fossil coccoliths. Coccolith-dimorphism is likely to have been prevalent to much the same extent in M. BLACK: TAXONOMIC PROBLEMS IN THE STUDY OF COCCOLITHS 795 fossil as in living taxa, and although final proof is usually beyond our reach, it is not difficult to find fossil examples of morphological variants that have all the earmarks of dimorphic pairs. These are not uncommon in Cretarhabdus (PI. 150, figs. 4, 5), Deflan- dr ius (PI. 151, figs. 4, 5), Kamptnerius (PI. 152, figs. 5, 6), and Rhabdolithina\ if fossil species with dimorphic coccoliths are not recognized as such, the worst that can happen is that two species with identical ranges appear in the fossil record instead of one. Pleomorphic life-histories Observations on the life-histories of species kept in culture have brought to light several examples in which a motile coccolith-bearing phase alternates with a sessile uncalcified phase which had previously been accepted as a member of a different genus (Parke 1961). This is particularly true of coastal species, whose non-motile phases tend to live attached to rocks. Distinct from these are the pelagic species, whose resting-stages take the form of calcified cysts, which in the absence of any solid support, remain suspended in sea-water. One of these species, Crystallolithus hyalinus, which bears organic scales, sometimes unmineralized (PI. 154, figs. 1, 2), and sometimes covered with minute rhombohedra of calcite, has proved to be the motile phase of Coccolithus pelagicus , an encysted form with coccoliths of a totally different kind (PI. 143, figs. 1, 2). These unexpectedly complicated life-histories are interesting from a taxonomic point of view, because the individual phases are usually so different from each other that they were put into separate genera when first discovered. To judge from our knowledge of living species, we can hardly expect to find tangible evidence of pleomorphism in the fossil record. In the coastal species, all the known sessile phases are uncalcified, and hence unlikely to be preserved as fossils. In the pelagic species, on the other hand, it is only the heavily calcified cysts, or their component coccoliths, that stand much chance of being preserved; delicate motile forms like Crystallolithus soon disintegrate, and are never found in bottom deposits. Indeed, material from living cultures needs careful handling in the laboratory to prevent the organic scales from losing their cover of calcite crystals. The cysts, on the other hand, are often preserved entire, and good examples are quite common in rocks as old as the Kimeridge Clay (PI. 143, fig. 6). Variation in response to external conditions Changes in external conditions can, in some circumstances, cause the organism to modify the appearance of its coccoliths, or even to cease calcification. In most species that have been experimented upon, the principal effect of temperature- change is to alter the rate of cell-multiplication, without any significant effect upon coccolith-morphology. In Coccolithus huxleyi, however, the coccoliths grown in warm water are reported to be distinguishable from those grown at low temperatures (PI. 145, figs. 1, 2); the two variants have been recorded in laboratory cultures (Watabe and Wilbur 1966) and in natural populations (McIntyre and Be 1967). C. huxleyi is kept in culture in several laboratories, and a number of separate strains are kept under constant observation. Some of these have been found quite suddenly to stop growing coccoliths, and to persist in a naked condition (Paasche 1964, p. 11). One cause of this change appears to be an enrichment in the supply of nutrients, par- ticularly phosphate. An interesting point is that whereas some strains will resume growth 796 PALAEONTOLOGY, VOLUME 11 of coccoliths if the culture medium is adjusted to a composition more like natural sea- water, others will not, and no known treatment will induce them to do so. There are several very suggestive possibilities here in connection with the evolution and taxonomy of these algae; fixation of a single phase out of a pleomorphic life-history, independent development of ecological variants, and sudden changes from a naked to a calcified condition, could all lead to the introduction of apparently cryptogenetic forms into the geological record. Furthermore, some of these changes could take place very suddenly, owing to the extraordinarily fast rate of multiplication that is prevalent in these organisms (Parke and Manton 1962). PHYLOGENETIC HISTORY OF SOME LIVING TAXA Attempts to trace living coccolithophorids back into geological history lead to curi- ously divergent results. Some species such as Coccolithus pe/agicus, C. huxleyi, and Braarudosphaera bigelowi lead back to flourishing and diverse complexes, which in the present oceans have been reduced to just one or two surviving representatives. The ancestral stocks show a good deal of evolutionary divergence, but a set of unifying family characters persists throughout. Pontosphaera provides an extreme example of divergence, for we not only get a hint of an evolutionary connection with the extinct Zygolithaceae, but we also find here the first suggestion of a far-distant linkage with another living family, the Helicosphaeraceae. Other stocks seem to have been peculiarly stable throughout their geological history. Calciosolenia can be traced back to mid- Cretaceous times with very little change, and the coccoliths of Braarudosphaera bigelowi are specifically indistinguishable from fossils preserved in early Cretaceous rocks. In contrast with these are two flourishing modern families, the Syracosphaeraceae which are abundantly represented in contemporary Globigerina Ooze, but have not been found in deposits older than Quaternary, and the Zygosphaeraceae which are not represented in bottom deposits at all. Coccolithus pelagicus (Wallich) Schiller Plate 143, figs. 1, 2 This is the most familiar of all coccoliths; it was discovered in samples of ooze raised from the floor of the North Atlantic Ocean during the first telegraph surveys, and was examined by Huxley and Wallich, and later by Murray and Blackman (1898). The living EXPLANATION OF PLATE 143 Figs. 1,2. Coccolithus pelagicus (Wallich) Schiller, recent Globigerina Ooze, Discovery II Sta. 4269, Biscay Abyssal Plain. 1, No. 20175, oblique distal view, x 4800. 2, No. 5000, details of central area, distal view, X 5000. Fig. 3. Coccolithus sp. cf. C. cavus Hay and Mohler; No. 22279, Upper Oligocene, core DWBG 10, Pacific Ocean; distal view, x 5300. Fig. 4. Coccolithus sp. cf. C. marismontium Black; No. 22284, Upper Oligocene, core DWBG 10, Pacific Ocean; proximal view, X6700. Figs. 5, 6. Ellipsagelosphaera sp. 5, No. 15230, Campanian Chalk, Belgorod, Russia; distal view, x8000. 6, No. 22874, Cambridge Greensand, Cherry Hinton Fields, near Cambridge; complete coccosphere (calcified cyst), x 3700. Palaeontology, Vol. 11 PLATE 143 BLACK, Living, Tertiary and Cretaceous coccoliths M. BLACK: TAXONOMIC PROBLEMS IN THE STUDY OF COCCOLITHS 797 alga is confined to the North Atlantic, and is the only species that is known to have a preference for cold water; all attempts to find it in Antarctic waters of suitable tempera- ture have failed. This was apparently not the case during the Pleistocene, for Allan Be and his colleagues at the Lamont Observatory have found it in core-samples from the southern oceans at levels dated as belonging to the middle of the Wisconsin glacial episode (McIntyre and Be 1967). In the Tertiary, there are very similar forms, but studies under the electron microscope show that they are specifically distinct (PI. 143, fig. 3). The same generic form can be traced back to the Eocene, but once we pass into the Mesozoic, the species that superficially resemble C. pelagicus are obviously different in the finer details of their construction. When the coccoliths are examined under a petro- logical microscope, with the fossil lying flat on the slide, Tertiary and living species give an interference figure in which the distal shield plays no part, since the optic axes of the component calcite crystals are at right-angles to the plane of the shield. In the Mesozoic genus Ellipsagelosphaera this is not the case, and the crystals of the distal shield show strong birefringence, so that the optical figure takes a complicated form, due to the superposition of two black crosses, one produced by each shield. In addition, the Meso- zoic species always have a corona of quadrate or keystone-shaped crystals lying on top of the distal shield, and marking off the central area (PI. 143, fig. 5). Under an ordinary microscope the Tertiary and Mesozoic species look very much alike, but under a polariz- ing microscope or an electron microscope they are so obviously distinct that they are now placed in separate subfamilies. In the Jurassic and Cretaceous there are also several other genera such as Sollasites (PI. 144, figs. 1, 2), which resemble Coccolithus in a general way, but differ considerably in the finer details of their structure. They are clearly not on the main line of descent towards the C. pelagicus stock, but some of them may prove to be ancestral to certain other Tertiary and living forms. Cyclococcolithus leptoporus (Murray and Blackman) Kamptner Plate 144, figs. 3, 4; Plate 147, fig. 1 On theoretical grounds, Kamptner has argued that the circular outline is more primi- tive than the elliptical, and has insisted upon the taxonomic separation of these two shapes (compare PI. 143, figs. 1-6, PI. 144, figs. 3-7). He remarks: ‘Above all else, one clear-cut taxonomic principle is decisive for subdivision . . . into tribes and subtribes: a sharper separation of the circular from the elliptical forms. A primitive character must be attributed to the circular outline, and a derivative character to the elliptical. ... It is also a priori conceivable that the change-over from the circular to elliptical types has been achieved polyphyletically, and so to speak on a broad front’ (Kamptner 1958, p. 64). In accordance with this principle, he removed Coccolithus leptoporus to a new genus, Cyclococcolithus, and similarly split up other genera so that species with circular coccoliths could be put into different subtribes from those with elliptical coccoliths. Cyclococcolithus leptoporus provides a convenient starting point from which to examine this taxonomic principle in relation to the geological record. This species is abundant in the living plankton, and is widely distributed. Very similar coccoliths are common in the Pliocene, and particularly so in core-samples from the ocean floor (PI. 144, fig. 4). 798 PALAEONTOLOGY, VOLUME 11 In such deposits, the two discs of the placolith are apt to become separated, and many records of Tiarolithus and Calcidiscus from the Upper Tertiary are probably based upon dismembered specimens of C. leptoporus or closely related species. Other species are found in the Lower Tertiary, and the earliest record of the genus is from the Danian. These early forms occur associated with characteristically elliptical species of Cocco- lithus, and the two stocks were quite distinct from each other at the time of their first appearance in the early Tertiary. Thus the Tertiary record does not give much support to the idea that the circular shape is primitive; indeed, it tends to emphasize the complete independence of the elliptical forms rather than to provide any suggestion of their descent from circular ancestors. This independent relationship is emphasized by a curiously similar state of affairs in the Upper Jurassic, for a different set of circular and elliptical types exist side by side in the Oxford and Kimeridge Clays. These include the genera Cyclagelosphaera and Ellipsagelosphaera (Noel 1965), which at first sight appear to be so much like C. lepto- porus and C. pelagicus that they were originally recorded under these names. This resemblance, however, is illusory, for there are significant differences in micro-structure. It is remarkable that the two Jurassic genera each differ from their Tertiary analogues in exactly the same way: they both have strongly birefringent distal shields, surmounted by a well-developed corona (compare PI. 143, fig. 3 and PI. 144, fig. 3 with PI. 143, fig. 5 and PI. 144, fig. 5). Consequently, if we focus our attention on the minutiae of crystal- arrangement, we find that the Jurassic circular forms resemble their elliptical contem- poraries more closely than they resemble modern circular forms. Are we then to unite the Jurassic genera, both circular and elliptical, into one family or subfamily, as Noel (1965) has done, or would it be more reasonable to keep the two shapes separate, regarding them as two parallel stocks that became independent at an early stage, and have remained so ever since? An attempt to answer this question requires a closer look at the geological history of these two stocks. The elliptical forms are abundant throughout the Mesozoic and Cainozoic, and it may be that within this multitude of species there exists a continuous evolutionary thread leading from the Jurassic to the present day. The only serious break in this record is at the Cretaceous-Paleocene boundary, and there is not enough information available at present to reach a decision. The history of the circular placoliths is rather different. After the great abundance of Cyclagelosphaera (PI. 144, fig. 5) in the Jurassic, there is a paucity of circular forms in the Cretaceous. The few examples that EXPLANATION OF PLATE 144 Figs. 1, 2. Sollasites horticus (Stradner et at.) comb. nov. ( Coccolithus horticus Stradner et at.), Cambridge Greensand near Cambridge, X9600. 1, No. 4706, Barrington, distal view. 2, No. 21739, Cherry Hinton Fields, proximal view. Figs. 3, 4. Cyclococcolithus spp.. Lower Pliocene, core LSDH 78P, Pacific Ocean. 3, cf. C. leptoporus (Murray and Blackman) Kamptner, No. 22224, distal view, x4000. 4, No. 22210, group with some specimens in the ‘Calcidiscus’ condition, x 2500. Fig. 5. Cyclagelosphaera margereli Noel; No. 17410, Oxford Clay (L. Oxfordian), Cambridge Experi- mental Borehole; distal view, x 10 000. Fig. 6. Markalius sp. cf. M. inversus (Deflandre) Bramlette and Martini; No. 17199, Belemnite Marl (L. Turonian), Cherry Hinton, near Cambridge; proximal view, x 6000. Fig. 7. Ericsonia sp.; No. 22295, Upper Oligocene, core DWBG 10, Pacific Ocean; proximal view, X 6000. Palaeontology , Vo l 11 PLATE 144 BLACK, Cretaceous, Tertiary and Jurassic coccoliths M. BLACK: TAXONOMIC PROBLEMS IN THE STUDY OF COCCOLITHS 799 are known, such as Markalius (PI. 144, fig. 6), can hardly be regarded as linking Cycla- gelosphaera with Cyclococcolithus, because they have a more complicated structure than either of these. They seem rather to be leading towards Ericsonia (PI. 144, fig. 7) and other still more specialized forms that flourished in the Lower Tertiary. There is thus a gap extending throughout the Cretaceous period, during which they were apparently no representatives of the hypothetical Cyclococcolithus lineage. It seems, therefore, unlikely that the Tertiary species originated as an offshoot from Cyclagelosphaera, and it is more probable that they developed from an unknown ancestor early in the Tertiary. The evidence from the geological record thus seems to alienate the Mesozoic circular forms from their modern analogues, and the evidence of morphology throws them into association with their elliptical contemporaries, in spite of the difference of shape. Coccolithus huxleyi (Lohmann) Kamptner (= Emiliania huxleyi (Lohm.) Hay and Mohler 1967) Plate 145, figs. 1, 2 The taxonomic position of this species is surrounded by interesting problems. It stands apart from C. pelagicus and all other living members of the genus in the peculiar construction of its shields, and its conspicuous central grid, which recalls the similar structures in many extinct forms. Mary Parke in Parke and Dixon (1964, p. 520) has pointed out that the motile phase lacks an obvious haptonema, and in this respect it resembles the Isochrysidales rather than the Coccolithophorales. Paasche (1964, p. 11) found in his cultures that coccolith-secreting individuals were without visible flagella, although these could be seen on some of the naked cells. In the living plankton, it is undoubtedly the most vigorous and successful of the coccolith-bearing algae. It is distributed over the whole area of the oceans from the Antarctic convergence to the Arctic, and can invade brackish waters inaccessible to other pelagic species. Because of its great vigour and tolerance, it has provided material for more laboratory experiments than any other species. C. huxleyi is apparently a very recent addition to the oceanic plankton. McIntyre and Be (1967) have reported its first appearance in deep-sea cores as taking place within Brunhes Normal Zone (that is, less than 700 000 years ago, and more probably nearer to 100 000 years). They also report the presence of a coocolith intermediate between C. huxleyi and Gephyrocapsa oceanica in cores spanning the interval between 300 000 and 100 000 years. G. oceanica (PI. 145, fig. 3) can be traced back into the Pliocene, and survives in the living plankton. Forms with the oblique bridge characteristic of Gephyro- capsa are unknown before the Pliocene, but the other characters of the genus, such as the large central opening with a bilateral grid, and the non-imbricate ray-elements, are strongly developed in a host of Lower Tertiary species of Tremalithus (PI. 145, figs. 4-6). There is a great diversity of species with elaborate grids in the Middle and Upper Eocene, and the same type of coccolith can be traced back still further into the Mesozoic. The earliest representatives, with rather simpler grids, are found in the Lower Gault j (Middle Albian) at Folkestone (PI. 145, figs. 7, 8). The taxonomic isolation of C. huxleyi from the rest of the genus Coccolithus, which is suspected from the peculiarities of the living alga, is thus confirmed by its long and independent geological history, and there can be little doubt that this species should be 800 PALAEONTOLOGY, VOLUME 11 removed to a separate genus, and possibly to a family independent of the Coccolitho- phoraceae.1 Pontosphaera discopova Schiller Plate 146, figs. 1 , 2 The genus Pontosphaera was created by Lohmann (1902) for species in which the coccoliths appear to be simple discs, with or without a shallow rim. The outline is elliptical, and the coccoliths are most commonly shaped like a shallow pie-dish. Specific diagnoses were originally based upon work under the biological microscope, and later studies in polarized light or under the electron microscope have led to the removal of several species to other genera. No electron micrographs of the type species, P. syracu- sana , have yet been published, but Mrs. Gaarder has generously allowed me to compare some of her excellent micrographs of this species with my Tertiary material. The cocco- liths are large, with a floor perforated by about 750 very delicate pores, and the wall is constructed of about 200 very slender and steeply inclined calcite laths. P. discopova has similarly constructed but smaller coccoliths, with larger and less numerous pores. The Tertiary coccoliths discussed below have the same general type of structure, and can with reasonable confidence be referred to the same genus as P. syracusana and P. discopova. Coccoliths of various species of Pontosphaera are not uncommon in modern Globi- gerina Ooze, particularly beneath the warmer parts of the oceans. In the Pliocene a form closely resembling P. discopova is occasionally found (PI. 146, figs. 4, 5), but other species are rare or absent, and none of the modern perforated species has yet been identified with certainty in pre-Pliocene sediments. Nevertheless, coccoliths with all the charac- teristic features of Pontosphaera except the circular pores in the floor, are widely dis- tributed throughout the Tertiary. There appear to be several lineages among these Tertiary species, each with a slightly different pattern in the structure of the floor. One of these is of special interest, since it appears to connect the living P. discopova with a complex of Eocene forms whose ancestry can probably be traced back to the Cretaceous. The proximal surface of the floor in this series has a bilateral arrangement of plates similar to that originally described by Kamptner in P. scutellum, and now known in a number of diverse Eocene forms (PI. 146, fig. 6). The distal side has an entirely different EXPLANATION OF PLATE 145 Figs. 1,2. Coccolithus huxleyi (Lohmann) Kamptner; recent oceanic deposits. 1, warm-water form, No. 3421, Challenger Sta. 338, S. Atlantic Ocean; distal view, x 16 000. 2, cold-water form, No. 11612, Discovery II Sta. 3809, Galicia Bank; distal view, X 20 000. Fig. 3. Gephyrocapsa oceanica Kamptner; No. 18072, modern Globigerina Ooze, Discovery II Sta. 4288, Biscay Mts., proximal view, X 8000. Figs. 4-8. Tremalithus spp. 4, T. danicus (Black) comb. nov. (= Dictyococcites danicus Black); No. 11819, Middle Oligocene, Grundfor, Denmark; proximal view, X 6000. 5, T. placomorphus Kamptner; No. 22838, Lower Oligocene, Rodstenseje nr. Odder, Denmark; proximal view, X 4000. 6, T. dictyodus (Defiandre and Fert) comb. nov. (= Discolithus dictyodus D. and F.); No. 22831, Lower Oligocene, Rodstenseje nr. Odder, Denmark (possibly reworked from M. Eocene); proximal view, X 6100. 7, T. burwellensis Black; No. 14622, Cambridge Greensand (Cenomanian), Barrington near Cambridge; distal view with proximal shield showing through, X 10 000. 8, T. sp., No. 13371, Lower Gault (M. Albian), Folkestone, Kent; proximal view, X 8000. 1 Since this lecture was delivered, a paper by W. W. Hay et al has been received, in which C. huxleyi has been formally transferred to a new genus, Emiliania. Palaeontology , Vol. 11 PLATE 145 BLACK, Living, Tertiary and Cretaceous coccoliths M. BLACK: TAXONOMIC PROBLEMS IN THE STUDY OF COCCOLITHS 801 appearance, with a pattern of concentrically arranged fibrous elements (PI. 146, figs. 3, 7). This combination of patterns is also seen in several species of Helicosphaera (PL 147, figs. 1-3), and will be considered further in connection with the ancestry of that genus. This constructional pattern can be picked up again in the late Cretaceous. Two species in the Santonian and Campanian have the same kind of wall-structure, and the same arrangement of plates on the proximal surface. One of these, Pontolithina moorevillensis, has two small perforations at the foci of the ellipse (PI. 149, fig. 4). This species may well prove to be the ancestor of several Tertiary linages. Obliteration of the two pores and an increase in the number of wall-elements would give a form very much like certain Eocene species, for example P. versa Bramlette and Sullivan (PI. 146, fig. 6), P. vadosa Hay et al., and possibly other forms like Discolithus oamaruensis Dell. Enlargement of the pores to make two circular windows would lead to such forms as D. duoeavus Bram. and Sull. (L. Eocene) and D. panarium Defl. (M. Eocene). There are also several Eocene species which differ in having numerous small circular pores, but which probably have a comparable floor-structure, since they give similar interference patterns in polarized light; their ultra-structure has not yet been examined under the electron microscope. D. punctosus Bram. and Sull. (L. Eocene) and D. dis- tinctly Bram. and Sull. (M. Eocene) have this structure; they foreshadow the modern perforate species of Pontosphaera, but in the absence of any recognizable intermediates in the Upper Eocene and Miocene, we cannot say whether there is any direct phylo- genetic connection. Possibly D. vigintiforatus, the type species of Discolithus from the Miocene of the Vienna basin, may yet prove to be on this line of descent. There can be little doubt that Pontosphaera has a long geological history, and that many Tertiary species that have been referred to the rather unsatisfactory form-genus Discolithus actually fit quite naturally into one or other of the lineages of this complex. A derivation from some branch of the Mesozoic Zygolithaceae seems quite possible, and is indeed suggested by the presence in the Upper Cretaceous of forms that are inter- mediate between the two families (PI. 149, fig. 4). Helicosphaera carteri (Wallich) Kamptner (= Helicopontosphaera kamptneri Hay and Mohler 1967) Plate 147, figs. 1, 2 In the present-day oceans, the genus Helicosphaera is represented by a single species whose coccoliths are peculiar in having a spiral brim which terminates in a charac- teristically flaring wing. Within this brim is an elliptical shield, shaped rather like the crown of a hat; it has bilaterally arranged plates on the proximal side, and concentric fibres on the distal, so that the structure is much the same as in many species of Ponto- sphaera. This species has been customarily regarded as closely related to, and possibly derived from taxa such as Coccolithus with typical placoliths. Study of H. carteri under the electron microscope does not give much support to this idea; the coccoliths are clearly not mis-shapen placoliths, as is often assumed, and it is difficult to see how the spiral flange could have been derived from the two shields of a placolith. Compared with other living coccolithophorids, H. carteri stands very much by itself, and its geological history abundantly confirms its independence from the Coccolithophoraceae. 802 PALAEONTOLOGY, VOLUME 11 At the present day H. carteri is confined to the parts of the oceans lying roughly between latitudes 50° N. and S. ; in bottom sediments its range is withdrawn a little towards the equator (McIntyre and Be 1967, p. 585). In the Pliocene, forms that would be difficult to separate from this species are widely distributed, both in deep-sea cores and in samples collected on land. Similar forms, still very much like H. carteri, are abundant in the Upper Miocene, but here they are associated with several extinct species which are obviously quite different. In the underlying parts of the Miocene and in the Oligocene and Eocene there is a varied complex of extinct species, and the genus was evidently undergoing a burst of evolutionary diversification, possibly at its climax during the Oligocene. In the early Eocene there are several species whose coccoliths lack the broad, flaring wing that is so characteristic of later species; they have a shape that does not depart much from an ovoid or even a regular ellipse. The earliest of these, H. seminulum , evolves from a Lower Eocene population in which this simple type (subspecies seminulum ) is predominant (PI. 147, fig. 4), to a Middle Eocene population in which the predominant subspecies lophotus is much more like a typical Helicosphaera (Bramlette and Sullivan 1961). The rim in subsp. seminulum shows little more than a suggestion of spiral struc- ture, and in fact resembles the wall of an early Tertiary Pontosphaera crossed by an oblique wrinkle. The resemblance between these primitive species of Helicosphaera and some of the contemporary species of Pontosphaera is so close that they might reasonably be placed in the same genus, were it not that forms like H. seminulum are so plainly ancestral to other more typical species of Helicosphaera. We have thus traced the ancestry of H. carteri back to an early stage when the genus was barely distinguishable from Pontosphaera, and we are now in a better position to assess its taxonomic status. It clearly has no close relationship with the Coccolithopho- raceae, and must be placed in some other family. It has sprung from the same ancestors as the living Pontosphaera, and for this reason might be included in the Pontosphaera- ceae. On the other hand, it can be argued that the divergence between the two stocks since the early Tertiary has been so profound that the living representatives can hardly be included in a single family, and I have proposed elsewhere that a new family, the Helicosphaeraceae, should be established for H. carteri and the numerous Tertiary species of which it is the sole survivor (Black 1968). Braarudosphaera bigelowi (Gran and Braarud) Deflandre Plate 147, fig. 5 In 1935 a new species, thought to be a Pontosphaera, was recorded in plankton hauls from the Bay of Fundy (Gran and Braarud 1935); this was re-examined by Deflandre, EXPLANATION OF PLATE 146 Figs. 1-7. Pontosphaera spp. 1-3, P. sp. cf. P. discopora Schiller, recent Globigerina Ooze, Challenger Sta. 338, S. Atlantic Ocean. 1, No. 11393, proximal view, X 7200. 2, No. 11318, distal view, X 8000. 3, No. 11319, details of distal surface, X 24 000. 4, 5, P. sp. cf. P. discopora Schiller, Pliocene, Cisano nr. Albenga, Italy. 4, No. 16852, proximal view, X4800. 5, No. 16854, oblique proximal view, X 4000. 6, 7, P. versa (Bramlette and Sullivan) comb. nov. (= Discolithus versus B. and S.), Eocene, Tuilerie de Donzacq, Landes, France. 6, No. 15855, proximal view, x6000. 7, No. 15834, distal view, X 6000. Palaeontology , Vol. 11 PLATE 146 BLACK, Living and Tertiary coccoliths I M. BLACK: TAXONOMIC PROBLEMS IN THE STUDY OF COCCOLITHS 803 who showed that its coccoliths were entirely different from any previously known form. They consist of five peculiarly shaped calcite plates arranged to form a regular pentagon, and twelve of these pentaliths fit together to make a dodecahedral skeleton enclosing an encysted cell. The species was re-named Braarudosphaera bigelowi , and a new family was based upon it (Deflandre 1947). At that time little was known about its geological history, which we now know to be quite remarkable. Coccoliths with exactly the same structure have been found at various levels in the Tertiary, and are abundant in the Upper Bracklesham Beds of the Sussex coast. They have also been found, though less abundantly, throughout the Cretaceous system, and except for an increase in size in the older specimens, are specifically indistinguishable from recent material. No other species of coccolithophorid has a stratigraphical range at all comparable with this, and at first sight it might be thought that we have here an extreme example of specific stability. This is no doubt true of the main lineage of B. bigelowi itself, but in the Tertiary we are confronted not only with a multiplicity of other species of Braarudosphaera itself, but in addition an exuberant development of two closely related genera, Micrantholithus and Pemma (PI. 147, fig. 6), all with coccoliths constructed on the same pentalith system. Detailed phylogenies have still to be worked out, but some of the simpler forms are so close to B. bigelowi that there can be little doubt about their origin. B. bigelowi is an unquestionable example of a single living species that is the sole survivor of an important and diversified Tertiary family whose ancestry can be traced back well into the Mesozoic. This point is stressed because there are several other species in the living plankton with very much the same kind of history, resulting in a taxonomic isolation which is equally real, but by no means always so obvious. calciosoleniaceae Kamptner Plate 148, figs. 1, 2 Most coccolithophorids have a spherical, egg-shaped, or pear-shaped body. The Calciosoleniaceae differ in being cylindrical or fusiform, and their coccoliths instead of being circular or elliptical, take the form of a narrow parallelogram (PI. 148, fig. 1). There are probably four or five living species, and although the family characters are unmistakable, their systematics at generic and specific levels are not easy. Fossil repre- sentatives are never common, but have been found at intervals in the geological column down to the Cretaceous, the earliest British occurrence being at the base of the Ceno- manian (PI. 148, fig. 2); Stradner (in Stradner el al 1968) has recently announced the dis- covery of similar specimens in the Albianof Flolland.The interesting feature of this record is that the earliest specimens differ so little from living material: the family characters with all their eccentricities are fully developed at the first appearance, there is no clue at all about relationships to other coccolith taxa, and no suggestion of any evolutionary change during the long interval from the middle Cretaceous to the present day. LIVING FAMILIES WITH NO KNOWN GEOLOGICAL RECORD We have now considered a number of isolated species in the living plankton, many of which have turned out to be survivors of once flourishing families that are otherwise C 6055 3 G 804 PALAEONTOLOGY, VOLUME 11 extinct. In contrast with these are the two largest living families, the Syracosphaeraceae and Zygosphaeraceae, which have no fossil record behind them. Both families have more than 50 species, and together they account for more than half the living coccolitho- phorids. The Zygosphaeraceae are all holococcoliths: that is, they are constructed of minute rhombohedra or prisms of calcite, each enclosed in an organic membrane. Such structures are very delicate, and would probably break down into their component crystals soon after falling to the sea floor; this may well be the reason why they are not found in the Globigerina Ooze or other oceanic deposits. This, however, can hardly be true of the Syracosphaeraceae, which are much more robustly constructed. Many species of Syracosphaera are, indeed, found abundantly in the modern Globigerina Ooze, and their absence from the Pliocene and older deposits therefore calls for some different explanation. (W. W. Hay and his colleagues have recently referred two Tertiary species to this genus (S. bisecta Hay et al. 1966, p. 393, and S. clathrata, 1967, p. 449). S', bisecta has been re-examined by Bramlette and Wilcoxon (1967, p. 102), who regard it as a species of Coecolithus, and until more is known about the wall- structure of S. clathrata, its reference to Syracosphaera cannot be regarded as fully established.) The obvious conclusion that the Syracosphaeraceae are, in fact, post- Tertiary additions to the oceanic plankton is probably correct, and will be considered later in relation to the abrupt appearance of other cryptogenic families in earlier times. EXPLANATION OF PLATE 147 Figs. 1-4. Helicosphaera spp. 1, 2, H. carteri (Wallich) Kamptner, recent Globigerina Ooze. 1, No. 18030, Discovery II Sta. 4288, Biscay Mts., proximal view, X 7200. Also Cyclococcolithus leptoporus (M. and B.) Kamptner, distal view. 2, No. 13137 from Challenger Sta. 338, S. Atlantic Ocean; distal view, x7200. 3, H. sp.. No. 22839, Lower Oligocene, Rodstenseje nr. Odder, Denmark; proximal view, X 3200. 4, H. seminulum Bramlette and Sullivan; No. 14293, Middle Eocene, core DWBG 23B, Pacific Ocean; proximal view, X6700. Figs. 5, 7. Braarudosphaera spp. 5, B. bigelowi (Gran and Braarud) Deflandre; No. 15730, Yazoo Formation (Upper Eocene), Clarke County, Mississippi; complete coccolith of five plates, X4000. 7, B. africana Stradner; No. 16057, Sutterby Marl (Aptian), borehole Alford, Lines., 101-7 ft.; complete coccolith, X 4000. Fig. 6. Pemma papillatum Martini; No. 15871, Yazoo Formation (Upper Eocene), Clarke County, Mississippi; single plate, x 5000. EXPLANATION OF PLATE 148 Fig. 1. Calciosolenia sp., No. 13131, recent Globigerina Ooze, Challenger Sta. 338, S. Atlantic Ocean; proximal view, X 10 000. Fig. 2. Scapholithus sp., No. 20678, Chalk Marl, (Cenomanian), Folkestone, Kent; proximal view, x 13 000. Figs. 3, 4. Syracosphaera spp. 3, S. hystrica Kamptner; No. 2573, recent Globigerina Ooze, Challenger Sta. 338, S. Atlantic Ocean; distal view, x 16 000. 4, S. pulchra Lohmann; No. 18062, recent Globigerina Ooze, Discovery II Sta. 4288, Biscay Mts. ; proximal view, X 6000. Fig. 5. Loxolithus armilla (Black) Noel, holotype, No. 2807, Burwell Rock (Cenomanian), Great Shel- ford nr. Cambridge; distal view, X 8000. Figs. 6, 9. Rhabdolithina spp. 6, R. sp.. No. 18616, Sutterby Marl (Aptian), borehole, Alford, Lines., 101-7 ft.; oblique proximal view, x6000. 9, R. sp.. No. 13501, Lower Gault (Albian), Folkestone, Kent; oblique proximal view, x 8000. Fig. 7. Staurolithites sp., No. 21747, Cambridge Greensand (Cenomanian), Cherry Hinton Fields nr. Cambridge; distal view, x 8000. Fig. 8. Zygolithus diplogrammus Deflandre; No. 22337, Mooreville Chalk (Santonian), nr. Eutaw, Alabama; distal view, X5300. Palaeontology, Vol. 11 PLATE 147 BLACK, Living, Tertiary and Cretaceous coccoliths Palaeontology, Vol. 11 PLATE 148 BLACK, Living and Cretaceous coccoliths M. BLACK: TAXONOMIC PROBLEMS IN THE STUDY OF COCCOLITHS 805 EXTINCT FAMILIES In examining the fossil representatives of living families, we have not found it necessary to employ a dual classification with the extinct species segregated into para- taxa. We can approach the extinct families in much the same way, and follow the well-tried palaeobotanical practice of treating them as being made up of closely related organ-genera. Much depends upon the choice of characters that are relied upon for the definition of genera or families, and there must always be an element of personal opinion in making this selection. Quite clearly, we cannot follow any hard and fast rules, for some kinds of criteria that are of great value in certain families are worthless in others. For example, in the Zygolithaceae and Podorhabdaceae, the number of crystals in the marginal shields or rings is variable, even within a single species; in the Deflandriaceae, however, the number is always 16 (PI. 151, fig. 2), and in the Braarudosphaeraceae it is 5 throughout the whole family (PI. 147, fig. 5). In a large number of families the structure of the outer margin of the coccolith follows a characteristic pattern, and can be used as a primary basis of classification. Thus, the Zygolithaceae all have an architecture based upon the loxolith-ring (see below and PI. 148, fig. 5); the Coccolithophoraceae have two marginal shields of radial elements (PI. 143, fig. 4; PI. 144, fig. 2; PI. 145, fig. 3), and the Podorhabdaceae have yet another type of rim (PI. 150, figs. 1, 6). The presence or absence of a central spine is also of unequal significance in ditfieient families. Such structures are characteristically present in the Rhabdosphaeraceae, but never in the Coccolithophoraceae. In the Podorhabdaceae, on the other hand, the coccoliths of some genera, such as Polypodorhabdus (PI. 150, fig. 1) always have spines, but in others, such as Cretarhabdus, some coccoliths in a single species may have spines and others not (PI. 150, figs. 4, 5). This state of affairs is also known in some living genera, such as Syracosphaera. ZYGOLITHACEAE Noel Plate 148, figs. 5-9; Plate 149, figs. 1-5 A large number of Mesozoic coccoliths have an outer wall that consists of narrow, steeply inclined laths of calcite, giving a structure rather like the wall of a barrel that has been given a sharp twist, so that the staves run obliquely instead of straight up and down (PI. 148, figs. 6, 9). The resulting loxolith-ring resembles the wall of Pontosphaera in its general plan of construction, but differs in normally having no floor, and in having a much smaller number of staves, which are also coarser and more robust. The numerous genera which make up the family are defined by the various structures that bridge or partly close the opening within the ring. The name Loxolithus was proposed by Denise Noel (1965) for species consisting of an unmodified ring (PI. 148, fig. 5); many fossils preserved in this condition are in fact damaged specimens of other genera, having lost the bridge or central cross that was originally present. In Zygolitbus there is a bridge along the shorter axis of the ellipse on the proximal side (PI. 148, fig. 8). Pontilithus has a grid on the proximal side and a distal bridge (PI. 149, figs. 2, 3), and Stawolithites has a central cross (PI. 148, fig. 7). In several genera the proximal side of the loxolith-ring is closed by a floor, sometimes solid, but frequently pierced by pores (PI. 149, fig. 1). In Rhagodiscus the floor consists 806 PALAEONTOLOGY, VOLUME 11 of a mosaic of little granules, and the addition of a prominent spine gives a rhabdolith- like appearance to Rhabdolithina (PI. 148, figs. 6, 9). The earliest Zygolithaceae are found in the Lias, but there is no great variety of struc- ture until the Lower Cretaceous, when specialized forms begin to appear, leading up to the spectacular diversity of genera and species in the Albian and Cenomanian. High in the Upper Cretaceous there are species with a new type of floor-structure, in which the plates are arranged very much as in early species of Pontosphaera and Helicosphaera (compare PI. 149, fig. 4 with PI. 146, fig. 6 and PI. 147, fig. 3). These have been mentioned already in connection with the ancestry of these two genera. PODORHABDACEAE Noel Plate 150, figs. 1-6 This is an exclusively Mesozoic family which made its first appearance in the Oxfordian, when five well-differentiated genera appeared simultaneously without any known ancestors. Most of these continue throughout the upper part of the Jurassic and the Lower Cretaceous. An additional genus, Cretarhabdus, appears in the Berriasian and continues with undiminished vigour to the top of the Maestrichtian, when this genus, and with it the whole family quite suddenly became extinct (Bramlette and Martini 1964). The characters of the family may be seen in Plate 150, figs. 1-6; there is an ellip- tical shield with a narrow rim consisting of two layers of keystone-shaped crystals sur- rounding a very large central area, from which arises a tall spine. The several genera which make up the family are distinguished from each other by the structures occupying the central area (compare PI. 150, figs. 1, 2, 3, 5); in most of them the central spine is supported by four buttresses constructed of fibrous crystals. This is an architectural feature which is not confined to the Podorhabdaceae (PI. 149, fig. 6; PI. 151, fig. 3; PI. 152, fig. 1), and its taxonomic significance will be discussed on a later page. EXPLANATION OF PLATE 149 Fig. 1 . Zygolithus sp. with perforated floor; No. 22407, Cambridge Greensand (Cenomanian), Hauxton Mill near Cambridge; proximal view, x 8000. Figs. 2, 3. Pontilitlws flabellosus (Stradner), comb. nov. ( = Coccolitlws flabellosus Stradner), Cam- bridge Greensand (Cenomanian). 2, No. 22881, Cherry Flinton Fields, proximal view, X 6000. 3, No. 11510, Barrington, distal view, X 8000. Figs. 4, 5. Pontolithinci spp. 4, P. moorevillensis Black, No. 22362, Mooreville Chalk (Santonian), nr. Eutaw, Alabama; proximal view, X4800. 5, P. sp., No. 22899, Prairie Bluff Formation (Mae- strichtian), Wilcox County, Alabama; distal view, X 4800. Fig. 6. Eiffellitlms turriseiffeli (Deflandre) Reinhardt, No. 22114, Prairie Bluff Formation (Maestrich- tian), Wilcox County, Alabama; distal view with characteristically oblique spine, X4000. EXPLANATION OF PLATE 150 Fig. 1. Podorhabdus cylindratus Noel, No. 22798, Upper Oxford Clay (L. Oxfordian), Cambridge Experimental Borehole, 220 ft. 3 in. -221 ft. 3 in.; distal view, X4800. Fig. 2. Polvpodorhabdus madingteyeusis Black, No. 17413, Oxford Clay, locality as 1; distal view, x 9600. ’ Fig. 3. Ethmorhabdus gallicus Noel, No. 22799, Oxford Clay, locality as 1; distal view, X7200. Figs. 4-6. Cretarhabdus spp., Cambridge Greensand (Cenomanian) nr. Cambridge. 4, No. 21766 with spine, Hauxton Mill; distal view, X4800. 5, No. 21763 without spine. Cherry Hinton Fields; distal view, x 9600. 6, No. 22867, Cherry Hinton Fields, side view, X4800. Palaeontology, Vol. 11 PLATE 149 BLACK, Cretaceous coccoliths Palaeontology, Vol. 11 PLATE 150 BLACK, Jurassic and Cretaceous coccoliths M. BLACK: TAXONOMIC PROBLEMS IN THE STUDY OF COCCOLITHS 807 DEFLANDRIACEAE Black 1967 Plate 151, figs. 1-5 More than a century ago Sorby described a remarkable coccolith from the Chalk; it has a ring of sixteen granules from which arise four flying buttresses supporting a peculiarly constructed stalk. Bramlette and Martini (1964) have shown that Sorby’s coccolith belongs to one of several closely similar species that existed in the Upper Cretaceous, and which they united into the genus Deflandrius. The earliest species is found in the Gault (PI. 151, fig. 1); it has all the family characters fully developed, and has no known ancestors. Three distinct, but very similar species are very abundant in the Upper Cretaceous right up to the top of the Maestrichtian, when the whole family became extinct. The main differences between this family and the Podorhabdaceae are to be found in details of ultra-structure. Otherwise the two families share the same basic architectural plan : a spine supported by four buttresses arising from a granular ring. Quite recently, this structural plan has been found to be developed in a remarkably similar way among the uncalcified Haptophyceae, notably in several species of Chrysochromulina. This discovery raises the possibility that the ancestors of both the Deflcindriaceae and Podor- habdaceae may have belonged to such uncalcified stocks in which the family characters were already developed, and only needed the ability to calcify in order to make their appearance on the geological record. goniolithaceae Deflandre Plate 151, fig. 6 This is perhaps the most remarkable of the extinct families. It was founded by Deflandre (1957) upon a single species, G.fluckigeri, which he discovered in the Eocene of north-west Germany and the Oligocene of Basses Alpes. The coccolith is pentagonal in outline, with a granular centre and radially constructed border. More recently, Martini (1964) has discovered a complete cyst in the Clayton Formation (Danian) of Alabama ; in this unique specimen, twelve of the pentagonal coccoliths are fitted together to form a regular dodecahedron, giving a configuration exactly like the cysts of Braaru- dosphaera , but with an entirely different coccolith-structure. Martini has shown that G.fluckigeri ranges from the Lower Maestrichtian to the Rupelian, in the later occur- rences apparently re-worked. The only observable difference between earliest and latest autochthonous specimens is a slight increase in size. Although Goniolithus is always a rare fossil, it has a wide geographical distribution : Denmark, France, Germany, English Channel (submarine core-samples), Tunisia, Alabama, and Texas. At the time of writing, the family still includes only the single species. stephanolithiaceae fam. nov. Plate 152, figs. 1-3 The genus Stephanolitliion was created by Deflandre (1939) for a peculiar species of coccolith from the Oxford Clay of Calvados, which he named S. bigoti. It differs from other coccoliths in having a set of radial spines extending beyond the marginal ring; 808 PALAEONTOLOGY, VOLUME 11 moreover, each spine consists of a single scalenohedron, a crystal habit that is most unusual among the coccolithophorids. Two more species have subsequently been dis- covered : S', speciosum, also from the Oxford Clay, and S. laffittei from the Portlandian of Algeria. S. bigot i is known from the Callovian, but is only common in the Oxfordian; in England it appears to be confined to the Upper Oxford Clay. C. laffittei is now known to occur throughout the Cretaceous System, becoming extinct at the top of the Maestrich- tian. Stephanolithion stands so far apart from other coccoliths that it cannot be fitted into any existing family, and I propose a new family name, Stephanolithiaceae, with the characters of the single included genus. discoasteraceae Tan Sin Hok Plate 153, figs. 1-9 The discoasters are usually considered to be taxonomically distinct from the cocco- liths, but it would be unrealistic to ignore the liklihood that they are the products of similar planktonic algae, possibly in an encysted condition. They are mentioned here because they illustrate particularly clearly the taxonomic importance of crystal-orienta- tion, which these organisms were able to control with great precision. In most species of Discoaster no crystal outlines are visible; nevertheless, the orientation of the crystals can be determined optically, and they are invariably found to be arranged so that the c-axis lies parallel with the axis of symmetry of the discoaster. In a smear-slide prepared so that the fossils lie flat against the cover-slip, they consequently remain dark in all positions between crossed nicols, and show the maximum refractive index for calcite. This property is constant for all discoasters, and at once distinguishes them from typical coccoliths. The general appearance of the discoasters can be seen in Plate 153, figs. 1-8. They EXPLANATION OF PLATE 151 Figs. 1-5. Deflandrius spp. 1, D. cantabrigensis Black, No. 22863, Cambridge Greensand (Cenomanian), Cherry Hinton Fields nr. Cambridge; side view, X 6000. 2, 3, D. spinosus Bramlette and Martini. 2, No. 22187, Prairie Bluff Formation (Maestrichtian), Wilcox County, Alabama; oblique distal view, x 4000. 3, No. 20144, Upper Chalk (Campanian), borehole at Holton nr. Halesworth, Suffolk; distal view, X 8000. 4, 5, D. cretaceus (Arkhangelski) Bramlette and Martini. 4, No. 18481, Upper Chalk (Santonian), nr. Salisbury, Wilts.; oblique proximal view of short-stalked form, x 6000. 5, No. 22321, Mooreville Chalk (Santonian). nr. Eutaw, Alabama; oblique proximal view, X 6000. Fig. 6. Goniolithiis fluckigeri Deflandre, No. 22888, Prairie Bluff Formation (Maestrichtian), Wilcox County, Alabama; damaged coccolith, X 10 000. EXPLANATION OF PLATE 152 Figs. 1-3. Stephanolithion spp. 1, S. bigoti Deflandre, No. 22783, Upper Oxford Clay (L. Oxfordian), Cambridge Experimental Borehole; distal view, X4000. 2, 3, S. laffittei Noel, Lower Gault (M. Albian), Folkestone, Kent. 2, No. 13503, proximal view, X 10 000. 3, No. 13537, oblique distal view, x 8000. Fig. 4. Neococcolithes sp.. No. 22923, Prairie Bluff Formation (Maestrichtian), Wilcox County, Alabama; proximal view, X 10 000. Figs. 5-7. Kamptnerius magnificus Deflandre, Prairie Bluff Formation (Maestrichtian), Wilcox County, Alabama, X 4000. 5, No. 22902, form with narrow wing, distal view. 6, No. 22928, form with expanded wing, distal view. 7, No. 22925, form with narrow wing, proximal view. Palaeontology, Vol. 11 PLATE 151 BLACK, Cretaceous coccoliths Palaeontology, Vol. 11 PLATE 152 BLACK, Jurassic and Cretaceous coccoliths M. BLACK: TAXONOMIC PROBLEMS IN THE STUDY OF COCCOLITHS 809 consist of calcite crystals grouped round a central axis, giving a snow-flake pattern. In Marthasterites (PI. 153, fig. 9) the whole structure appears to consist of a single calcite crystal, fn Discoaster each ray is crystallographically distinct, as may be seen from the etch-patterns which are all too common on corroded specimens, or from the crystal faces that are sometimes seen in exceptionally well-preserved specimens (PI. 153, fig. 1). Marthasterites is found in the Upper Cretaceous and in the Lower Eocene; Dis- coaster is confined to the Tertiary, appearing in the Paleocene and becoming extinct at the top of the Pliocene or just within the Pleistocene, according to the horizon adopted for the base of the Quaternary. Reported discoveries of living Discoasteraceae are at present unconfirmed. Constancy of crystal-arrangement is equally prevalent among the coccoliths; its optical effects, however, are much more complicated when the crystal-axes are not all parallel, but they are none the less of great taxonomic significance, and are relied upon extensively in systematic work. In the Braarudosphaeraceae the crystals are arranged round a central axis, much as in the Discoasteraceae, but the orientation is quite dif- ferent, with the rhombohedron faces lying approximately in the plane of the coccolith (PI. 147, fig. 5). In Stephanolithion the radial spines are scalenohedra with the optic axes at right angles to the central axis of the coccolith (PI. 152, fig. 1). In the majority of coccoliths, the arrangement is much more complicated, with the optic axes arranged in a pattern that almost invariably involves some kind of screw symmetry. RELATIONS WITH UNCALCIFIED STOCKS The geological histories of the various coccolithophorid stocks discussed above clearly do not all follow the same pattern. On the one hand there are long-ranging stocks such as the Coccolithophoraceae, Pontosphaeraceae, and Zygolithaceae, with a great number of species spread out through a considerable span of geological time. These pass through phases when generic and specific distinctions are blurred, and evolutionary divergence was evidently taking place. The Coccolithophoraceae, for example, passed through such phases in Albian-Cenomanian times, again in the Eocene, and apparently once more in the Pleistocene. The taxonomic problems involved in this state of affairs are those familiar in any actively evolving group of organisms, and call for no special comment here. In contrast with such normally evolving stocks, there are others which behave quite differently; these arrive on the scene abruptly, with all their characters ready-made. Some of them, for example Goniolithus, Deflandrius , and Stephanolithion , persist perhaps for the length of a geological period with little or no change, and then just as abruptly become extinct; others, such as the Podorhabdaceae, appear equally abruptly, but undergo noticeable evolutionary changes before they become extinct. In each of these cryptogenic stocks, there is a well-defined architectural plan which we have taken as characteristic of the family, and which is fully perfected in the earliest known species. It looks very much as though the architectural plan had already been elaborated before the fossil record started, and I think we have to consider seriously the possibility that the beginning of the fossil record for such a family simply marks the change-over to calcifi- cation in a stock whose scales were previously unmineralized. Of course, we can hardly hope to find any tangible evidence of a palaeontological 810 PALAEONTOLOGY, VOLUME 11 nature to support this speculation, but fortunately there are other lines of investigation open to us. The most promising of these is an inquiry into the existing unmineralized stocks themselves, with a view to finding out what linkages may exist between these and the coccolith-bearing algae. Recent work on the uncalcified Haptophyceae has indeed brought to light some very interesting parallels with the coccolithophorids. The various species of Chrysochromulina have been found to bear delicate organic scales, which in their architectural construction are very similar to some coccoliths. It is a little surprising to find that the most elaborately constructed scales, although quite unlike the majority of living forms, can be readily matched among extinct taxa. In spite of their extreme tenuity, these scales apparently consist of two layers which are differently sculptured, and if calcified would produce entirely different crystal patterns. The simplest scales are elliptical discs, with sharply etched radiating lines on the proximal surface, and a less clearly defined series of concentric rings o,r a vague spider’s web pattern on the distal side (PI. 154, figs. 1, 2). Scales of this kind are also well-known in Crystallolitlms hyalinus , where they provide the support upon which calcite crystals are secreted (Parke and Adams 1960). In addition to these simple scales, most species of Chrysochromulina also carry larger and more elaborate scales, in which the distal layer is modified in various ways, often producing a central spine. The large scales of C. chiton (Parke et al. 1958, figs. 16, 17, 28, 29) are shaped like pie-dishes, with sloping walls on the distal surface, much like those of Pontosphaera (PI. 146, fig. 5). C. ericina (Parke et al. 1956, figs. 17-19) has scales shaped like rhabdoliths with very long spines, reminiscent of some Tertiary species of Rhabdosphaera. In several other species the spines are differently constructed, and do not resemble those of Rhabdosphaera , but are EXPLANATION OF PLATE 153 Figs. 1-8. Discoaster spp. 1, D. sp. showing crystal faces; each of the six rays is a separate rhombo- hedron, with triad axis at right-angles to page; No. 22291, Upper Oligocene, core DWBG 10, Pacific Ocean ; X6100. 2, D. elegans Bramlette and Sullivan, No. 11890, Middle Oligocene (reworked from Eocene?), Grundfor, Denmark; X 3000. 3, D. barbadiensis Tan Sin Hok, No. 11949, Middle Oligocene (reworked from Eocene?), Grundfor, Denmark; X4500. 4, D. lodoensis Bramlette and Riedel, No. 15771, Eocene, Tuilerie de Donzacq, Landes, France; X 1900. 5, D. strictus Stradner, No. 11899, Middle Oligocene (reworked from Eocene?), Grundfor, Denmark; x 3000. 6, D. deflcmdrei Bramlette and Riedel, No. 15830, Eocene, Tuilerie de Donzacq, Landes, France; x 4000. 7, D. hamatus Martini and Bramlette, No. 16969, damaged specimen. Pliocene, core DWBP 119, Pacific Ocean; : 4000. 8, D. brouweri Tan Sin Hok, No. 22197, Pliocene, core LSDH 78P, Pacific Ocean ; x 3000. Fig. 9. Marthasterites tribrachiatus (Bramlette and Riedel) Deflandre, No. 10485, Lower Eocene, Rosnaes, Denmark; X2500. EXPLANATION OF PLATE 154 Figs. 1-3. Crystcillolithus hyalinus Gaarder and Markali (= motile phase of Coccolithus pelagicus ), strain No. LB 913/12 (Plymouth 182), Culture Collection of Algae and Protozoa, Botany School, Cambridge; x 16 000. 1, No. 22909, two scales in proximal view. 2, No. 22911, two scales in distal view. 3, No. 22915, scale and calcite rhombohedra detached during preparation. Fig. 4. Scyphosphaera sp., No. 18028, ordinary coccolith, proximal view; recent Globigerina Ooze, Discovery II Sta. 4269, Biscay Mts.; x 6000. Figs. 5-7. Scyphosphaera sp., ordinary coccoliths; modern Globigerina Ooze, Challenger Sta. 338, S. Atlantic Ocean. 5, No. 11316, distal view, x 6000. 6, No. 11317, detail, distal surface, x 20 000. 7, No. 11308, detail, proximal surface, x 80 000. Palaeontology, Vol. 11 PLATE 153 BLACK, Tertiary discoasters Palaeontology, Vol. 11 PLATE 154 I 3 BLACK, Living uncalcified Haptophyceae and coccoliths M. BLACK: TAXONOMIC PROBLEMS IN THE STUDY OF COCCOLITHS 811 remarkably like those of certain Mesozoic genera. In these species the spine is supported by four buttresses which arise from the distal surface of the scale. C. alifera (Parke et a/. 1956, fig. 76) has buttresses which do not reach the marginal rim, much as in the Cretaceous Eijfellithus (PI. 149, fig. 6). In C. ephippium (Parke et a/. 1956, figs. 38, 39) the buttresses are arranged exactly as in the Jurassic Polypodorhabdus (PI. 150, fig. 2); they extend from the marginal rim to the centre, dividing the scale into four quadrants, each of which is ornamented by radial striae on the proximal surface. In the large spined scales of C. pringsheimii (Parke et al. 1962, figs. 21, 22, 29) the supports of the tall spine are no longer in contact with the surface of the scale except at the outer margin, but behave like flying buttresses; the whole arrangement is an astonishing duplication of the Upper Cretaceous Deflandrius (PI. 151, fig. 3). C. pringsheimii is unusual in having so many as four different kinds of scale, each of which could serve as a template for the orderly disposition of calcite-crystals to give a familiar coccolith pattern. Enough has been said to show that there is a very significant parallel between the calcified and uncalcified Haptophyceae. Several of the living species of Chrysochromu- lina could indeed be regarded as uncalcified survivors of familiar Mesozoic families. It would be asking too much of coincidence to believe that the few living species of Chrysoehromulina have contrived to reconstruct in their uncalcified scales just those features that were required to start off the Podorhabdaceae in the Oxfordian, the Deflandriaceae in the Albian, and the Pontosphaeraceae and Helicosphaeraceae some- time near the beginning of the Tertiary. 1 think it is at least as reasonable to suppose that this remarkable genus is indeed a survival from a protean stock that existed in early Mesozoic times, and by selective calcification of its organic scales has given birth to these ancient families. REFERENCES black, m. 1968. The systematics of coccoliths in relation to the geological record. In The micro- palaeontology of oceans , edited by B. M. Funnell and W. R. Riedel, Cambridge University Press. bramlette, m. n. and martini, e. 1964. The great change in calcareous nannoplankton fossils between the Maestrichtian and the Danian. Micropaleontology , 10, 291-322. and sullivan, f. r. 1961. Coccolithophorids and related nannoplankton of the early Tertiary in California. Ibid. 7, 129-88. and wilcoxon, j. a. 1967. Middle Tertiary calcareous nannoplankton of the Cipero section, Trinidad, W.I. Tulane Stud. Geol. 5, 93-131. deflandre, g. 1939. Les stephanolithes, representants d'un type nouveau de coccolithes du Jurassique superieur. C. r. hebd. Seanc. Acad. Sci. Paris , 208, 1331-3. 1947. Braarudosphaera nov. gen., type d’une famille nouvelle de coccolithophorides actuels a elements composites. Ibid. 225, 439-441. 1952. Coccolithophoridae. In Traite de Zoologie, ed. P. P. Grasse. Paris: Masson. 1959. Sur les nannofossiles calcaires et leur systematique. Rev. Micropaleont. 2, 127-52. 1966. Commentaire sur la systematique et la nomenclature des nannofossiles calcaires. Cahiers Micropaleont., ser. 1, no. 3, C.R.N.S., Paris. 9 pp. and fert, c. 1954. Observations sur les coccolithophorides actuels et fossiles en microscopie ordinaire et electronique. Ann/s. Paleont. 40, 1 15-76. gaarder, k. r. 1967. Observations on the genus Ophiaster Gran (Coccolithineae). Sarsia, 29, 183-92. gran, h. h. and braarud, t. 1935. A quantitative study of the phytoplankton in the Bay of Fundy and the Gulf of Maine. J. biol. Bd. Can. 1, 297-467. halldal, p. and markali, j. 1955. Electron microscope studies on coccolithophorids from the 812 PALAEONTOLOGY, VOLUME 11 Norwegian Sea, the Gulf Stream and the Mediterranean. Avh. norske VidenskAkad. Oslo, 1955 (1), 30 pp. hay, w. w., mohler, h. p., roth, p. h., schmidt, r. r., and Boudreaux, j. e. 1967. Calcareous nanno- plankton zonationof the Cenozoic of the Gulf Coast and Caribbean-Antillean area and transoceanic correlation. Trans. Gulf-Cst Ass. geol. Socs. 17, 428-80. and wade, mary e. 1966. Calcareous nannofossils from Nal’chik (Northwest Caucasus). Eel. Geol. Helv. 59, 379-400. kamptner, e. 1958. Betrachtungen zur Systematik der Kalkflagellaten, nebst Versuch einer neuen Gruppierung der Chrysomonadales. Arch. Protistenk. 103, 54-116. lohmann, h. 1902. Die Coccolithophoridae, eine Monographic der Coccolithen bildenden Flagellaten. Ibid. 1, 89-165. mcintyre, a. and be, a. w. h. 1967. Modern Coccolithophoridae of the Atlantic Ocean — I. Placoliths and Cyrtoliths. Deep Sea Res. 14, 561-97. martini, e. 1964. Ein vollstandiges Gehause von Goniolitluis fluckigeri Deflandre. Neues Jb. Geol. Paldont. Abh. 119, 19-21. Murray, g. and blackman, v. h. 1898. On the nature of the coccospheres and rhabdospheres. Phil. Trans. R. Soc., B, 190, 427-41. noel, denise. 1965. Sur les coccolithes Jurassiques Europeens et d'Afrique du Nord. C.N.R.S., Paris. 209 pp. paasche, e. 1964. A tracer study of the inorganic carbon uptake during coccolith formation and photosynthesis in the coccolithophorid Coccolithus huxleyi. Physio/ogia PL, suppl. 3, 82 pp. parke, mary. 1961. Some remarks concerning the class Chrysophyceae. Br. phycol. Ball. 2, 47-55. and adams, Irene. 1960. The motile ( Crystallolithus hyalinus Gaarder and Markali) and non- motile phases in the life history of Coccolithus pelagicus (Wallich) Schiller. J. mar. biol. Ass. U.K. 39, 263-74. — and dixon, p. s. 1964. A revised check-list of British marine algae. Ibid. 44, 499-542. manton, Irene, and clarke, b. 1956-62. Studies on marine flagellates. Ibid. 35, 387-414; 37, 209-28; 42, 391-404. schiller, j. 1930. Coccolithineae. In Rabenhorst’s Kryptogamen-Flora, 10 (2), 89-266. stradner, h„ adamiker, d., and maresch, o. 1968. Electron microscope studies on Albian calcareous nannoplankton. Verb. K. ned. Akad. Wet. 24 (4). 107 pp. watabe, n. and wilbur, k. m. 1966. Effects of temperature on growth, calcification and coccolith form in Coccolithus huxleyi (Coccolithineae). Limnol. Oceanogr. 11, 567-75. wilbur, k. m. and watabe, n. 1963. Experimental studies on calcification in molluscs and the alga Coccolithus huxleyi. Ann. N.Y. Acad. Sci. 109 (1), 82-112. APPENDIX Localities of deep-sea samples mentioned in the explanation of plates H.M.S. Challenger Station 338: dredge, 21° 15' S., 14° 02' W., about 900 miles S. of Ascension Is., South Atlantic Ocean. R.R.S. Discovery II Station 3809: dredge, northern slope of Galicia Bank, 42° 56' N., 11° 47' W., about 120 miles W. of Cape Finisterre. Station 4269: trigger-weight core, Biscay Mts., W. of Bay of Biscay. Station 4288 : dredge, Biscay Abyssal Plain, Bay of Biscay. Scripps Institution of Oceanography DWBG 10: core sample, specimens from 11 to 27 cm. 6° 54' N., 131° 00' W., central Pacific Ocean. DWBG 23B: core sample, specimens from 10 to 12 cm. 16° 42' S., 145° 48' W., between Tuamotu Archipelago and Tahiti. M. BLACK: TAXONOMIC PROBLEMS IN THE STUDY OF COCCOLITHS 813 DWBP 119: core sample, specimens from 398 to 400 cm. 27° 54' S., 106° 53' W., about 200 miles ESE. of Easter Is. LSDH 78PG core sample, specimens from 250 to 270 cm. 4° 30' S., 168° 03' E., between Gilbert Is. and Solomon Is. MAURICE BLACK Department of Geology Sedgwick Museum Typescript received 21 May 1968 Cambridge THE PALAEONTOLOGICAL ASSOCIATION Extracts from the Annual Report of the Council for 1967-8 Membership. On 31 December 1967 there were 1,353 members (725 Ordinary, 131 Student, and 497 Institutional), a net increase of 15 members during the year. Finance. During 1967 expenditure rose by £2,724 due largely to the cost of reprinting Volume 3, Part 2 and producing Special Papers. This increase was more than offset by an increase in sales from £1,381 to £4,348. There was also an increased income from subscriptions, interest, and donations to specific papers, but general donations at £525 reached their lowest level. Special Paper No. 1 was published at a cost of £3,227 ; donations of £1,500 from the N.C.B. and £250 from Shell Oil Company together with sales of £912 left £565 as a charge on the Publications Reserve at the end of the year. Standing orders for Special Papers now number 135. Volume 3, Parts 1, 3, and 4 will have to be reprinted as soon as possible and £1,500 has been set aside for this purpose from the Publications Reserve. This now stands at £5,720, enough to cover 62 per cent of a single volume compared with 63 per cent at the end of 1966. Subscriptions for Ordinary and Student Members were raised to meet the increasing cost of printing Palaeontology which will rise to over £9,000 in 1968. Publications. The four parts of Volume 10 of Palaeontology were published during 1967; they con- tained 37 papers plus ‘Notes for Authors’. Special Papers in Palaeontology No. 1 for 1967 was published; it contained 324 pages and 27 plates. Meetings. Four meetings took place during 1967-8. The Association is grateful to the Council of the Geological Society of Fondon, Professor J. Sutton (Imperial College, Fondon), and Professor Scott Simpson (University of Exeter) for generously granting facilities for meetings, and to Mr. G. W. Green and Mr. M. Mitchell for leading the field meeting. a. The Tenth Annua! Genera! Meeting was held in the Rooms of the Geological Society of London, Burlington House, London, W. 1, on Wednesday, 1 March 1967, at 5.00 p.m. The Annual Report of the Council for 1966-7 was adopted and the Council for 1967-8 was elected. Professor Alwyn Williams of the Queen’s University, Belfast, delivered the Tenth Annual Address on ‘Evolution of the shell structure of articulate brachiopods’. b. A Field Demonstration Meeting was held at Burrington Combe in the Mendips on Saturday, 6 May 1967. Mr. G. W. Green and Mr. M. Mitchell demonstrated the Carboniferous Limestone succession. c. The Tenth Anniversary Meeting and Dinner was held at Imperial College, London, on Wednesday, 8 November 1967. The discussion meeting was on ‘The explosive development of palaeontology’ and the introductory speakers were Professor P. C. Sylvester-Bradley, Dr. D. N. Wood, and Dr. W. S. McKerrow. d. A Symposium on ‘Devonian floras and faunas’ was held in the Department of Geology, Univer- sity of Exeter, on Thursday to Saturday, 14-16 December 1967. About 45 persons attended to hear 16 papers and see 13 demonstrations. Professor Scott Simpson was local secretary. Special General Meeting. Immediately before the Discussion Meeting at Imperial College, London, on Wednesday, 8 November 1 967, a Special General Meeting was held at 3.00 p.m. to consider a change in Rule 2 of the Constitution. The following motion was proposed from the Chair and passed (nem. con.): That, effective from 1 January 1968, Rule 2 of the Constitution be amended to read as follows: 'There shall be Ordinary members. Institutional members, and Student members, and their annual subscriptions shall be Five Pounds, Seven Pounds, and Three Pounds respectively. Each subscriber shall be considered a member of the Association, but Institutional members shall not be eligible to take part in the government of the Association.’ THE PALAEONTOLOGICAL ASSOCIATION 815 Council. The following were elected members of the Council of the Association for 1967-8 at the Annual General Meeting on 1 March 1967: President : Professor T. S. Westoll, F.R.S.; Vice-Presidents : Dr. W. S. McKerrow, Professor F. H. T. Rhodes, Professor C. H. Holland; Treasurer : Dr. C. Downie; Secretary: Dr. J. M. Hancock; Editors: Mr. N. F. Hughes, Dr. Gwyn Thomas, Dr. I. Strachan, Professor M. R. House, Dr. R. Goldring; Other members: Mr. M. A. Calver, Dr. C. B. Cox, Mr. D. Curry, Dr. Grace Dunlop, Mr. G. F. Elliott, Dr. T. D. Ford, Dr. A. Hallam, Dr. R. P. S. Jefferies, Dr. G. A. L. Johnson, Dr. W. D. I. Rolfe, Dr. A. H. Smout, Dr. L. B. H. Tarlo, Professor H. B. Whittington. BALANCE SHEET AND ACCOUNTS FOR THE YEAR ENDING 31 DECEMBER 1967 Balance Sheet Liabilities Publications Reserve Account Balance as per Account ...... Special Papers Fund Account ....... Provision for Reprinting ........ Amounts received in advance of Subscriptions for 1968 Provision for cost of publication of Palaeontology , Vol. 10, as per Income and Expenditure Account ...... Less Expenditure incurred to 31 December 1967 Sundry Creditors ......... £ s. d. £ s. d. 5,720 1 8 1,500 0 0 451 10 8 7,960 8 10 4,860 8 10 3,100 0 0 25 5 1 Note. No amount has been included in these accounts for subscriptions unpaid at 31 December 1967 £10,796 17 5 Assets Office Equipment at cost ..... Less Depreciation to date .... Investments at cost : Equities Investment Fund for Charities — 958 units 5 per cent Defence Bonds .... Liverpool 6| per cent Bonds Wagon Investment Loan 6f per cent 5i per cent National Development Bonds Sundry Debtors and Payments in Advance: Authors for offprints ..... Advance Payments: For Palaeontology, Vol. 11. For Palaeontology, offprints .... Cash at Bank: Deposit Account ..... Current Account ..... £ s. d. 27 6 6 15 6 6 999 18 3 2,000 0 0 1,500 0 0 2,000 0 0 2,000 0 0 83 6 8 84 16 6 106 17 6 1,462 15 7 £ s. d. 12 0 0 8,499 18 3 547 211 168 3 2 1,569 13 1 £10,796 17 5 Report of the Auditors to the Members of the Palaeontological Association. We have examined the above Balance Sheet and annexed Income and Expenditure Account which in our opinion give respectively a true and fair view of the state of the Association’s affairs as at 31 December 1967 and of its income and expenditure for the year ended on that date. JOSHUA WORTLEY & CO. Chartered Accountants 816 PALAEONTOLOGY, VOLUME 11 Income and Expenditure Account for the Year ended 31 December 1967 Expenditure £ s. d. £ s. d. To Provision for the cost of Publication of Palaeontology, Vol. 10 Part 1 ...... 1,925 10 3 Part 2 ...... 2,034 18 7 Part 3 2,000 0 0 Part 4 ...... 2,000 0 0 7,960 8 10 Add extra costs Vol. 9, Part 4 146 3 4 8,106 12 2 Less over-provision for Vol. 9, Part 3 118 16 9 7,987 15 5 Reprinting Vol. 3, Part 2 554 8 6 Administrative Expenses: Insurance ..... 2 0 3 Postages and Stationery 191 13 6 Audit Fee ..... 14 14 0 Honorarium to Assistant Treasurer . 120 0 0 Duplicating and dispatch of circulars . 181 10 2 Meeting expenses .... 32 9 3 542 7 2 Depreciation of Equipment . 6 0 0 Special Papers: Printing costs ..... 3,319 7 9 Less Sales ..... . 920 9 11 Donations: Shell . 250 0 0 Foreman ..... . 475 8 9 1,645 18 8 1,673 9 1 Excess of Income over Expenditure transferred to Publications Reserve Account ........ 1,578 13 9 £12,342 13 11 THE PALAEONTOLOGICAL ASSOCIATION 817 Income By Subscriptions for 1967 .... Subscriptions for 1966 ..... Sales of Publications ..... Interest received : 5 per cent Defence Bonds .... Bank Deposit Account .... Equities Investment Fund for Charities Liverpool Corporation .... Wagon Investments ..... Special Donations ..... General Donations: British Petroleum Co. .... Texaco ....... Burmah Oil Co., Ltd. .... Miscellaneous Receipts .... Special Papers Fund ..... Excess of Expenditure over Income transferred Reserve Account ..... £ s . d. £ s. d. 5,923 11 8 202 0 3 6,125 11 11 3,428 15 2 260 0 0 30 18 2 76 4 10 101 9 0 124 18 4 593 10 4 217 14 7 250 0 0 175 0 0 100 0 0 525 0 0 66 11 11 1,385 10 0 to Publications £12,342 13 11 Publications Reserve Account To Balance as per Balance Sheet .... 5,720 1 8 Excess of Expenditure over Income . - - Provision for Reprinting ..... . - - £5,720 1 8 By Balance at 31 December 1966 .... 5,141 7 11 Add Excess of Income over Expenditure 1,578 13 9 Less Provision for Reprinting .... 1,000 0 0 578 13 9 £5,720 1 8 INDEX Pages 1-162 are contained in Part 1; pages 163-327 in part 2; pages 329-490 in Part 3; pages 491-642 in Part 4; pages 643-824 in Part 5. Figures in Bold Type indicate plate numbers. A Acltlysopsis, 218; hemitora ; 219, 42; liokata, 219, 42. Africa: non-marine ostracods from Ghana, 295. Albertella, 212; eiloitys, 213, 39; judithi, 212, 39; lata, 214, 39. Albertelloides, 214; mischi, 215, 38; pandispinata, 216, 39; sp., 217, 39. Alectryonia, 482. Algae: Cretaceous coccoliths, 361; taxonomy of coc- coliths, 793; Tethyan Dasycladaceae, 491. Allen, N. W. See Kilenyi, T. I. Alokistocare alta, 226, 41. Alveolites, 49; multiperforatus, 50, 13; obtortus, 51, 17 ; suborbicularis, 49, 12, 14, 15 ; s. forma gemmans, 50, 16. Amecephalus laticaudum, 227, 40. Ammobaculites sp., 156, 30. Ammonoidea: Famennian ammonoids, 535; new Kimmeridgian ammonite, 16; Visean/Namurian goniatites, 264. Andrews, H. N. See Phillips, T. L. Angiosperm: probable pollen in British Cretaceous, 421. Antagmus arenosus, 220, 36. Aquilapollenites, 549; attenuatus, 550, 106; pachypolus, 551, 106; spinulosus, 550, 106. Archangelsky, S. Studies on Triassic fossil plants from Argentina. IV. The leaf genus Dicroidium and its possible relation to Rhexoxylon stems, 500. Archengraulis, 255 ; productus, 256. Arctostrea, 458; alaefonnis, 87; colubrina, 85, 86, 88; c. ricordeana, 459, 85-87; diluviana, 463, 85; pusilla, 87, 90; ungulata, 463, 85-87. Argentina: leaf genus Dicroidium, 500; trunk of Rhexoxylon, 236. Armstrong, J. The unusual brachial skeleton of Atten- uatella convexa sp. nov. (Brachiopoda), 783. Arthropoda. See Ostracoda, Trilobita. Assam : Eocene Nummulites from, 669. Astartila cytherea, 20. Asterocyclina asterisca, 287. Astrocystites, 513; distans, 515, 99, 100. Athabaskia sp., 203, 43. Attenuatella, 786; convexa, 788, 142; cf. incurvata, 791, 142; sp. A, 790, 142. Aulopora, 60; ? tubaeformis, 61, 17. Auroraspora macro, 124, 27. Australia: Carboniferous brachiopods, 389; Devonian ammonoids, 535; Devonian brachiopods, 627, 731 ; Devonian foraminifera, 601 ; Ordovician edrio- blastoid, 513; Permian bivalve ligaments, 94; Permian brachiopods, 783; Silurian trilobite, 691. Australirhynchia, 731 ; cudalensis, 732, 141. Austrospirifer variabilis, 64. Ayalaina rntteni, 103. B Bactrynium, 329; bicarinatum, 66-68. Baculatisporites fusticulus, 117, 25. Bajanorthis tukolandica, 91. Banner, F. T. See Eames, F. E. Barringtonella bulmani, 149. Bates, D. E. On ‘ Dendrocrinus ’ cambriensis Flicks, the earliest known crinoid, 406. Bathyuriscus pe talus, 203, 43. Batten, D. J. Probable dispersed spores of Cretaceous Equisetites, 633. Batten, R. L. See Rollins, H. B. Biscalitheca musata, 105, 21-24. Bivalvia: epizoic oysters, 19; functional study of oyster, 458; ligaments in Permian, 94; micro- structure of shell, 163. Black, M. Taxonomic problems in the study of cocco- liths, 793. Blake, D. B. Pedicellariae of two Silurian echinoids from western England, 576. Blow, W. FI. See Eames, F. E. Bobinella kulumbensis, 91, 92. Bonnia copia, 194, 36. Boucot, A. J. See Walmsley, V. G. Boulter, M. C. A species of compressed lycopod sporophyll from the Upper Coal Measures of Somerset, 445. Braarudosphaera, 802; bigelowi, 802, 147; africana, 147. Brachial skeleton: unusual in Attenuatella, 783. Brachiopoda: Carboniferous schizophoriids, 64, 389; Dayia navicula, 612; delthyrial cover in Mucrospi- rifer, 317; feeding mechanisms, 329; mantle canal patterns, 389; new resserellid genus, 306; new rhynchonelloid genus, 731; septate dalmanellid, 627 ; shell structure, 486 ; unusual brachial skeleton, 783. Brett, D. W. Studies on Triassic fossil plants from Argentina. III. The trunk of Rhexoxylon, 236. Bulman, O. M. B., and Rickards, R. B. Some new diplograptids from the Llandovery of Britain and Scandinavia, 1. Bythicheilus? sp. indet., 229, 42. C Caborcella, 221; clinolimbata, 221, 41; granosa, 221, 39; sp., 222, 43. Calciosolenia sp., 803, 148. 820 INDEX Callocystites jewetti, 136. Cambrian: brachiopod shell structure, 486; revision of two trilobites, 410; trilobites from Nevada, 183. Cantheliophorus spp., 454. Carboniferous: fern, 104; goniatites, 264; lycopod Eskdalia, 439; lycopod sporophyll, 445; mantle canal patterns, 389; schizophoriid brachiopods, 64; Tournaisian spore flora, 116. Carter, R. M. Functional studies on the Cretaceous oyster, Arctostrea, 458. Caryocrinites; ornatus, 138, 139; roemeri, 139; sep- tentrionalis, 139. Cassigerinella, 368; chipolensis, 369; eocaenica, 368. Central America: Tertiary larger foraminifera from, 283. Chaetetes, 46; lonsdalei, 47, 14; multitabulatus, 47, 9; cf. rotundus, 48, 9. Chanda vemista, 230, 40. Cheirocrinus, 593; anatiformis, 135; giganteus, 134; granulatus, 134, 135; jamesii, 140; languedodanus, 134; spp., 134, 135. Chlamys asperrimus, 87. Chubbina, 527; cardenasensis, 529; jamaicensis, 527, 101, 102; macgillavryi, 529, 102, 103. Clarke, W. J. See Eames, F. E. Classopollis sp., 106. Clavatipollenites, 422; hughesii, 426, 80; rotundus, 424, 79, 80. Cleidogonia antiqua, 256. Climacograptus scalaris, 10. Coccoliths: Cretaceous from Zululand, 361; taxo- nomic problems in, 793. Coccolithus, 362; cf. cavus, 143; cribosphaereUa, 362, 70; huxleyi, 799, 145; cf. marismontium, 143; pelagicus, 796, 143; zuluensium, 363, 69. Cocks, L. R. M. See Ziegler, A. M. Coelenterata : Devonian tabulate corals, 44; Silurian nredusoid (?), 610. Coenites, 51 ; escharoides, 52, 10, 13; gradatus, 53, 17; sp., 54, 14. Colour markings: in Devonian trilobites, 498. Conkin, B. M. See Conkin, J. E. Conkin, J. E., and Conkin, B. M. A revision of some Upper Devonian foraminifera from Western Australia, 601. Convolutispora, 122; cf. mellita, 122, 26; cf. tuberosa, 123, 26. Cope, J. C. W. Epizoic oysters on Kimmeridgian ammonites, 19. — Propectinatites, a new Lower Kimmeridgian am- monite genus, 16. Corals : Devonian tabulates, 44. Cordey, W. G. A new Eocene Cassigerinella from Florida, 368. — Morphology and phylogeny of Orbulinoides beck- mannii (Saito 1962), 371. Corynexochides prolatus, 204, 42. Cowen, R. A new type of delthyrial cover in the Devonian brachiopod Mucrospirifer, 317. Crassifimbria walcotti ?, 223, 37. Crassostrea, 479; ameghinoi rocana, 473; echinata, 88. Cravenoceras leion, 270, 47. Cretaceous: alveolinid, 526; angiosperm pollen, 421; coccolithophorids, 361; Equisetites spores, 633; functional study of oyster, 458; non-marine ostra- cods, 259; Wealden marine-brackish bands, 141. Cretarhabdus spp., 150. Crinoids: earliest known, 406; function of stem, 275. Crystallolithus hyalinus, 154. Cyclagelosphaera margereli, 144. Cyclococcolithus leptoporus, 797, 144. Cyclocypris? sp. A, 261, 45. Cyclolithus zulua, 363, 70. Cymaclymenia pseudogoniatites, 538. Cypridea, 158, 260; marina, 29; pumila, 29; sp. A, 260, 45; sp. B, 260, 45; sp. C, 261, 45; tuberculata, 29. ‘ Cypris ’ henfieldensis, 144, 29. Cystoids: dichoporite pore-structure, 697; Macro- cystella, 580. D Darwinula, 144; leguminella, 144, 29; oblonga, 145, 29. Davidsonella sinuata, 68. Dayia, 613; navicula, 614, 118-21; /;. cf. bohemica, 120; sp., 119. Deflandrius, 807; cantabrigensis, 151; cretaceus, 151; spinosus, 151. Denckmannites, 691 ; rutherfordi, 692, 133. ‘ Dendrocrinus' cambriensis, 406. Devon: tabulate corals from, 44. Devonian, colour markings in trilobites, 498; del- thyrial cover, 317; Famennian ammonoids, 535; foraminifera from Australia, 601; new Lower O.R.S. plant, 683; new rhynchonelloid, 731 ; septate dalmanellid, 627 ; sinus-bearing monoplacophoran, 132; tabulate corals, 44. Dicroidium, 501; coriaceum, 506, 97, 98; elongation, 503, 97, 98; sp., 508, 97; zuberi, 502, 98. Discoaster, 808; barbadiensis, 153; brouweri, 153; deflaiulrei, 153; elegans, 153; hamatus, 153; lodoen- sis, 153; strictus, 153; sp., 153. Discohelix, 554; ( D .) albinatiensis, 566, 107, 108; cf. calculiformis, 571, 110; centricosta, 569, 109; conica, 569, 108; costata, 570, 109; cotswoldiae, 572, 110; cf. crenulata, 568, 108; dictyota, 565, 107; cf. gum- beli, 573, 110; levis, 570, 109; ( Pentagonodiscus ) angusta, 573, 110; reussii, 574, 110. Discolithus, 364; cristallinus, 364, 69, 70; rliabdo- sphaericus, 364, 69, 71; spiralis, 365, 70, 71. Drozdzewski, G. See Seilacher, A. Duodecimedusina palmer i, 610. E Eames, F. E. Sindulites, a new genus of the Nummuli- tidae (Foraminifera), 435. — Clarke, W. J., Banner, F. T., Smout, A. H., and Blow, W. FI. Some larger foraminifera from the Tertiary of Central America, 283. Echinocystites pomum, 576. Echinodermata. See Crinoids, Cystoids, Echinoids, Edrioblastoids. INDEX 821 Echinoencrinites ; angulosus, 140; reticulatus, 135; senckenbergii, 135. Echinoids: Silurian pedicellariae of, 576. Edrioblastoids: Ordovician from Australia, 513. Edwards, D. A new plant from the Lower Old Red Sandstone of South Wales, 683. Eiffellithus turriseiffeli, 149. Elliott, G. F. Three new Tethyan Dasycladaceae (Calcareous Algae), 491. Elliottina deslongchampsi, 68. Ellipsagelosphaera sp., 143. Emiliana huxleyi, 799, 145. Encrinus liliiformis, 48. England: Bathonian fish otoliths, 246; Cretaceous angiosperm pollen, 421; ichthyosaur gastric con- tents, 376; lycopod sporophyll from Somerset, 445; new Llandovery diplograptids, 1 ; Silurian echinoid pedicellariae, 576; Silurian medusoid (?), 610; tabulate corals from Devon, 44. Epimastopora malaysiana, 491, 93. Equisetites, 633. Ericsonia sp ., 144. Eskdalia, 439; minuta, 442, 82. Esker, G. C., III. Cloour markings in Phacops and Greenops from the Devonian of New York, 498. Etheria elliptica, 88. Ethmorhabdus gallicus, 150. Eudesella mayalis, 68. Eulepidina, 285; favosa, 296, 57; undosa, 285, 49; u. laramblaensis , 296, 57. Eumorphoceras, 268; ( Edmooroceras ) dichalocium, 269, 46; pseudocoronula, 270, 46; tornquisti, 268, 46; ( Eumorphoceras ) medusa, 272, 47 ; m. sinuosum, 47. Europe: Carboniferous schizophoriid brachiopods from, 64. F Fabanella bononiensis, 146, 29. Feeding mechanisms: in Triassic brachiopods, 329. Figge, K. A goniatite fauna from the Visean/Namur- ian boundary, 264. Fish: Bathonian otoliths, 246; Silurian ostracoderm, 21. Florida: Eocene Cassigerinella, 368. Foraminifera: Cretaceous alveolinid, 526; Eocene Cassigerinella , 368; Indian nummulitid, 669; larger Tertiary, 283; new nummulitid genus, 435; Orbulin- oides, 371 ; Upper Devonian, 601. Fritz, W. H. Lower and early Middle Cambrian trilo- bites from the Pioche Shale, east-central Nevada, U.S.A., 183. G Gastropoda: Discohelix, Tethyan index fossil, 554; sinus-bearing monoplacophoran, 132. Genuclymenia, 536; angelini, 538; f rechi, 538; karpin- skii, 538; keepitensis, 536, 104. Gephyrocapsa oceanica, 145. Germany: pseudoplanktonic crinoid, 275; Visean / Namurian goniatites, 264. Ghana: non-marine Cretaceous ostracods, 259. Glansicystis baccata, 137. Glauconome rugosa, 98. Globigeraspis kugleri, 372. Globigerina ampliapertura, 369. Globigerinatheka barri, 'ill. Glossopleura, 205; sp. 1, 205, 43; sp. 2, 205, 43. Glyptocystites multiporus, 135. Glyptograptus ( Pseudoglyptograptus ) vas, 13. Goniatite: Visean/Namurian fauna, 264. Goniatites schaelkensis, 267 , 47. Goniolithus, 807 ; fluckigeri, 151. Grandispora echinata, 126, 27. Graptolites: assemblages from Birkhill Shales, 654; new Llandovery diplograptids, 1. Greenops boothi, 96. Gyrosteus subdeltoideus, 249. H Halkyardia bikiniensis, 292, 51. Harlanjohnsonella annulata, 494, 93, 94. Harper, C. W. See Walmsley, V. G. Haude, R. See Seilacher, A. Helicolepidina paucispira, 287, 292, 51. Helicosphaera carteri, 801, 147; seminulum, 147; sp. 147. Helicostegina soldadensis, 287. Hemicosmites', extraneus, 138; malum, 139; cf. pyri- formis, 139; cf. verrucosus, 138; spp., 138-40. Heterostegina, 290; ( Heterostegina ), 290; ( Vlerkina ), 290; ( Vlerkinella ) kugleri, 291, 51. Holwill, F. J. W. Tabulate corals from the Ilfracombe Beds (Middle-Upper Devonian) of North Devon, 44. Homocystites constrictus, 134. Hudson, J. D. The microstructure and mineralogy of the shell of a Jurassic mytilid (Bivalvia), 163. Hutsonia capelensis, 153, 30. Hymenozonotriletes, 125; hast ulus, 126, 26. I Ichthyosaur, 376, 72, 73. Illinois: Pennsylvanian fern from, 104. India: Eocene Nummulites from Assam, 669; structure of Vertebraria, 643. J Jaekelocystis hartleyi, 137. Jamaica: new Cretaceous alveolinid genus from, 526. Jamoytius kerwoodi, 21, 3-6. Jenkins, T. B. H. Famennian ammonoids from New South Wales, 535. Jurassic: dinosaur, 40; epizoic oysters, 19; fish oto- liths, 246; function of crinoid stem, 275; gastric contents of ichthyosaur, 376; index gastropod, 554; Kimmeridgian ammonite, 16; microstructure of mytilid shell, 163. K Kamptnerius magnificus, 152. Kemp, E. M. Probable angiosperm pollen from the British Barremian to Albian strata, 421. 822 INDEX Kilenyi, T. I., and Allen, N. W. Marine-brackish bands and their microfauna from the lower part of the Weald Clay of Sussex and Surrey, 141. Kistocare campbellensis, 230, 43. Knoxisporites pristinus , 123, 27. Kootenia, 195; aculacauda, 195, 41; brevispina , 196, 40; convoluta, 197, 41; crassa, 198,42; crassinucha, 198, 41; sp., 199, 43. Kotujella calva, 92. Krithodeophyton croftii, 684, 130-2. Krommelbein, K. The first non-marine Lower Cret- aceous ostracods from Ghana, West Africa, 259. L Lepidocarpon mazonensis, 454. Lepidocyclina, 284; armata, 285, 50; gigas, 299; man- telli, 295, 56, 57 ; montgomeriensis, 287 ; rdouvillei, 291, 55; yurnagunensis, 284, 49; y. morganopsis, 293, 295, 55, 56; ( Lepidocyclina ), 297 ; canellei, 297, 59; cf. crassicosta, 297, 58; sachsi, 299, 58 ; (Nephro- lepidina), 294; bikiniensis, 300, 58; b. pumilipapilla, 301, 59; wilsoni, 294, 56. Lepidophylloides acwninatus, 454. Lepidophyllum gracile, 454. Lepidostrobophyllum, 447 ; alatum, 448, 83, 84. Lepidostrobus, 453; brevifolius, 454; goodei, 454; lancifolius, 454; triangularis, 453. Leptolepis, 254; densus, 254; roddenensis, 255; tenuirostris, 254. Ligaments: preserved in Permian bivalves, 94. Liostrea, 472; anabarensis, 472; ex. gr. delta, 472; praeanabarensis, 472. Lipsanocystis magnus, 140. Lopha, 482; folium, 89; semiplana, 90. Lovenicystis angelini, 136. Loxolitluis armilla, 148. Lycopod: Carboniferous genus Eskdalia, 439; com- pressed sporophyll, 445. Lycospora torulosa, 123, 26. Lyttonia sp., 68. M Macrocystella, 580; azaisi, 111-13; a. midticristata, 113; mariae, 583, 111-13. Markalius cf. inversus, 144. Marthasterites tribrachiatus, 153. Martin, A. R. H. Aquilapollenites in the British Isles, 549. Maslovella, 365; africana, 365, 69-71; blackii, 366, 69; pulchra, 366, 69. McKerrow, W. S. See Ziegler, A. M. Megadesmus', cuneatus, 20; nobilissimus, 20. Megalosaurus bucklandi, 7, 8. Metacypris, 262; sp. A, 262, 45; sp. B, 262, 45. Mexico: new Cretaceous alveolinid genus, 526. Mimocystites, 589. Miogypsina cushmani, 295. Mollusca. See Ammonoidea, Bivalvia, Gastropoda. Monoplacophora : sinus-bearing, 132. Moorellina leptaenoides, 68. Mucrospirifer, 317; mucronatus, 63, 64. Myonia; morrisi, 20; valida, 19, 20. N Neococcolithus sp., 152. Nevada: Cambrian trilobites from Pioche Shale, 183. Newman, B. H. The Jurassic dinosaur Scelidosaurus Owen, 40. New South Wales: brachiopod mantle canal patterns, 389; Famennian ammonoids, 535; new rhynchon- elloid, 731; septate dalmanellid, 627; Silurian tri- lobite, 691. New York: colour markings in Devonian trilobites, 498. Nisusia ferganensis, 91. Nummulites, 669; chavannesi, 670, 128, 129; fabianii, 673, 129; pengaronensis, 676, 128. O Ogygopsis sp., 201, 40. Old Red Sandstone: new plant from South Wales, 683. Olenellus gilberti, 193, 36. Olenoides, 199; steptoensis, 199, 39; sp., 200, 43. Onchocephalus papulus, 224, 36. Orbulinoides beckmannii, 371. Ordovician: earliest known crinoid, 406; edrioblas- toid from Australia, 513; Macrocystella, 580. Oryctocephalites typicalis, 202, 41. Oryctocephalus, 201; maladensis, 202, 41; cf. primus, 201, 40. Ostracoda: non-marine from Ghana, 259; Wealden marine-brackish, 141. Ostracoderm : new evidence on Jamoytius, 21 . Ostrea ; bononiae, 2; multiformis, 2. Oxinoxis ampullacea, 605, 117. Oysters: epizoic on ammonites, 19; functional study of, 458. P Pachyaspis, 231; gallagari, 231, 40; longa, 231, 40. Paedumias sp., 193, 36. Pagetia, 189; arenosa, 189, 43; clytia, 190, 43; mala- densis, 190, 43; mucrogena, 191, 43; resseri, 192, 38. Palaeodiscus ferox, 577. Palaeonummulites, 287; antiguensis, 301, 51; kugleri, 287, 50; palmarealensis, 288, 50; stainforthi, 288, 51. Palynology: Aquilapollenites in Britain, 549; Cret- aceous angiosperm pollen, 421; dispersed spores of Equisetites, 633; Tournaisian spore flora, 116. Pant, D. D., and Singh, R. S. The structure of Verte- braria indica Royle, 643. Parabolinoides, 410; bucephalus, 410, 77; contractus, 413; hebe, 413; palatus, 413. Paul, C. R. C. Macrocystella Callaway, the earliest glyptocystitid cystoid, 580. — Morphology and function of dichoporite pore- structures in cystoids, 697. Pectinatites ( Virgatosphinctoides ), 19; elegans, 2; reisiformis densicostatus, 2. Pemma papillatum, 147. Permian: brachial skeleton of Attenuatella, 783; pre- INDEX 823 served ligaments in bivalves, 94; structure of Vertebraria, 643. Peytonoceras ? involution, 271, 46. Phacops rana, 96. Phillips, T. L., and Andrews, H. N. Biscalitheca (Coenopteridales) from the Upper Pennsylvanian of Illinois, 104. Pholidophorus, 251 ; paradoxicus, 251 ; prae-elops, 252. Pienaar, R. N. Upper Cretaceous coccolithophorids from Zululand, South Africa, 361. Pilasporites allenii, 638, 123. Pioche Shale: trilobites from, 183. Planicardinia, 628; carroli, 629, 122. Plants: Carboniferous lycopod, 439, 445; new Lower Devonian, 683; Pennsylvanian fern, 104; structure of Vertebraria, 643; Triassic from Argentina, 236, 500. See also Algae, Palynology. Platyclymenia, 537; (P.) annulata annulata, 538, 104; a. densicosta, 540, 104; alterna, 543, 105; pattinsoni, 542; teicherti, 541, 104, 105. Pleurocystites-, elegans, 1 34 ; filitextus, 134, 140; rugeri, 134. Pleurodictyum sp., 60, 17. Pliolepidina, 289; tobleri panamensis, 289, 292, 52-55; ? sp., 289, 54. Pocock, Y. P. Carboniferous schizophoriid brachio- pods from Western Europe, 64. Podorhabdus cylindratus, 150. Poliella, 206; denticulata, 206, 38; germana, 207, 37; leipalox, 208, 38. Pollard, J. E. The gastric contents of an ichthyosaur from the Lower Lias of Lyme Regis, Dorset, 376. Poly podorhabdus madingleyensis, 150. Pontolithina moorevillensis, 149. Pontosphaera, 800; discopora, 800, 146; versa, 146. Praemytilus str at hair densis, 163, 31-35. Propectinatites, 16; websteri, 17, 1. Pseudo climacograptus, 2; ( Clinoclimacograptus ) retro- versus, 8; ( Metaclimacograptus ) hughesi, 3; undu- latus, 6. Pseudocrinites; gordoni, 136; perdewi, 136; pyriformis, 136. Pseudo phragmina ( Proporocyclina ) flint ensis, 290. Ptarmiganoides, 209; araneicauda, 209, 37; lepida?, 209, 37; sp., 210, 38. Pustulatisporites gibberosus, 118, 25. Pycnodonta hyotis, 89. Pyramus laevis, 19, 20. Q Queensland: Permian brachiopods from, 783. R Raistrickia, 118; clavata, 118, 25; corynoges, 118, 25. Ramseyocrinus, 406; cambriensis, 407, 76. Rectoclymenia ? sp., 544, 105. 1 Reptiles: ichthyosaur gastric contents, 376; Jurassic j] dinosaur, 40. i Rhabdolithina spp., 148, ] Rhexoxylon, 236, 509; africanum, 239; piatnitzkyi, 240, 44; tetrapteridoides, 239. Rhinocypris jurassica jurassica, 148, 29. Rickards, R. B. See Bulman, O. M. B. Ritchie, A. New evidence on Jamoytius kerwoodi White, an important ostracoderm from the Silurian of Lanarkshire, Scotland, 21. Roberts, J. Mantle canal patterns in Schizophoria (Brachiopoda) from the Lower Carboniferous of New South Wales, 389. Robinson, E. Chubbina, a new Cretaceous alveolinid genus from Jamaica and Mexico, 526. Rollins, H. B., and Batten, R. L. A sinus-bearing monoplacophoran and its role in the classification of primitive molluscs, 132. Rudwick, M. J. S. The feeding mechanisms and affinities of the Triassic brachiopods Thecospira Zugmayer and Bactrynium Emmrich, 329. Runnegar, B. Preserved ligaments in Australian Permian bivalves, 94. Rushton, A. W. A. Revision of two Upper Cambrian trilobites, 410. S Saccammina glenisteri, 603, 116. Samanta, B. K. Nummulites (Foraminifera) from the Upper Eocene Kohili Formation of Assam, India, 669. Savage, N. M. Australirhynchia, a new Lower Devon- ian rhynchonelloid brachiopod from New South Wales, 731. — Planicardinia, a new septate dalmanellid brachio- pod from the Lower Devonian of New South Wales, 627. — See also Walmsley, V. G. Scandinavia: new Llandovery diplograptids from, 1. Scapholithus sp., 148. Scelidosaurus harrisoni, 40, 7, 8. Schizocystis armata, 137. Schizophoria, 64, 389; annectans, 77, 18; connivens, 64, 18; gibbera, 69, 18; linguata, 72, 18; resupinata, 80, 18; verulamensis, 74, 75; woodi, 86, 18. Schopfites claviger, 121, 25. Schuleridea ( Eoschuleridea ) wealdensis, 151, 30. Scotland: Lower Silurian graptolite assemblages, 654; Silurian ostracoderm, 21; Tournaisian spore flora, 1 16. Scyphosphaera sp., 154. Seilacher, A., Drozdzewski, G., and Haude, R. Form and function of the stem in a pseudoplanktonic crinoid ( Seirocrinus ), 275. Seirocrinus subangularis, 275, 48. Shell structure: Billingsellacean brachiopods, 486; Jurassic mytilid, 163. Sherwin, L. Denckmannites (Trilobita) from the Silurian of New South Wales, 691. Silurian: Birkhill graptolite assemblages, 654; Dayia navicula, 612; Denckmannites from Australia, 691; echinoid pedicellariae, 576; Llandovery transgres- sion, 736; medusoid (?), 610; new Llandovery diplograptids, 1 ; ostracoderm, 21 ; resserellid brachiopod, 306. Sindulites sindensis, 435. 824 INDEX Singh, R. S. See Pant, D. D. Sinuitopsis acutilira, 134, 28. Smout, A. H. See Eames, F. E. Sollasites horticus, 144. Sorosphaeroidea adhaerens, 604, 116. South Africa: Cretaceous coccoliths from Zululand, 361. South Wales: new plant from Lower Old Red Sand- stone, 683. Spencia quadrata, 232, 40. Sphaerocystites multifasciata, 136. Sphaeronchus, 250; circularis, 251; dorsetensis, 250. Sphaerophthalmus, 414; alatus, 417, 78; humilis , 417, 78; major, 416, 78; majusculus, 417. Sporadoceras, 545; inflexion, 545, 105; cf. rotundum, 546, 105. Staurocystis quadrifasciatus, 137. Staurolithites sp., 148. Stephanolithion, 807; bigoti, 152; laffittei, 152. Sternbergella ( Parasternbergella ), 147; wolburgi, 148, 30. Stinton, F. C., and Torrens, H. S. Fish otoliths from the Bathonian of southern England, 246. Strachan, I. A new medusoid (?) from the Silurian of England, 610. Strobilocystites calvini, 136, 140. Sullivan, H. J. A Tournaisian spore flora from the Cementstone Group of Ayrshire, Scotland, 116. Suppiluliumaella polyreme, 495, 95. Syracosphaera; hystrica, 148; pulchra, 148. Syspacephalus ?, 225; cf. uncus, 225, 36; sp., 226, 37. T Taxonomy: problems in study of coccoliths, 793. Tertiary: AquilapoUenites, 549; Eocene Cassigerinella, 368; Indian Nummulites, 669; larger foraminifera, 283; new nummulitid, 435; Orbulinoides, 371. Tethys: calcareous algae of, 491; index gastropod in Jurassic of, 554. Thamnopora , 54; boloniensis, 55, 12, 14; cervicornis, 56, 11, 12; cronigera, 57, 12; irregularis, 58, 11; polyforata, 57, 11; polymorpha, 59, 10; tumefacta, 60, 11. Thecospira, 329; haidingeri, 65; tyrolensis, 65; sp., 66, 68. Theriosynoecum, 158. Thomacystis tuberculata, 139. Thomas, B. A. A revision of the Carboniferous lyco- pod genus Eskdalia Kidston, 439. Toghill, P. The graptolite assemblages and zones of the Birkhill Shales (Lower Silurian) at Dobb’s Linn, 654. Tolypammina, 606; devoniana, 606, 114; helina, 607, 115; nexuosa, 607, 115; sp., 608, 114. Torrens, H. S. See Stinton, F. C. Tremalithus; burwellensis, 145 ; danicus, 145; dictyodus, 145; placomorphus, 145. Triassic: brachiopod feeding mechanisms, 329; leaf genus Dicroidium, 500; trunk of Rhexoxylon, 236. Tricolpites, 430; albiensis, 430, 81 ; sp. 1 , 431, 81 ; sp. 2, 432, 81. Trilobita: Cambrian from Nevada, 183; colour mark- ings in Devonian, 498; Silurian Denckmannites, 691 ; two Upper Cambrian, 410. Tucker, E. V. The atrypidine brachiopod Dayia navi- cula (J. de C. Sowerby), 612. U ?Utriculith, 257. V Vacunella ; curvata, 20; ?sp. nov., 20. Vallatisporites vallatus, 123, 26. Verrucosisporites, 121; scoticus, 121, 25; variotuber- culatus, 121, 26. Vertebraria indica, 643, 124-7. Vertebrata: fish otoliths, 246; ichthyosaur gastric contents, 376; Jurassic dinosaur, 40; Silurian ostracoderm, 21. Visbyella, 306; cumnockensis, 313, 61; nana, 311, 62; pygmaea, 310, 61; visbyensis, 307, 60. W Walmsley, V. G., Boucot, A. J., Harper, C. W., and Savage, N. M. Visbyella — a new genus of resserellid brachiopod, 306. Wealden: marine-brackish bands from Sussex and Surrey, 141. Webby, B. D. Astrocystites distans sp. nov., an edrio- blastoid from the Ordovician of eastern Australia, 513. Welsh Borderland: Llandovery transgression of, 736. Wendt, J. Discohelix (Archaeogastropoda, Euompha- lacea) as an index fossil in the Tethyan Jurassic, 554. Williams, A. Shell structure of the Billingsellacean brachiopods, 486. Z Zacanthoides, 211; demissus, 211, 42; sp., 212, 40. Zacanthopsis levis, 217, 36. Ziegler, A. M., Cocks, L. R. M., and McKerrow, W. S. The Llandovery transgression of the Welsh Borderland, 736. Zones: Silurian graptolite, 654. Zululand: Cretaceous coccoliths from, 361. Zygolithus] diplogrammus, 148; sp., 149. THE PALAEONTOLOGICAL ASSOCIATION COUNCIL 1968-9 President Professor Alwyn Williams, The Queen’s University, Belfast Vice-President Dr. W. S. McKerrow, University Museum, Oxford Treasurer Dr. C. Downie, Department of Geology, The University, Mappin Street, Sheffield 1 Membership Treasurer Dr. A. J. Lloyd, Department of Geology, University College, Gower Street, London, W.C.l Secretary Dr. J. M. Hancock, Department of Geology, King’s College, London Assistant Secretary Dr. W. D. I. Rolfe, Hunterian Museum, The University, Glasgow, W. 2 Editors Mr. N. F. Hughes, Sedgwick Museum, Cambridge Dr. Gwyn Thomas, Department of Geology, Imperial College, London, S.W.7 . Dr. I. Strachan, Department of Geology, The University, Birmingham, 15 Professor M. R. House, The University, Kingston upon Hull, Yorkshire Dr. R. Goldring, Department of Geology, The University, Reading Other members of Council Dr. F. M. Broadhurst, The University, Manchester Mr. M. A. Calver, Institute of Geological Sciences, Leeds Dr. C. B. Cox, King’s College, London Mr. D. Curry, Eastbury Grange, Northwood, Middlesex Dr. Grace Dunlop, Bedford College, London Mr. G. F. Elliott, 60 Fitzjohn Avenue, Barnet, Herts. Dr. A. Hallam, University Museum, Oxford Dr. Julia Hubbard, King’s College, London Dr. J. D. Hudson, The University, Leicester Dr. R. P. S. Jefferies, British Museum (Natural History), London Dr. J. D. Lawson, The University, Glasgow Dr. A. H. Smout, British Petroleum Company, Sunbury-on-Thames Professor H. B. Whittington, Sedgwick Museum, Cambridge Overseas Representatives Australia : Professor Dorothy Hill, Department of Geology, University of Queensland, Brisbane Canada : Dr. D. J. McLaren, Institute of Sedimentary and Petroleum Geology, 3303-33rd Street NW., Calgary, Alberta India : Professor M. R. Sahni, 98 The Mall, Lucknow (U.P.), India New Zealand: Dr. C. A. Fleming, New Zealand Geological Survey, P.O. Box 368, Lower Hutt West Indies and Central America: Mr. John B. Saunders, Geological Laboratory, Texaco Trinidad, Inc., Pointe-^-Pierre, Trinidad, West Indies Western U.S.A. : Professor J. Wyatt Durham, Department of Paleontology, University of California, Berkeley 4, Calif. Eastern U.S.A. : Professor J. W. Wells, Department of Geology, Cornell University, Ithaca, New York PALAEONTOLOGY VOLUME 1 1 ' PART 5 CONTENTS The structure of Vertebraria indica Royle. By D. d. pant and R. s. singh The graptolite assemblages and zones of the Birkhill Shales (Lower Silurian) at Dobb’s Linn. By p. toghill Nummulites (F or ammif era.) from the Upper Eocene Kopili Formation of Assam India. By b. k. samanta i /A new plant from the Lower Old Red Sandstone of South Wales. By DIANNE EDWARDS Denckmannites (Trilobita) from the Silurian of New South Wales. By L. sherwin Morphology and function of dichoporite pore-structures in cystoids. By C. R. C. PAUL Australirhynchia, a new Lower Devonian rhynchonelloid brachiopod from New South Wales. By N. M. savage The Llandovery transgression of the Welsh Borderland. By A. M. ziegler, l. r. m. cocks, and w. s. mckerrow The unusual brachial skeleton of Attenuatella convexa sp. nov. (Brachiopoda). By 3. ARMSTRONG Taxonomic problems in the study of coccoliths. By M. black 643 654 669 683 691 697 673 736 783 793 PRINTED IN GREAT BRITAIN AT THE UNIVERSITY PRESS, OXFORD BY VIVIAN RIDLER, PRINTER TO THE UNIVERSITY CO Z CO -•'* Z CO Z to ion NouniiiSNi_NviNosHims S3iavyan libraries Smithsonian institution Noun to ~T \ to — CO — CO sn libraries Smithsonian institution NouniiiSNi nvinoshxiins saiavaan libri ^ ^ — t/) \ ^ (/) 'ion NouniiiSNi nvinoshiiws saiavaan libraries smithsonian institution Noun to Z CO Z . tO -r ... CO X A' O ^ Z vo***'^/ *. an libraries smithsonian institution NouniiiSNi nvinoshjliws °S3 i avya nzu bri CO Z 07 to X tO As ^ M c c < ,y 5 2 W I | - "V S o ion'' NOliniUSNlAviNOSHiinS S3iavaanJLIBRARIESZSMITHS0NIANJ|NSTITUTI0N:rN0Un z r- v z r- z — — ^ — GO _ C/> ? an libraries smithsonian institution NouniiiSNi nvinoshiiins S3iavaan libr; 5 ^ z to z to z Z A —I IT .V _» T ZcAstoCX. — i /■ o to X H- s g » § . 4?^' i '®wv/ a , » A I V»y :f > s s XaA > ION NOimiliSNI_ NVIN0SH1IINS S3IMVaan libraries SMITHSONIAN INSTITUTION NOlin: to z \ 00 „ = to — to ^ CO X w to UJ JS LU y^I! c H v _ O ~ O XJ^SZ “■ -* 2 J Z _j — an LIBRARIES SMITHSONIAN INSTITUTION NOlinillSNI NVIN0SH1IWS S3iavaaH LIBRA — Z r* z r~ z CO 31 > 3J m xgNDCjz m m w CO rn (O £ co 5 — CO \ 5 CO ION NOlinillSNI NVINOSHHWS S3IMVaan LIBRARIES SMITHSONIAN INSTITUTION NOI10 to Z to z to z ... co < .. s < a\.. s X to tf. > 2 S v > ^inT^y ^ > 'W a n_ U 8 rar i es smithsonian_institution NoiiniusNi^NviNOSHinNs^s 3 1 y vy a \\_ li b r/ ^ ^ - .... *.«§ I .v*fe ? /^USks ■ >> •A' NVIN0SH1IWS°°S3 I a vy a nZU BRAR I ES^SMITHSONIAN^INSTITUTION NO!iniliSNI_NVINQSHlllNS to ~ co — to 2: - ^ uj co ^ CK — f < _J O NJV\o5 _ ’ O Xftos*£ -I 2 _j 2 -J ^ SMITHSONIAN INSTITUTION NOIlfllllSNI NVIN0SH1IWS S3iaVM9n LIBRARIES SMITHSONIAIs 2 r~ 2 C z O /T\asvTT\ “ . O — xtwo.TN O co ■*» ^=* •~) 1*^ ^ 3 m rn co ± co \ ± <0 nvinoshiiws S3iavaan libraries Smithsonian institution NoiinmsNi nvinoshihns CO 2 »,. CO 2 ... CO — ^ > vp* 5 ^ > Nsjpms!^ 2 ®\ > '‘W^ 2 SMITHSONIAN INSTITUTION MOIlfUIJlSm NVINOSHIIWS^S 3 I BVB 9 11 LIBRARIES SMITHSONIAN CO — CO 2 CO DL < «rmr H VcW.07 to Z! " W* m =j m '<» MviNOSHims^sa lavaan^UBRAR i es^ Smithsonian^ institution 2 NouniiiSNi~JNviMOSHiii/MS v 2 r~ 2 r* 2 «“ v m x^vosvvvx rn co 2 co _ co SMITHSONIAN INSTITUTION NOlinillSNI NVIN0SHJL1WS S3iaVMail LIBRARIES SMITHSONIAIS 2 » CO 2 CO 2 s 3 Wm>, s £$rm\ 3 «k z - ■4"m o P rw w p -x- «A-V w ss/t/' -tff:, X VC -5 ' W > ' 3E X^voscOJ^ > co • ■*“' 2 co 2 to nvinoshiiws S3 lava an libraries Smithsonian institution NoiinuxsNi_NviNOSHim$ CO — CO ~ CO c -> 2 _J 2 -1 ^ SMITHSONIAN^ INSTITUTION NOlinillSNI NVIN0SH1MS S3iaVHail LIBRARIES SMITHSONIAf yv- ^ /'vfn^v I o AfL> | p ./ rn ^ in g 2i > W 30 rn VLjtDv^ w Jr; W u; rn CO _ — co X E to NVIN0SH1IINS S3iavaan LIBRARIES SMITHSONIAN. INSTITUTION NOlinillSNI NVINOSHIIIM! CO 2 CO 2 V- CO 2 SSS&X E , . .*s s ■ny I t I O t 1 - bMITHSO1^’ INSTITUTION NOIlfllllSNI NVINOSHIIINS^SB I aVH 9 l*lf LI B RAR I ES^SMITHSONIAf .v tu