Skip to main content

Full text of "Palaeontology"

See other formats


IT  I ON  Z NOimillSNI^NVINOSHlI  WS  ^S3  lyvaail^LIBRARI  ES^  SMITHSON  IAN-*  INSTITUTION  2 NOIlfl 
z r-  v Z r~  z r;  z 

O Z 58k.  o xC$t7t£>v  Z O „ X'ogjvnX  O 


m 

ugn  LIBRARIES  SMITHSONIAN  INSTITUTION  NOIinillSNI-NVINOSHllWS  S3  I b \fU  B 11  ~LI  B R; 
z in  z t in  z m z 


\\  X 

in 

o x 'X  o iff/:, ' x Vs'Gr.  -3*7  ° 

> ^ > ‘ 5 > 

TION  NOlinillSNI  NVINOSHimS</>S3iyVMail  LIBRARIES^SMITHSONIAN  INSTITUTION  NOliO 
to  5 't  «o  . ^ ^ co  ^ to 


„^v.  c 

o z 5 “-r  5 

Z _l  Z _|  z -» 

Jail  LIBRARIES  SMITHSONIAN  INSTITUTION  NOlinillSNI  NVINQSH1IWS  S3I  tflVH  a H LIBR 
z z r-  z "" 


rn  \2i£S>^ 

CO  — CO  £ CO  _ 

ITION  NOlinillSNI  NVIN0SH1IINS  S3IHVM8n  LIBRARIES  SMITHSONIAN  INSTITUTION  NOIlf 

Z _ CO  z vy.  CO  Z 


yan  libraries  Smithsonian  institution  NOlinillSNI  nvinoshiiws  saiavaan  libr 

CO  ^ CO  ~ CO 


o _ ><0£ \^y  o 

JTION*-  NOIinillSNI_NVINOSHllWS  S3  I 8 V8 8 II “V I B R AR I E S*2  SMITHSONIAN^  I NSTITUTIO  N ^NOIlfl 
z I-  V Z r-  z i~  z 


y g n LIBRARIES  SMITHSONIAN  INSTITUTION  NOlinillSNI  NVINOSHimS  S3!8VHan  LIBR 

Z , CO  z CO  Z CO  z ■ 

< \»  S < V Z 2 


:A  z Ok  ^ z -H 

jo 

c 

«E  ''  3>  ^ «>W  ■>  5 3»’  S' 

to  W Z CO  '•’**  Z CO  Z CO 

ITION  NOlinillSNI  NVIN0SH1IWS  SBIBVaail  LIBRARIES  SMITHSONIAN  INSTITUTION  NOlin 
co  ~ in  — co  5;  co  _ 

~ ^ ” jj§|5|y  ^ 

5 W'  “ zl 

z 


S^>  - O Z ' O 

-J  Z _1  z -J 

nan  libraries  Smithsonian  institution  NouniiisNi  nvinoshiiws  S3iavaan  libr 

z r-  z r-  ...  z C” 


2 


iNOSHims  ssiavaan  libraries  Smithsonian  institution  NoiinmsNi  nvinoshiiws  s: 

,.2:  r*  2:  r-  z _ [I  > g 

O O — ^<Tn7^  o ^STIT(7>V  \v  o 


ithsonian  institution  NoiinmsNi  nvinoshiiws  saiavaan  LIBRARIES  SMITHSONIAN  ir 

2 > CO  2 " 


™ a.  ■»  I ^ X EC  CWi/  O 

i vx^  >'  5 >'■  S 2 

J/}  2 CO  2 CO  N 2 CO 

IN0SH1IIMS  S3iavaan  LIBRARIES  SMITHSONIAN  INSTITUTION  NOiiniusNi_NviNOSHiiws  s 

CO  — to  — CO  Z ,n 


O x^osv^  O 

2 _i  2 

I ITHSONIAN  INSTITUTION  NOlinillSNI  NVIN0SH1IWS  S3IHVaan  LIBRARIES  SMITHSONIAN  II 
2 1-  z 


^ rn  ^ Ngtmssx  m 

IN0SH1IWS  SBIdVaan- LIBRARIES  SMITHSONIAN  ""INSTITUTION  NOlinillSNI  NVIN0SH1MS  S 

CO  2 .....  CO  -p-  CO  2 


=]5 

ITHSONIAN  INSTITUTION  NOlinillSNI  NVINOSHIIWS^SB  I a VH  a 11  LIBRARIES  SMITHSONIAN  II 

- CO  — CO  X to  2 

^ 'm  ^ co  w , co 


e/  C V'£® 

Z!  %?  _ _ _ . 

O Ng*.  _ X^lllS^X  O v II  X^UllS^^  q XON  pcy  O 

IN0SHlllNS:ZS3iavaan"JLIBRAR  I ESZ SMITHSONIAN^ INSTITUTION ^NOimillSNI^NVINOSHimS  S 
^2  r~  2 z 


0 O 

TO 

> IpI 

JO  x ^ 

rn  !£  * ' ’ n ^ 

— CO  ± CO  £ — — 

IITHSONIAN  INSTITUTION  NOlinillSNI  NVIN0SH1IWS  S3iavaail  LIBRARIES  SMITHSONIAN  II 
co  2 co  2 > <5  — 2 

< V S < 'fc,  I < 

A 2 


— xTog.A'ox  _ 

^ o 

co  £1  _ 

X V5&,  •3FSJ  O 

2 , "y  / ' ' t 5 

2 ' W > ''  2 >'  2 

to  2 CO  2 co  A 2 co 

rIN0SHllWS_S3  I ava  a n_LI  B RAR  I ES  SMITHSONIANJNSTITUTION  _NOIinillSNI_NVINOSHimS/rtS 

O “ O cj 

2 _J  2 -I  z 

IITHSONIAN  INSTITUTION  NOlinillSNI  NVIN0SH1MS  S3iavaail  LIBRARIES  SMITHSONIAN  I 


VOLUME  25 


Palaeontology 


1982 


PUBLISHED  BY  THE 
PALAEONTOLOGICAL  ASSOCIATION 


LONDON 


Dates  of  Publication  of  Parts  of  Volume  25 


Part  l,pp.  1-225,  pis.  1-24 

1 January 

1982 

Part  2,  pp.  227-442,  pis.  25-45 

1 April 

1982 

Part  3,  pp.  443-679,  pis.  46-65 

1 July 

1982 

Part  4,  pp.  681-923,  pis.  66-100 

1 November  1982 

THIS  VOLUME  EDITED  BY  M.  G.  BASSETT,  K.  C.  ALLEN,  R.  A.  FORTEY  AND  A.  L.  PANCHEN 


Dates  of  Publication  of  Special  Papers  in  Palaeontology 

Special  Paper  No.  28  August  1982 
Special  Paper  No.  29  October  1982 


© The  Palaeontological  Association  1982 


Printed  in  Great  Britain 
at  the  University  Press,  Oxford 
by  Eric  Buckley 
Printer  to  the  University 


CONTENTS 


Part  Page 

Akpan,  E.  B.,  Farrow,  G.  E.  and  Morris,  N.  J.  Limpet  grazing  on  Cretaceous  algal-bored 
ammonites  2 361 

Aldridge,  R.  J.  A fused  cluster  of  coniform  conodont  elements  from  the  late  Ordovician  of 
Washington  Land,  Western  North  Greenland  2 425 

Aldridge,  R.  J.  See  Swift,  A.  and  Aldridge,  R.  J. 

Allen,  K.  C.  See  Marshall,  J.  E.  A.  and  Allen,  K.  C. 

Balinski,  A.  See  Biernat,  G.  and  Balinski,  A. 

Balson,  P.  S.  and  Taylor,  P.  D.  Palaeobiology  and  systematics  of  large  cyclostome  bryozoans 
from  the  Pliocene  Coralline  Crag  of  Suffolk  3 529 

Barwick,  R.  E.  See  Campbell,  K.  S.  W.  and  Barwick,  R.  E. 

Biernat,  G.  and  Balinski,  A.  Shell  structure  of  the  Devonian  retziid  brachiopod  Plectospira 
ferita  4 857 

Blake,  D.  B.  Somasteroidea,  Asteroidea,  and  the  affinities  of  Luidia  ( Platasterias ) latiradiata  1 167 

Boyd,  M.  J.  Morphology  and  relationships  of  the  Upper  Carboniferous  ai'stopod  amphibian 
Ophiderpeton  nanum  1 209 

Brenchley,  P.  J.  and  Cocks,  L.  R.  M.  Ecological  associations  in  a regressive  sequence:  the  latest 
Ordovician  of  the  Oslo-Asker  district,  Norway  4 783 

Campbell,  K.  S.  W.  and  Barwick,  R.  E.  A new  species  of  lungfish  Dipnorhynchus  from  New 
South  Wales  3 509 

Cocks,  L.  R.  M.  The  commoner  brachiopods  of  the  latest  Ordovician  of  the  Oslo-Asker 
district,  Norway  4 755 

Cocks,  L.  R.  M.  See  Brenchley,  P.  J.  and  Cocks,  L.  R.  M. 

Coombs,  W.  P.  Jr.  Juvenile  specimens  of  the  ornithischian  dinosaur  Psittacosaurus  1 89 

Cope,  J.  C.  W.  See  Paul,  C.  R.  C.  and  Cope,  J.  C.  W. 

Crame,  J.  A.  Late  Jurassic  inoceramid  bivalves  from  the  Antarctic  Peninsula  and  their 
stratigraphic  use  3 555 

Curry,  G.  B.  Ecology  and  population  structure  of  the  Recent  brachiopod  Terebratulina  from 
Scotland  2 227 

De  Jersey,  N.  J.  An  evolutionary  sequence  in  Aratrisporites  miospores  from  the  Triassic  of 
Queensland,  Australia  3 665 

Donovan,  D.  T.  and  Howarth,  M.  K.  A rare  lytoceratid  ammonite  from  the  Lower  Lias  of 
Radstock  2 439 

Elliott,  G.  F.  A new  calcareous  green  alga  from  the  Middle  Jurassic  of  England:  its 
relationships  and  evolutionary  position  2 43 1 

Farrow,  G.  E.  See  Akpan,  E.  B.,  Farrow,  G.  E.  and  Morris,  N.  J. 

Forsten,  A.  The  taxonomic  status  of  the  Miocene  horse  genus  Sinohippus  3 673 

Fraser,  N.  C.  A new  rhynchocephalian  from  the  British  Upper  Trias  4 709 

Freeman,  E.  F.  Fossil  bone  recovery  from  sediment  residues  by  the  ‘Interfacial  Method’  3 471 

Fursich,  F.  T.  and  Palmer,  T.  J.  The  first  true  anomiid  bivalve?  4 897 

Gale,  A.  S.  and  Smith,  A.  B.  The  palaeobiology  of  the  Cretaceous  irregular  echinoids  Infulaster 
and  Hagenowia  1 1 1 

Geys,  J.  F.  Two  salenoid  echinoids  in  the  Danian  of  the  Maastricht  area  2 265 

Harland,  R.  A review  of  Recent  and  Quaternary  organic-walled  dinoflagellate  cysts  of  the 
genus  Protoperidinium  2 369 

Harland,  T.  L.  and  Torrens,  H.  S.  A redescription  of  the  Bathonian  red  alga  Solenopora 
jurassica  from  Gloucestershire,  with  remarks  on  its  preservation  4 905 

Henry,  J.-L.  See  Romano,  M.  and  Henry,  J.-L. 

Higgins,  A.  C.  and  Varker,  W.  J.  Lower  Carboniferous  conodont  faunas  from  Ravenstone- 
dale,  Cumbria  1 145 


CONTENTS 


iv 

Part  Page 

Higgins,  A.  C.  and  Wagner-Gentis,  C.  H.  T.  Conodonts,  goniatites  and  the  biostratigraphy  of 

the  earlier  Carboniferous  from  the  Cantabrian  Mountains,  Spain  2 313 

Howarth,  M.  K.  See  Donovan,  D.  T.  and  Howarth,  M.  K. 

Jefferson,  T.  H.  Fossil  forests  from  the  Lower  Cretaceous  of  Alexander  Island,  Antarctica  4 681 

Kier,  P.  M.  Rapid  evolution  in  echinoids  1 1 

Lund,  R.  and  Melton,  W.  G.  Jr.  A new  actinopterygian  fish  from  the  Mississippian  Bear  Gulch 

Limestone  of  Montana  3 485 

Marshall,  J.  E.  A.  and  Allen,  K.  C.  Devonian  miospore  assemblages  from  Fair  Isle,  Shetland  2 277 

Mateer,  N.  J.  Osteology  of  the  Jurassic  lizard  Ardeosaurus  brevipes  (Meyer)  3 461 

Melton,  W.  G.  See  Lund,  R.  and  Melton,  W.  G. 

Milner,  A.  R.  Small  temnospondyl  amphibians  from  the  Middle  Pennsylvanian  of  Illinois  3 635 

Morris,  N.  J.  See  Akpan,  E.  B.,  Farrow,  G.  E.  and  Morris,  N.  J. 

Palmer,  T.  J.  See  Fursich,  F.  T.  and  Palmer,  T.  J. 

Paul,  C.  R.  C.  and  Cope,  J.  C.  W.  A parablastoid  from  the  Arenig  of  South  Wales  3 499 

Riding,  R.  and  Voronova,  L.  Affinity  of  the  Cambrian  alga  Tubomorphophyton  and  its 
significance  for  the  Epiphytaceae  4 869 

Rieppel,  O.  A new  genus  of  shark  from  the  Middle  Triassic  of  Monte  San  Giorgio,  Switzerland  2 399 

Rogers,  M.  J.  A description  of  the  generating  curve  of  bivalves  with  straight  hinges  1 109 

Romano,  M.  and  Henry,  J.-L.  The  trilobite  genus  Eoharpes  from  the  Ordovician  of  Brittany 
and  Portugal  3 623 

Smith,  A.  B.  Tooth  structure  of  the  pygasteroid  sea  urchin  Plesiechinus  4 891 

Smith,  A.  B.  See  Gale,  A.  S.  and  Smith,  A.  B. 

Stein,  W.  E.  Jr.  The  Devonian  plant  Reimannia,  with  a discussion  of  the  Class  Progymno- 

spermopsida  3 605 

Surlyk,  F.  and  Zakharov,  V.  A.  Buchiid  bivalves  from  the  Upper  Jurassic  and  Lower 
Cretaceous  of  East  Greenland  4 727 

Swift,  A.  and  Aldridge,  R.  J.  Conodonts  from  the  Upper  Permian  strata  of  Nottinghamshire 
and  North  Yorkshire  4 845 

Taylor,  P.  D.  See  Balson,  P.  S.  and  Taylor,  P.  D. 

Thulborn,  R.  A.  Liassic  plesiosaur  embryos  reinterpreted  as  shrimp  burrows  2 351 

Torrens,  H.  S.  See  Harland,  T.  L.  and  Torrens,  H.  S. 

Tunnicliff,  S.  P.  A revision  of  late  Ordovician  bivalves  from  Pomeroy,  Co.  Tyrone,  Ireland  1 43 

Turner,  R.  E.  Reworked  acritarchs  from  the  type  section  of  the  Ordovician  Caradoc  Series, 

Shropshire  1 119 

Turner,  S.  A new  articulated  thelodont  (Agnatha)  from  the  early  Devonian  of  Britain  4 879 

Varker,  W.  J.  See  Higgins,  A.  C.  and  Varker,  W.  J. 

Voronova,  L.  See  Riding,  R.  and  Voronova,  L. 

Wagner-Gentis,  C.  H.  T.  See  Higgins,  A.  C.  and  Wagner-Gentis,  C.  H.  T. 

Wang  Lixin.  See  Wang  Ziqiang  and  Wang  Lixin 

Wang  Ziqiang  and  Wang  Lixin.  A new  species  of  the  lycopsid  Pleuromeia  from  the  early 
Triassic  of  Shanxi,  China,  and  its  ecology  1 215 

Wellstead,  C.  F.  A Lower  Carboniferous  ai'stopod  amphibian  from  Scotland  1 193 

Wilson,  M.  V.  H.  A new  species  of  the  fish  Amia  from  the  Middle  Eocene  of  British  Columbia  2 413 

Young,  G.  C.  Devonian  sharks  from  south-eastern  Australia  and  Antarctica  4 817 

Zakharov,  V.  A.  See  Surlyk,  F.  and  Zakharov,  V.  A. 

Zdebska,  D.  A new  zosterophyll  from  the  Lower  Devonian  of  Poland  2 247 

Zhang  Zhongying.  Upper  Proterozoic  microfossils  from  the  Summer  Isles,  N.  W.  Scotland  3 443 


Palaeontology 

VOLUME  25  ■ PART  1 JANUARY  1982 


Published  by 

The  Palaeontological  Association  London 
Price  £1 7 


THE  PALAEONTOLOGICAL  ASSOCIATION 


The  Association  publishes  Palaeontology  and  Special  Papers  in  Palaeontology.  Details  of  membership  and  subscription  rates 
may  be  found  inside  the  back  cover. 

The  journal  Palaeontology  is  devoted  to  the  publication  of  papers  on  all  aspects  of  palaeontology.  Review  articles  are 
particularly  welcome,  and  short  papers  can  often  be  published  rapidly.  A high  standard  of  illustration  is  a feature  of  the 
journal.  Four  parts  are  published  each  year  and  are  sent  free  to  all  members  of  the  Association.  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  Dr.  R.  A.  Fortey,  Palaeontological  Association,  Department  of  Palaeontology,  British  Museum 
(Natural  History),  Cromwell  Road,  London  SW7  5BD,  England,  who  will  supply  detailed  instructions  for  authors  on  request 
(these  were  published  in  Palaeontology  1977,  20,  pp.  921-929). 

Special  Papers  in  Palaeontology  is  a series  of  substantial  separate  works;  the  following  are  available  (post  free). 

1.  (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  plates.  Price  £8  (U.S.  $17-50). 

2.  (for  1968):  Evolution  of  the  Shell  Structure  of  Articulate  Brachiopods,  by  a.  williams.  55  pp.,  27  text-figs.,  24  plates. 
Price  £5  (U.S.  $11). 

3.  (for  1968):  Upper  Maestrichtian  Radiolaria  of  California,  by  Helen  p.  foreman.  82  pp.,  8 plates.  Price  £3  (U.S.  $6-50). 

4.  (for  1969):  Lower  Turonian  Ammonites  from  Israel,  by  R.  freund  and  M.  raab.  83  pp.,  15  text-figs.,  10  plates.  Price  £3 
(U.S.  $6-50). 

5.  (for  1969):  Chitinozoa  from  the  OrdoVician  Viola  and  Fernvale  Limestones  of  the  Arbuckle  Mountains,  Oklahoma, 
by  w.  a.  M.  jenkins.  44  pp.,  10  text-figs.,  9 plates.  Price  £2  (U.S.  $4-50). 

6.  (for  1969):  Ammonoidea  from  the  Mata  Series  (Santonian-Maastrichtian)  of  New  Zealand,  by  r.  a.  Henderson.  82  pp., 
13  text-figs.,  15  plates.  Price  £3  (U.S.  $6-50). 

7.  (for  1970):  Shell  Structure  of  the  Craniacea  and  other  Calcareous  Inarticulate  Brachiopoda,  by  a.  williams  and 
a.  d.  wright.  51  pp.,  17  text-figs.,  15  plates.  Price  £1-50  (U.S.  $3-50). 

8.  (for  1970):  Cenomanian  Ammonites  from  Southern  England,  by  w.  j.  Kennedy.  272  pp.,  5 tables,  64  plates.  Price  £8 
(U.S.  $17-50). 

9.  (for  1971):  Fish  from  the  Freshwater  Lower  Cretaceous  of  Victoria,  Australia,  with  Comments  on  the  Palaeo- 
environment,  by  M.  waldman.  130  pp.,  37  text-figs.,  18  plates.  Price  £5  (U.S.  $11). 

10.  (for  1971):  Upper  Cretaceous  Ostracoda  from  the  Carnarvon  Basin,  Western  Australia,  by  r.  h.  bate.  148  pp.,  43  text- 
figs.,  27  plates.  Price  £5  (U.S.  $11). 

11.  (for  1972):  Stromatolites  and  the  Biostratigraphy  of  the  Australian  Precambrian  and  Cambrian,  by  m.  r.  Walter. 
268  pp.,  55  text-figs.,  34  plates.  Price  £10  (U.S.  $22). 

12.  (for  1973):  Organisms  and  Continents  through  Time.  A Symposium  of  23  papers  edited  by  N.  F.  hughes.  340  pp., 
132  text-figs.  Price  £10  (U.S.  $22)  (published  with  the  Systematics  Association). 

13.  (for  1 974) : Graptolite  studies  in  honour  of  O.  M.  B.  Bulman.  Edited  by  r.  b.  rickards,  d.  e.  jackson,  and  c.  p.  hughes. 
261  pp.,  26  plates.  Price  £10  (U.S.  $22). 

14.  (for  1974):  Palaeogene  Foraminiferida  and  Palaeoecology,  Hampshire  and  Paris  Basins  and  the  English  Channel,  by 
j.  w.  Murray  and  c.  a.  wright.  171  pp.,  45  text-figs.,  20  plates.  Price  £8  (U.S.  $17-50). 

15.  (for  1975):  Lower  and  Middle  Devonian  Conodonts  from  the  Broken  River  Embayment,  North  Queensland,  Australia, 
by  p.  G.  telford.  100  pp.,  9 text-figs.,  16  plates.  Price  £5-50  (U.S.  $12). 

16.  (for  1975):  The  Ostracod  Fauna  from  the  Santonian  Chalk  (Upper  Cretaceous)  of  Gingin,  Western  Australia,  by 
j.  w.  neale.  131  pp.,  40  text-figs.,  22  plates.  Price  £6-50  (U.S.  $14-50). 

17.  (for  1976):  Aspects  of  Ammonite  Biology,  Biogeography,  and  Biostratigraphy,  by  w.  j.  Kennedy  and  w.  a.  cobban. 
94  pp.,  24  text-figs.,  11  plates.  Price  £6  (U.S.  $13). 

18.  (for  1976):  Ostracoderm  Faunas  of  the  Delorme  and  Associated  Siluro-Devonian  Formations,  North  West  Territories, 
Canada,  by  d.  l.  dineley  and  e.  j.  loeffler.  218  pp.,  78  text-figs.,  33  plates.  Price  £20  (U.S.  $44). 

19.  (for  1977):  The  Palynology  of  Early  Tertiary  Sediments,  Ninetyeast  Ridge,  Indian  Ocean,  6yE.  M.  kemp  and  w.  K.  Harris. 
74  pp.,  2 text-figs.,  8 plates.  Price  £7  (U.S.  $15-50). 

20.  (for  1977):  Fossil  Priapulid  Worms,  by  s.  c.  morris.  159  pp.,  99  text-figs.,  30  plates.  Price  £16  (U.S.  $35). 

21.  (for  1978):  Devonian  Ammonoids  from  the  Appalachians  and  their  bearing  on  International  Zonation  and  Correla- 
tion, by  M.  r.  house.  70  pp.,  12  text-figs.,  10  plates.  Price  £12  (U.S.  $26-50). 

22.  (for  1978,  published  1979):  Curation  of  palaeontological  collections.  A joint  colloquium  of  The  Palaeontological  Associa- 
tion and  Geological  Curators’  Group.  Edited  by  m.  g.  bassett.  280  pp.,  53  text-figs.  Price  £25  (U.S.  $55). 

23.  (for  1979):  The  Devonian  System.  A Palaeontological  Association  International  Symposium.  Edited  by  M.  R.  house,  c.  t. 
scrutton,  and  M.  G.  bassett.  353  pp.,  102  text-figs.,  1 plate.  Price  £30  (U.S.  $66). 

24.  (for  1980):  Dinofiagellate  Cysts  and  Acritarchs  from  the  Eocene  of  southern  England,  by  j.  b.  bujak,  c.  downie,  g.  l. 
eaton,  and  G.  L.  williams.  104  pp.,  24  text-figs.,  22  plates.  Price  £15  (U.S.  $33). 

25.  (for  1980):  Stereom  Microstructure  of  the  Echinoid  Test,  by  a.  b.  smith.  81  pp.,  20  text-figs.,  24  plates.  Price  £15  ($33). 

26.  (for  1981):  The  fine  structure  of  Graptolite  periderm,  by  p.  r.  crowther.  1 19  pp.,  37  text-figs.,  20  plates.  Price  £25  ($55). 

27.  (for  1981):  Late  Devonian  Acritarchs  from  the  Carnarvon  Basin, Western  Australia,  by  G.  playford  andw..  s.  dring.  78  pp., 
10  text-figs.,  19  plates.  Price  £15  ($33). 

© The  Palaeontological  Association,  1982 


Cover:  The  acritarch  Umbellasphaeridium  saharicum  Jardine  et  al.  1972  from  the  Tournaisian  Bedford  Shale  of  Ohio, 
U.S.A.,  x 1000.  I.G.S.  specimen  number  MPK  3152  deposited  in  the  micropalaeontological  collection  of  the  Institute  of 
Geological  Sciences,  Leeds.  Photograph  by  S.  G.  Molyneux.  This  group  of  fossils  was  unknown  when  the  Association  was 

founded,  twenty-five  years  ago. 


RAPID  EVOLUTION  IN  ECHINOIDS 


by  PORTER  M.  KIER 


Abstract.  The  evolution  of  the  irregular  echinoid  and  of  the  sand  dollar  occurred  in  a very  short  time.  The  first 
irregular  echinoid  appears  abruptly  in  the  Early  Jurassic  (Sinemurian);  and  by  the  Toarcian,  only  ten  million 
years  later,  irregular  echinoids  possess  all  the  features  necessary  to  permit  them  to  live  buried  in  the  sediment. 
The  first  clypeasteroid  appears  in  the  Paleocene.  By  the  middle  Eocene  its  very  specialized  descendants,  the  sand 
dollars,  have  a worldwide  distribution.  This  rapid  evolution  and  diversification  seem  to  result  from  a sudden 
adaptive  breakthrough.  The  presence  of  so  few  intermediates  indicates  the  evolutionary  steps  must  have  been 
large. 

Two  of  the  most  significant  events  in  the  evolution  of  the  echinoids  are  the  development  of  the 
irregular  echinoid  and  the  subsequent  appearance  of  the  sand  dollar.  Both  of  these  events  were 
believed  to  have  occurred  over  a long  period  of  time  (Durham  1966,  p.  U289)  but  new  evidence 
suggests  otherwise.  The  irregular  echinoid  was  assumed  to  have  evolved  during  the  Triassic  and 
perhaps  during  the  latter  part  of  the  Paleozoic.  However,  study  of  the  Triassic  faunas  indicates  that 
no  irregulars  were  present  then  and  that  the  great  changes  from  the  regular  to  the  irregular  echinoid 
occurred  during  the  early  part  of  the  Early  Jurassic.  Likewise,  it  was  believed  (Durham  1966, 
p.  U290)  that  the  clypeasteroids  arose  during  the  Late  Cretaceous  and  Paleocene.  New  evidence 
suggests  that  the  first  clypeasteroid  actually  appeared  in  the  Paleocene,  and  that  the  great  evolution  to 
the  sand  dollar  occurred  during  the  early  Eocene. 

EVOLUTION  OF  THE  IRREGULAR  ECHINOID 

The  first  irregular  echinoid,  Plesiechinus  hawkinsi  Jesionek-Szymanska  (text-fig.  1b),  occurs  in  the 
Early  Jurassic  (Sinemurian).  This  echinoid,  a pygasterid,  differs  from  all  other  echinoids  of  the  same 
age  or  older  in  having  an  asymmetrical  test  with  small  tuberculation,  short  and  numerous  spines, 
differentiated  pores  with  the  adapical  pores  larger  than  the  adoral  ones,  posteriorly  eccentric 
periproct,  and  presumably  keeled  teeth  (although  the  lantern  was  not  found  with  any  specimens  of 
this  species,  Melville  (1961)  found  keeled  teeth  in  a Pygaster). 

One  might  assume  that  the  great  changes  necessary  to  derive  this  irregular  echinoid  from  a regular 
form  would  have  taken  a very  long  time.  Many  workers  believe  that  these  changes  occurred  during 
the  Triassic  or  possibly  in  the  Paleozoic.  They  attributed  the  lack  of  intermediates  found  during  the 
Triassic  to  the  poor  fossil  record  of  that  period.  However,  that  does  not  appear  to  be  the  case. 
Although  it  is  true  that  few  echinoids  are  known  from  the  Triassic,  particularly  from  the  Early  and 
Middle  Triassic,  a prolific  echinoid  fauna  occurs  in  the  Late  Triassic  St.  Cassian  beds  of  Italy.  These 
beds  (Kier  19776)  have  been  painstakingly  searched  by  Rinaldo  Zardini  who  has  found  many 
complete  tests  and  thousands  of  fragments.  I have  searched  through  all  this  material  and  have  failed 
to  find  any  keeled  teeth  or  any  test  fragments  with  the  fine  tuberculation  of  an  irregular  echinoid.  As 
shown  by  Kier  (1977a),  irregular  echinoids  are  much  more  likely  to  be  preserved  than  regular 
echinoids.  Had  they  been  living  during  St.  Cassian  time,  they  should  have  been  preserved  as  fossils. 
Furthermore,  no  irregular  echinoids  or  any  form  resembling  them  have  been  found  in  the  lowermost 
Jurassic  (Hettangian). 

I am  convinced  that  this  change  from  a regular  echinoid  to  Plesiechinus  hawkinsi  must  have 
occurred  extremely  quickly  during  the  latter  part  of  the  Hettangian  or  early  Sinemurian.  Evolution 
then  continued  at  a rapid  rate,  because  by  the  Toarcian,  still  within  the  Early  Jurassic,  the  cassiduloid 
Galeropygus  dumortieri  (Paris)  (text-fig.  lc)  has  appeared  with  all  the  features  of  an  irregular 

IPalaeontology,  Vol.  25,  Part  1,  1982,  pp.  1-9,  pis.  1-2.] 


2 


PALAEONTOLOGY,  VOLUME  25 


echinoid.  It  is  more  advanced  than  P.  hawkinsi  in  having  a more  flattened  and  elongate  test,  a more 
eccentric  periproct,  larger  adapical  pores  forming  incipient  petals,  and  smaller,  more  numerous, 
tubercles  and  spines.  Its  peristome  is  smaller,  elongate,  and  eccentric  anteriorly.  There  are  no  gills. 
The  most  profound  differences  are  the  lack  of  a lantern  and  lantern  supports  and  the  presence  of  well- 
developed  phyllodes. 

Presumably,  one  of  the  reasons  for  this  rapid  evolution  was  that  these  changes  enabled  echinoids  to 
occupy  a habitat  not  available  to  them  before.  They  were  now  able  to  burrow  (Smith  1978)  into  the 
sediment  and  extract  the  organic  material  contained  within.  Simpson  (1944,  1953),  and  later  Stanley 
(1979),  proposed  that  a higher  taxon  arises  rapidly  through  the  occurrence  of  a sudden  adaptive 
breakthrough.  During  the  Triassic  all  echinoids  apparently  lived  on  the  surface  of  the  sea  floor  and 
could  not  burrow  into  the  sediment.  They  all  had  jaws  and  grooved  teeth  which  were  used 
(presumably  like  modern  regular  echinoids)  to  tear  off  and  chew  organic  material  which  was  then 
passed  into  the  gut.  The  small  amount  of  faecal  discharge  could  be  easily  carried  away  by  water 
currents.  Most  modern  irregular  echinoids  lacking  teeth  feed  differently.  They  burrow  (text-fig.  1) 


LOWER  JURASSIC 


HETTANGIAN  SINEMURI  AN  TOARCI  AN 


ELONGATION  OF  THE  TEST ► 

DECREASE  IN  SIZE  OF  TUBERCLES  AND  SPINES  ► 

INCREASE  IN  NUMBER  OF  TUBERCLES  AND  SPINES ► 

SHIFTING  OF  ANUS  POSTERIORLY ► 

SHIFTING  OF  PERISTOME  ANTERIORLY ► 

DECREASE  IN  SIZE  OF  PERISTOME ► 

DEVELOPMENT  OF  PETALS ► 

DEVELOPMENT  OF  PHYLLODES ► 

GROOVED  TEETH ► KEELED  TEETH ► NO  TEETH 

text-fig.  i . Evolution  of  the  irregular  echinoid  showing  the  changes  that  enabled  the  echinoid  to  live  buried 
in  the  substrate,  a,  an  Hettangian  regular  echinoid  such  as  Diademopsis;  b,  the  earliest  known  irregular, 
Plesiechinus  hawkinsi  Jesionek-Szymanska;  c,  a cassiduloid,  Galeropygus. 


EXPLANATION  OF  PLATE  1 

Figs.  1-5.  Togocyamus  seefriedi  (Oppenheim).  Paleocene,  Ekekoro  Formation,  Ekekoro  quarry,  55  km  north- 
west of  Lagos,  Nigeria.  1,  accessory  pore  on  dorsal  side  of  USNM  312503  just  beyond  petal  I.  The  large 
pores  on  the  lower  right  side  are  at  the  end  of  petal  I,  x 74.  2,  enlarged  view  of  accessory  pore  in  USNM 

312504  showing  large  neural  pore,  x 500.  3,  side  view  of  USNM  312505,  x 19.  4,  top  view  of  USNM 

312503,  x 13.  5,  bottom  view  of  USNM  312504,  x 15. 


PLATE  1 


KIER,  Togocyamus 


4 


PALAEONTOLOGY,  VOLUME  25 


into  the  sediment,  collect  large  amounts  of  sediment  with  their  tube-feet,  and  pass  it  through  the  gut 
while  extracting  the  organic  material.  The  sediment  is  then  expelled  through  the  anus  (periproct). 
Most  of  the  differences  between  an  irregular  and  regular  echinoid  relate  to  these  differences  in  mode 
of  feeding. 

The  posterior  migration  of  the  periproct  made  it  possible  for  the  irregular  echinoid  to  leave  this 
large  volume  of  discharged  sediment  in  its  trail  rather  than  over  its  dorsal  surface.  At  the  same  time 
that  the  periproct  migrated,  most  irregular  echinoids  increased  the  oxygen-gathering  capability  of 
their  dorsal  tube-feet  by  greatly  broadening  them.  The  result  was  the  formation  of  the  ‘petals’  so 
typical  of  most  irregulars.  Specialized  tube-feet  were  also  produced  around  the  mouth.  They  were 
larger,  more  numerous,  and  were  used  to  collect  sediment  that  was  then  passed  to  the  mouth. 


EVOLUTION  OF  THE  CLYPEASTEROID  ECHINOID 

The  first  clypeasteroid  echinoid,  Togocyamus  (PI.  1,  figs.  1-5),  appears  in  Paleocene  strata.  By  the 
middle  Eocene  the  highly  specialized  sand  dollar  had  evolved.  Until  now  it  was  believed  that  these 
developments  required  a long  time  from  the  Cretaceous  through  the  Eocene.  New  evidence  indicates 
that  the  change  was  much  more  rapid.  In  fact  the  change  from  a cassiduloid  ancestor  to  a 
clypeasteroid  probably  occurred  within  the  Paleocene;  and  the  change  from  a primitive  clypeasteroid 
to  a sand  dollar  occurred  during  the  early  Eocene. 

The  earlier  ‘Cretaceous’  origin  of  the  clypeasteroids  was  based  on  the  supposed  occurrences  of 
three  species  of  clypeasteroids  in  the  Late  Cretaceous.  These  occurrences  are  probably  erroneous. 
One  of  the  three  species  is  too  poorly  preserved  to  be  identified,  and  the  stratigraphic  data  with  the 
other  two  are  inadequate.  Extensive  collecting  has  been  done  in  the  Late  Cretaceous  beds  where  these 
two  species  supposedly  were  found  and  neither  Meijer  (1965)  nor  Ernst  (1972)  have  found  any 
clypeasteroids.  Ernst,  in  the  course  of  his  study,  examined  over  15000  echinoids  from  the  Late 
Cretaceous. 

The  chronologically  later  origin  of  the  clypeasteroids  is  supported  by  their  absence  from  the 
Paleocene  of  the  Western  Hemisphere.  An  exhaustive  search  by  Kier  through  washings  of  the 
Paleocene  Vincentown,  Aquia,  and  Clayton  Formations  (from  which  eighteen  echinoid  species  are 
known)  revealed  no  fragments  that  could  be  identified  as  clypeasteroid,  although  echinoid  fragments 
are  very  common.  It  is  noteworthy  that  the  small  fibularids,  like  those  found  in  the  Paleocene  of 
Africa,  were  absent  but  normally  would  be  expected  in  this  material.  They  usually  live  buried  in  this 
type  of  sediment  and  are  small  enough  to  have  their  tests  preserved  intact  in  these  sands. 

The  earliest  confirmed  clypeasteroid,  Togocyamus,  occurs  in  the  Paleocene  of  West  Africa.  It 
differs  so  markedly  from  all  previous  echinoids  that  there  is  disagreement  as  to  its  origin.  Some 
workers  consider  the  clypeasteroids  to  be  derived  from  a holectypoid,  but  Phelan  (1977,  p.  419) 
makes  a strong  case  for  their  derivation  from  a juvenile  stage  of  a cassiduloid.  Its  evolution  must  have 
been  extremely  rapid  and  probably  occurred  within  the  Paleocene.  The  Late  Cretaceous  echinoid 
record  is  not  only  extensive  but  very  well  studied  (Ernst  1972)  and  no  intermediates  have  been 
found  there.  Current  information  indicates  that  the  subsequent  diversification  and  radiation  of 
the  clypeasteroids  from  a fibularid  to  a sand  dollar  occurred  very  abruptly  (text-fig.  2)  in  the  early  part 
of  the  middle  Eocene.  Only  two  species,  both  fibularids,  are  known  in  the  Paleocene:  Togocyamus 


EXPLANATION  OF  PLATE  2 

Figs.  1-3.  Sismondia  logotheti  Fraas.  Early  Eocene,  from  Siout  ( = Assiout),  Egypt.  1,  2,  3,  top,  bottom,  side 
views  of  topotype  B22908,  Museum  National  d’Histoire  Naturelle,  Paris,  x 5. 

Figs.  4-6.  Periarchus  lyelli  (Conrad).  Middle  Eocene,  Castle  Hayne  Formation,  from  North  Carolina  Lime 
Company  pit,  adjacent  to  Tuckahoe  Church,  3-8  miles  (61  km)  west  of  Comfort,  Jones  County,  North 
Carolina.  4,  5,  6,  side,  top,  bottom  views  of  USNM  312506,  x 1. 


PLATE  2 


3 


KIER,  Sismondia,  Periarchus 


6 


PALAEONTOLOGY,  VOLUME  25 


seefriedi  (Oppenheim)  and  T.  alloiteaui  Roman  and  Gorodiski.  During  the  early  Eocene  the 
clypeasteroids  (text-fig.  2)  were  confined  to  Africa  and  India  and  consist  of  six  species  of  fibularids 
and  two  species  of  Sismondia,  the  most  primitive  member  of  the  laganids.  By  the  end  of  the  middle 
Eocene,  the  clypeasteroids  had  worldwide  distribution  with  over  62  species  representing  20  genera 
and  4 families. 

In  summary,  the  first  clypeasteroid  appeared  in  the  Paleocene,  and  by  the  middle  Eocene  its  far 
more  complex  and  specialized  descendants  were  present  all  around  the  world. 


MIDDLE 

EOCENE 


EARLY 

EOCENE 


PALEOCENE 


text-fig.  2.  Evolution  and  radiation  of  the  clypeasteroid  echinoid.  The  Paleocene 
clypeasteroids  are  confined  to  West  Africa  and  are  a small,  high  species  with  incipient  petals. 
In  the  early  Eocene  they  are  more  flattened,  have  more-developed  petals,  and  occur  in  Africa 
and  India.  By  the  middle  Eocene  they  are  fully  developed  sand  dollars  and  are  present  all 
around  the  world. 


The  Cretaceous  origin  of  the  clypeasteroids  was  suggested  by  previous  workers  not  only  because  of  j 
the  supposed  occurrence  of  Cretaceous  species,  but  also  because  it  was  believed  that  the  Palaeocene 
fibularids  were  not  primitive  enough  to  be  ancestral  to  all  later  clypeasteroids.  New  information  is 
now  available  on  Togocyamus.  It  is  more  primitive  than  previously  thought  and  could  be  close  to  the 
ancestral  stock  of  all  clypeasteroids.  Although  it  has  been  assumed  that  its  lantern  supports  were  like 
those  found  in  later  fibularids  with  each  support  composed  of  a single  interambulacral  plate,  the 
supports  in  T.  seefriedi  (Oppenheim)  are  both  interambulacral  and  ambulacral  in  origin.  Each  ; 
support  ( text-fig.  3c)  is  formed  by  the  extension  of  the  primordinal  interambulacral  plate  and  the 
adjacent  half-ambulacral  plates.  This  discovery  is  important  for  it  has  been  suggested  (Philip  1965,  p. 

58;  Kier  1970,  p.  105)  that  the  clypeasteroids  could  be  divided  into  two  orders  on  the  basis  of  the 
character  of  the  lantern  supports.  The  suborder  Clypeasterina  includes  all  those  clypeasteroids 


KIER:  ECHINOID  EVOLUTION 


7 


EOCENE 


text-fig.  3.  The  lantern  supports  in  the  clypeasteroids.  c,  the 
oldest  clypeasteroid,  the  Paleocene  Togocyamus  seefriedi 
(Oppenheim)  has  lantern  supports  (indicated  by  a solid  line) 
composed  of  ambulacral  (shaded)  and  interambulacral  plates; 
a,  b.  Eocene  clypeasteroids  have  lantern  supports  composed  of 
interambulacral  plates  as  in  the  suborder  Scutellina  (a)  or 
ambulacral  plates  as  in  Clypeasterina  (b). 


having  ambulacral  lantern  supports  (text-fig.  3b);  in  the  suborder  Scutellina  the  supports  are 
interambulacral  (text-fig.  3a).  The  presence  of  supports,  both  interambulacral  and  ambulacral  in 
origin,  in  the  oldest  known  and  most  primitive  clypeasteroid  adds  weight  to  the  supposition  that 
Togocyamus  is  close  to  the  ancestral  stock  of  both  suborders.  By  simply  reducing  the  size  of  the 
ambulacral  extensions  and  increasing  the  size  of  the  interambulacral  ones,  supports  could  be 
produced  that  are  typical  of  later  species  of  the  fibularids  and  the  rest  of  the  Scutellina.  Conversely, 
the  reduction  in  the  size  of  the  interambulacral  and  increase  in  the  ambulacral  extensions  would 
produce  the  typical  Clypeasterina  supports  that  first  appear  in  the  late  Eocene. 

A second  discovery  in  Togocyamus  is  that  its  accessory  pores  are  few  in  number  and  are  restricted 
to  the  border  of  the  ambulacra  (PI.  1,  fig.  1).  As  pointed  out  by  Durham  (1966,  p.  U451),  accessory 
pores  are  an  exclusive  feature  of  the  clypeasteroids,  occurring  in  all  species.  The  fact  that  they  are  less 
well  developed  in  this  species  than  in  any  other  is  further  evidence  of  the  primitiveness  of  this  form. 

In  the  light  of  the  primitive  features  of  Togocyamus , we  can  now  postulate  the  evolutionary  history 
of  the  earliest  clypeasteroids: 

1.  The  Paleocene  Togocyamus  (PI.  1,  figs.  3-5)  has  a small,  high  test  with  its  periproct  in  a primitive  dorsal 
position,  slightly  developed  petals  with  simple  nonconjugate  pores,  a very  erect  lantern  with  supports  of 
interambulacral  and  ambulacral  origin.  Its  few  accessory  pores  are  confined  to  the  borders  of  the  ambulacra.  It 
has  no  food  grooves  and  has  a large  peristome. 

2.  By  the  early  Eocene,  Sismondia  (PI.  2,  figs.  1-3)  has  a larger,  more  flattened  test,  a lower  lantern, 
interambulacral  lantern  supports,  and  a ventral  periproct.  The  petals  are  better  developed  with  conjugate  pores; 
the  accessory  pores  are  far  more  numerous. 


PALAEONTOLOGY,  VOLUME  25 


3.  The  middle  Eocene  Protoscutella  and  Periarchus  (PI.  2,  figs.  4-6)  are  typical  sand  dollars  having  a large,  very 
flattened  test,  food  grooves,  very  wide  and  low  lantern,  and  a very  small  peristome.  Accessory  pores  are  spread 
all  over  the  ambulacra,  and  the  test  is  strongly  reinforced  by  calcareous  supports  that  are  pierced  by  many  canals 
for  the  water  vascular  system  serving  these  pores.  The  adoral  plate  arrangement  is  now  distinctive;  there  are  far 
fewer  and  larger  plates  than  in  earlier  clypeasteroids.  As  pointed  out  by  Durham  (1966,  p.  U450),  these  changes 
in  the  adoral  plates  result  from  the  flattening  of  the  test.  In  flattened  species  the  number  of  plates  on  the  adoral 
surface  is  determined  at  an  early  ontogenetic  stage  and  thereafter  growth  is  only  by  enlargement  of  the  plates. 

The  morphological  changes  that  produced  the  sand  dollar  are  specializations  that  enabled  the 
echinoid  to  live  more  efficiently  in  sand  (Seilacher  1979).  The  flattened  test  made  it  easier  for  the 
echinoid  to  burrow.  The  accessory  tube-feet  were  used  to  pass  sand  over  the  top  of  the  test  and  to 
convey  food  to  the  food  grooves.  The  better-developed  petals  increased  the  respiratory  capability  of 
the  petaloid  tube-feet  by  increasing  their  area.  The  internal  supports  strengthened  the  test,  enabling 
the  sand  dollar  to  live  in  environments  of  higher  energy.  The  change  from  the  erect  lantern  and  large 
peristome  in  the  primitive  fibularid  to  the  low  lantern  with  horizontal  teeth  and  small  peristome  in  the  \ 
sand  dollar  reflects  a change  in  eating  habits.  According  to  Markel  (1974,  1978;  Markel,  Gorny,  and 
Abraham  1977)  and  Nichols  (1959),  the  fibularid  uses  its  lantern  to  scrape  organic  material  from  sand 
grains.  This  feeding  method  puts  little  stress  on  the  teeth.  The  sand  dollar,  however,  uses  the  teeth  for 
grinding  and  chewing.  The  great  stress  can  be  withstood  because  the  teeth  are  horizontal,  and  the 
stress  is  transmitted  to  the  long  axis  of  the  teeth.  The  larger  peristome  in  the  fibularid  permits  the  teeth 
to  extend  further  out  of  the  test  to  grasp  food;  in  the  sand  dollar  the  sand  is  passed  to  the  teeth  within 
the  test. 

If  it  is  true,  as  suggested  herein,  that  the  clypeasteroids  originated  in  the  Paleocene  then  all  these 
changes  necessary  to  derive  a sand  dollar  from  a cassiduloid  ancestor  occurred  within  20  million 
years.  Certainly  there  can  be  little  question  that  the  evolution  from  a fibularid  to  a typical  sand  dollar 
occurred  between  the  beginning  of  the  early  Eocene  and  the  latter  part  of  the  middle  Eocene,  a period 
of  less  than  10  million  years. 


CONCLUSIONS 

The  sudden  appearance  of  the  first  irregular  and  the  first  clypeasteroid  echinoids  and  their  rapid 
diversification  indicate  a rate  of  evolution  much  faster  than  previously  supposed.  The  mechanisms 
producing  these  great  changes  are  uncertain,  but  the  evolutionary  steps  must  have  been  large.  If  each 
speciation  event  produced  only  small  morphological  change,  than  a multitude  of  transitional  species 
would  have  resulted.  The  fossil  record  of  the  irregular  echinoids  is  excellent  (Kier  1977a).  Even  if  this 
rapid  evolution  occurred  in  peripherally  isolated  populations,  somewhere  in  the  world  we  should 
have  found  more  of  these  transitional  species.  I believe  the  absence  of  a large  number  of  transitional 
species  is  explained  not  because  they  have  not  been  preserved  as  fossils,  but  because  they  never 
existed.  This  conclusion  supports  Stanley’s  (1979,  p.  212)  statement  that  ‘rates  of  evolution  are 
highest  early  in  adaptive  radiation,  when  degree  of  divergence  per  speciation  event  is  high  . . .’. 

Acknowledgements.  I thank  Kenneth  Towe  and  Thomas  Waller  (Smithsonian  Institution),  and  Thomas  Phelan 
and  Andrew  Smith,  for  helpful  discussions.  The  manuscript  was  read  by  Richard  Boardman,  Richard  Grant, 
and  David  Pawson.  I thank  them  for  their  suggestions.  Mary  Hurd  Lawson  did  the  light  photography;  Walter 
Brown  did  the  SEM  photography.  Larry  Isham  made  all  the  text-figures  with  his  customary  skill  and 
inventiveness.  Jean  Roman  (Museum  National  d’Histoire  Naturelle,  Paris)  and  Richard  Jefferies  (British 
Museum,  Natural  History)  very  kindly  lent  specimens. 


REFERENCES 

Durham,  j.  w.  1966  Classification,  U270-U295;  Clypeasteroids,  U450-U491.  In  Moore,  R.  C.  (ed.).  Treatise  on 
invertebrate  paleontology , Part  U,  Echinodermata  3,  1-2,  Geol.  Soc.  Am.  and  Univ.  Kansas  Press. 
ernst,  G.  1972.  Grundfragen  der  Stammesgeschichte  bei  irregularen  Echiniden  der  nordwesteuropaischen 
Oberkreide.  Geol.  Jb.  A4,  63-175,  pis.  1-7. 


KIER:  ECHINOID  EVOLUTION 


9 


gould,  s.  J.  and  eldredge,  n.  1977.  Punctuated  equilibria:  The  tempo  and  mode  of  evolution  reconsidered. 
Paleobiology , 3,  115-151. 

kier,  p.  m.  1970.  Lantern  support  structures  in  the  clypeasteroid  echinoids.  J.  Paleont.  44,  98-109,  pis.  23,  24. 

— 1977a.  The  poor  fossil  record  of  the  regular  echinoid.  Paleobiology,  3,  168-174. 

19777>.Triassic  echinoids.  Smithson.  Contr.  Paleobiol.  30,  1-88,  pis.  1-21. 

MARKEL,  K-  1974.  Morphologie  der  Seeigelzahne.  V.  Die  Zahne  der  Clypeastroida  (Echinodermata, 
Echinoidea).  Zh.  morph.  Tiere,  78,  221-256. 

1978.  On  the  teeth  of  the  recent  cassiduloid  Echnolampas  depressa  Gray,  and  on  some  Liassic  fossil  teeth 

nearly  identical  in  structure  (Echinodermata,  Echinoidea).  Zoomorphologie,  89,  125-144. 

— gorny,  p.  and  abraham,  k.  1977.  Microarchitecture  of  sea  urchin  teeth.  Fortschr.  Zool.  24,  103-1 14. 
meijer,  m.  1965.  The  stratigraphical  distribution  of  echinoids  in  the  chalk  and  tuffaceous  chalk  in  the 

neighborhood  of  Maastricht  (Netherlands).  Meded.  geol.  Sticht.  17,  21-25. 
melville,  r.  v.  1961.  Dentition  and  relationships  of  the  echinoid  genus  Pygaster  J.  L.  R.  Agassiz,  1836. 
Palaeontology,  4,  243-246. 

nichols,  d.  1959.  The  histology  and  activities  of  the  tube-feet  of  Echinocyamus  pusillus.  Q.  Jl  microsc.  Sci.  100, 
539-555. 

phelan,  t.  f.  1977.  Comments  on  the  water  vascular  system,  food  grooves,  and  ancestry  of  the  clypeasteroid 
echinoids.  Bull.  mar.  Sci.  Gulf  Caribb.  27,  400-422. 
philip,  G.  M.  1965.  Classification  of  echinoids.  J.  Paleont.  39,  45-62. 
seilacher,  a.  1979.  Constructional  morphology  of  sand  dollars.  Paleobiology,  5,  191-221. 
simpson,  G.  G.  1944.  Tempo  and  mode  in  evolution,  237  pp.  New  York,  Columbia  Univ.  Press. 

— 1953.  The  major  features  of  evolution,  434  pp.  New  York,  Columbia  Univ.  Press. 

smith,  a.  b.  1978.  A comparative  study  of  the  life  style  of  two  Jurassic  irregular  echinoids.  Lethaia,  11,  57-66. 
Stanley,  s.  m.  1979.  Macroevolution,  pattern  and  process,  332  pp.  San  Francisco,  Freeman. 


Typescript  received  28  March  1981 
Revised  typescript  received  30  April  1981 


PORTER  M.  KIER 


Department  of  Paleobiology 
Smithsonian  Institution 
Washington  D.C.  20560 
U.S.A. 


THE  PALAEOBIOLOGY  OF  THE 
CRETACEOUS  IRREGULAR  ECHINOIDS 
INFULASTER  AND  HAGENO  WIA 

by  ANDREW  S.  GALE  and  ANDREW  B.  SMITH 


Abstract.  The  taxonomy  of  the  infaunal  holasteroids  Infulaster  and  Hagenowia  is  revised  in  the  light  of  new 
material  from  the  Senonian  Chalk  of  southern  England.  Investigation  of  the  plating  structure  in  Hagenowia 
permits  a more  precise  definition  of  its  species,  and  a better  understanding  of  the  evolution  of  the  rostrum. 
The  majority  of  structural  modifications  in  the  Infulaster-Hagenowia  lineage  were  caused  directly,  or 
indirectly,  by  apical  elongation  and  size  reduction.  A detailed  study  of  test  and  ambulacral  pore  morphology, 
tuberculation,  and  spines,  taken  in  comparison  with  living  echinoids,  allows  reconstruction  of  the  life  habits 
of  Infulaster  and  Hagenowia.  The  evolutionary  changes  are  related  to  the  adoption  of  a specialized  method 
of  feeding  and  an  attempt  to  avoid  predation. 


T he  echinoid  genus  Hagenowia , in  which  the  apical  part  of  the  test  is  drawn  out  to  form  a rostrum, 
is  the  most  bizarre  of  a wide  range  of  holasteroids  which  inhabited  the  Upper  Cretaceous  Chalk 
Sea  of  north-west  Europe.  The  ancestry  of  this  genus  was  traced  by  Wright  and  Wright  (1949)  to 
Infulaster,  and  the  evolutionary  story  further  discussed  and  elaborated  by  Ernst  and  Schulz  (1971). 

The  Upper  Chalk  in  England  falls  into  two  distinct  faunal  provinces,  the  boundary  of  which 
ran  approximately  east-west  through  north  Norfolk  (Peake  and  Hancock  1961;  Reid  1976;  Rawson 
et  al.  1978).  Infulaster  and  Hagenowia  have  been  regarded  as  characteristic  members  of  the  northern 
province  fauna  (where  they  occur  commonly)  and  recorded  hitherto  only  as  occasional  rarities  in 
the  south.  Hagenowia  is  the  most  common  fossil  in  one  interval  of  the  Yorkshire  Upper  Chalk 
and  has  been  used  there  as  a local  zonal  index  in  place  of  Micros  ter  coranguinum  (Rowe  1904; 
Wright  and  Wright  1942). 

Collecting  during  recent  years  from  the  Senonian  Chalk  of  south-east  England  has  yielded 
abundant  material  of  both  genera  and  shows  that  they  occur  frequently  at  certain  stratigraphical 
levels  in  at  least  part  of  the  southern  province.  This  material  has  provided  much  new  information 
(particularly  concerning  the  rostrum  of  Hagenowia)  which  has  allowed  a reappraisal  of  the  taxonomy 
and  phylogeny  of  Infulaster  and  Hagenowia. 

Irregular  echinoid  lineages  in  the  Chalk  have  provided  a number  of  classical  evolutionary  stories, 
of  which  the  best  known  is  in  Micros  ter  (Rowe  1899).  Progressive  morphological  changes  in 
Micraster  were  interpreted  by  Nichols  (1959)  as  adaptations  to  an  increasing  depth  of  burial.  A 
similar  explanation  has  been  advanced  for  the  Infulaster-Hagenowia  story  (Nichols  1959;  Ernst 
and  Schulz  1971).  Our  work  suggests  that  changes  in  this  lineage,  particularly  the  development  and 
modification  of  the  rostrum,  are  related  to  the  adoption  of  a specialized  method  of  feeding, 
and  avoidance  of  predation. 

METHODS  OF  STUDY 

Although  the  surface  details  of  Infulaster  and  Hagenowia  from  southern  England  are  often 
well  preserved,  the  tests  are  invariably  distorted  by  compactional  crushing.  For  this  reason, 
biometrical  studies  on  the  material  have  not  been  attempted. 

In  a detailed  study  of  the  rostrum  of  H.  elongata  (Nielsen)  Schmid  (1972)  used  a special  technique 
to  elucidate  the  plating  structure.  This  took  advantage  of  the  naturally  separate  reflectivity  of 


IPalaeontology,  Vol.  25,  Part  1,  1982,  pp.  11-42,  pis.  3-6| 


12 


PALAEONTOLOGY,  VOLUME  25 


individual  plates,  accentuated  by  gold  coating.  This  method  proved  to  be  effective  on  H.  blackmorei 
Wright  and  Wright,  but  unsuccessful  on  earlier  members  of  the  genus.  To  work  out  the  plating 
arrangements  of  these,  naturally  weathered  rostra  were  stained  with  black  ink  to  pick  out  the 
sutures.  Surface  details  of  the  echinoids,  particularly  tubercles  and  pores,  were  studied  with  the 
scanning  electron  microscope 

Abbreviations  used  for  museum  collections  are  as  follows:  BMNH— British  Museum  (Natural  History); 
IGS— Institute  of  Geological  Sciences,  London;  MMH— Mineralogical-Geological  Museum,  Copenhagen. 

STRATIGRAPHY 

In  this  paper  the  usage  of  stages  and  zones  within  the  Upper  Chalk  of  England  have  been  adopted  from 
Rawson  et  al.  (1978).  Text-fig.  1 gives  a generalized  succession  of  all  but  the  lowest  Senonian  in  east  Kent, 
and  is  based  on  the  cliff  sections  between  Dover  and  Kingsdown,  and  on  the  Isle  of  Thanet.  The  section  is 
abstracted  from  detailed  measurements  made  by  one  of  us  (A.  S.  G.)  and  gives  only  selected  marker  horizons, 
to  which  records  of  species  are  related. 

Details  of  the  distribution  of  Infulaster  and  Hagenowia  in  north-west  Germany  are  taken  from  Ernst  and 
Schulz  (1971,  1974).  In  connection  with  this,  it  is  important  to  note  that  the  Turonian-Coniacian  boundary 
in  Germany  is  taken  at  a different  level  from  that  generally  used  in  England,  where  the  Holaster  planus-Micraster 
cortestudinarium  zonal  junction  is  taken  as  the  stage  boundary  (Rawson  et  al.  1978).  The  base  of  the  Coniacian 
in  the  German  sense  probably  falls  within  the  basal  few  metres  of  the  coranguinum  Zone  at  Dover,  on  the 
basis  of  evidence  from  Inoceramus  faunas. 

SYSTEMATIC  PALAEONTOLOGY 

Order  holasteroida  Durham  and  Melville,  1957 
Family  holasteridae  Pictet,  1857 
Genus  infulaster  Desor,  1858 

Type  Species.  Cardiaster  hagenowi  d’Orbigny  1853  (=  Spatangus  excentricus  Woodward  1833;  see  Wright  and 
Wright  1949,  p.  455). 

Diagnosis.  Small  to  medium-sized  holasterids,  proportionately  narrow  and  tall,  with  apex  positioned 
anteriorly.  Anterior  ambulacrum  set  in  a sulcus  which  runs  from  the  apex  to  the  transversely 
rounded  peristome  on  the  adoral  surface.  Paired  ambulacra  flush  with  test,  non-petaloid.  Apical 
system  uninterrupted,  elongated,  with  four  genital  pores.  Periproct  situated  at  summit  of  steep 
posterior  truncation.  Plastron  metasternal,  weakly  keeled.  Marginal  fasciole  present. 

Remarks.  The  diagnosis  given  above  essentially  follows  the  description  of  Infulaster  given  by  Wright 
and  Wright  (1949),  with  some  elaboration. 

Wright  and  Wright  (1949)  regarded  /.  excentricus  as  the  only  valid  species  of  Infulaster,  and 
placed  I.  krausei  Desor,  I.  borchardi  Desor,  I.  hagenowi  (d’Orbigny),  and  I.  tuberculatus  Valette  in 
synonomy  with  it.  I.  tuberculatus  is  here  considered  to  be  a separate  species,  and  its  descendant, 
Hagenowia  infulasteroides  Wright  and  Wright  is  transferred  to  Infulaster.  This  species  displays  a 
number  of  features  previously  thought  to  be  exclusive  to  Hagenowia,  notably  interruption  of  the 
pastron  and  elongation  of  the  dorsolateral  plate  columns.  Infulaster  thus  differs  from  Hagenowia 
only  in  its  uninterrupted  apical  system  and  lack  of  a rostrum. 

Infulaster  excentricus  (Woodward,  1833) 

Text-fig.  2 (1) 

Remarks.  Large  forms  of  Infulaster  occur  commonly  in  the  Middle  and  Upper  Turonian  of  the 
northern  province  in  England  (north  Norfolk,  Lincolnshire,  Yorkshire)  and  in  the  eastern  extension 
of  the  province  in  north-west  Germany  (Ernst  and  Schulz  1971),  Poland  (Nietsch  1921)  and  the 
Caucasus  (Moskveena  1959).  For  these,  the  name  I.  excentricus  is  used  provisionally,  and  it  is  not 
intended  in  this  paper  to  revise  and  redescribe  this  material,  which  may  include  more  than  one  species. 


text-fig.  1 . Generalized  succession  in  the  Senonian  Chalk  (lowest  beds  omitted)  of  the  Kent  coast,  with  selected 
marker  bands  only,  showing  the  distribution  of  species  of  Infulaster  and  Hagenowia.  Scale  on  left  of  column 
in  5 m intervals.  Flints  solid  black. 


14 


PALAEONTOLOGY,  VOLUME  25 


Only  one  specimen  of  I.  excentricus  is  known  from  southern  England,  from  the  Coniacian  M. 
cortestudinarium  Zone  at  Dover,  3-2  m below  the  Lower  East  Cliff  Marl  (text-fig.  2 (1);  BMNH 
E76832,  A.  S.  Gale  Coll.) 

Infulaster  tuberculatus  Valette,  1913 
Plate  4,  figs.  2,  11;  Plate  5,  fig.  3;  Plate  6,  fig.  5;  text-fig.  2 (2-4) 

1913  Infulaster  tuberculatus  Valette,  p.  5,  fig.  1. 
p.  1949  Infulaster  excentricus  (Woodward)  Wright  and  Wright,  p.  456. 

Holotype.  Valette’s  solitary  specimen  came  from  the  Coniacian  zone  H,  of  Rosoy  near  Sens,  France.  The 
specimen  is  presumably  with  Valette’s  collection,  now  in  the  University  of  Dijon. 

Diagnosis.  Small  Infulaster  in  which  the  test  is  short  and  proportionally  tall.  The  apex  may  be 
acutely  angled.  Plastron  uninterrupted  by  ambulacra  I,  V. 


text-fig.  2.  Infulaster  spp.  from  east  Kent:  1,  Infulaster  excentricus  (Woodward),  oral  view,  from 
3-2  m below  summit  of  Micraster  cortestudinarium  Zone,  East  Cliff,  Dover.  BMNH  E76832  A.  S. 
Gale  Coll.  2-4, 1.  tuberculatus  Valette.  2,  group  of  specimens  from  basal  M.  coranguinum  Zone,  0-8 
m above  Upper  East  Cliff  Marl,  cliffs  north-east  of  St.  Margaret’s  Bay,  near  Dover,  Kent.  BMNH 
E76833,  A.  S.  Gale  Coll.  3 a,  b , oral  and  lateral  views  of  individual  without  stratigraphical  location, 
east  of  Dover,  Kent.  BMNH  E10300  Cockburn  Coll.  4 a-c,  dorsal,  oral,  and  lateral  views  of  specimen 
from  0-5  m below  Lower  East  Cliff  Marl,  top  of  M.  cortestudinarium  Zone,  East  Cliff,  Dover,  Kent. 

BMNH  E76834,  A.  S.  Gale  Coll.  All  specimens  x 2.  Coated  in  ammonium  chloride. 


GALE  AND  SMITH:  CRETACEOUS  ECHINOIDS 


15 


Remarks.  Small  Infulaster  with  lengths  of  10-20  mm  are  common  in  the  Lower  coranguinum  Zone 
of  east  Kent.  These  are  almost  invariably  crushed  and  fragmentary,  but  the  few  well-preserved 
individuals  show  close  similarities  in  shape  with  Valette’s  figures  of  I.  tuber culatus.  The  species 
differs  from  I.  excentricus  in  having  a proportionally  shorter  test  with  a steep  posterior  slope  in 
addition  to  the  consistent  difference  in  size.  Many  individuals  are  slightly  inflated  just  posterior  to 
the  apex.  Specimens  from  the  higher  part  of  the  range  of  the  species  at  Dover,  although  distorted, 
have  acutely  angled  apices,  and  can  only  be  distinguished  from  I.  infulasteroides  (Wright  and 
Wright)  by  their  uninterrupted  plastrons. 

Occurrence.  At  Dover  the  species  occurs  in  the  highest  1-5  m of  the  M.  cortestudinarium  Zone, 
and  the  basal  18  m of  the  M.  coranguinum  Zone  (text-fig.  1).  The  species  is  most  common  just 
beneath  a band  of  large  flints  about  1 m above  the  upper  East  Cliff  Marl,  where  clusters  of  individuals 
are  found.  The  type  specimen  came  from  the  Coniacian  H of  Rosoy,  a level  equivalent  to 
approximately  the  lower  half  of  the  coranguinum  Zone. 

Infulaster  infulasteroides  (Wright  and  Wright  1949) 

Plate  4,  fig.  2 

1949  Hagenowia  infulasteroides  Wright  and  Wright,  p.  470,  figs.  17,  18. 

1971  Hagenowia  infulasteroides  Wright  and  Wright;  Ernst  and  Schulz,  p.  138,  pi.  13,  figs.  1-4; 
text-fig.  6. 

Types.  The  holotype  is  a flint  steinkern  from  the  Haldon  Gravel  of  Devon  (Wright  and  Wright  1949,  fig.  17; 
BMNH  E8403).  A paratype,  similarly  preserved,  comes  from  flint  gravel  at  Lulworth,  Dorset  (BMNH  El 709). 
Both  were  probably  derived  originally  from  the  upper  part  of  the  coranguinum  Zone.  A second  paratype,  a 
crushed,  incomplete  test  from  the  coranguinum  Zone  of  the  North  Foreland,  near  Broadstairs,  Kent  (BMNH 
E33886,  Rowe  Coll.)  almost  certainly  came  from  a level  2-4  m below  Whitaker’s  3-inch  band,  the  only  horizon 
at  which  the  species  is  common  in  east  Kent. 

Diagnosis.  Test  with  acutely  angled  apex.  Plates  of  interambulacra  1 and  4 on  sides  of  test  elongated. 
Plastron  interrupted  by  ambulacra  I and  V. 

Remarks.  This  species  is  transferred  from  Hagenowia  to  Infulaster  on  account  of  its  undivided 
apical  system,  and  the  absence  of  a rostrum.  The  general  morphology  and  variation  is  well  illustrated 
by  Ernst  and  Schulz  (1971). 

Occurrence.  In  east  Kent  /.  infulasteroides  ranges  from  10  m below  Bed  well’s  Columnar  Band,  up 
to  the  base  of  the  Uintacrinus  Zone  (text-fig.  1).  It  is  only  common  2-4  m below  Whitaker’s  3-inch 
band,  where  it  occurs  in  clusters.  I.  infulasteroides  is  also  known  from  the  upper  coranguinum  Zone 
of  Berkshire,  Hampshire,  Sussex  and  the  Isle  of  Wight,  and  the  Uintacrinus  Zone  of  Hampshire. 
Ernst  and  Schulz  (1971)  record  this  species  from  the  Middle  Santonian  of  Lagerdorf. 

Genus  hagenowia  Duncan,  1889 
(=  Martinosigra  Nielsen,  1942) 

Type  species.  Cardiaster  rostratus  Forbes  1852  by  original  designation. 

Diagnosis.  Small  holasterids  in  which  the  apical  part  of  the  test  is  elongated  antero-dorsally  to 
form  a tapering  rostrum.  Narrow,  well-defined  sulcus  runs  from  apex  of  rostrum  to  circular 
peristome.  Apical  system  disjunct,  with  two  posterior  oculars  at  base  of  rostrum  separated  from 
remainder  of  system  (which  is  on  rostral  tip)  by  interambulacra  1 and  4.  Two  or  four  genital  pores. 

Longitudinally  oval  periproct  is  positioned  at  the  summit  of  the  posterior  truncation.  Plastron 
metasternal,  keeled.  Marginal  fasciole. 

Remarks.  The  diagnosis  of  Hagenowia  is  amended  from  that  of  Wright  and  Wright  (1949),  to 
exclude  I.  infulasteroides  from  the  genus.  The  most  important  diagnostic  feature  of  Hagenowia  is 


16 


PALAEONTOLOGY,  VOLUME  25 


III 


text-fig.  3.  Plating  structure  and  cross-sectional  shape  of  the  rostrum  in  Hagenowia  rostrata  (Forbes):  a and 
b,  dorsal  and  lateral  views  of  a small  individual  from  1 m above  Bedwell’s  Columnar  Band,  East  Cliff,  Dover. 
BMNH  E76835,  A.  S.  Gale  Coll,  c,  cross-section  of  the  rostrum,  taken  at  the  3rd  plate  of  interambulacral 
row  1 b.  Specimen  from  1-5  m above  Bedwell’s  Columnar  Band,  Ramsgate.  BMNH  E76836,  A.  S.  Gale  Coll. 
d,  lateral  view  of  specimen  from  2 m above  the  East  Cliff  Semitabular,  East  Cliff,  Dover.  BMNH  E76837, 

A.  S.  Gale  Coll. 


GALE  AND  SMITH:  CRETACEOUS  ECHINOIDS 


17 


the  separation  of  the  apical  system.  Interruption  of  the  plastron  by  ambulacra  I and  V is  seen  in 
all  Hagenowia  and  is  also  present  in  I.  infulasteroides. 

Hagenowia  rostrata  (Forbes,  1852) 

Plate  3,  figs.  1-5;  Plate  4,  fig.  3;  Plate  5,  figs.  1,2,  5;  text-figs,  3,  4 

1852  Cardiaster  rostratus  Forbes,  p.  3,  pi.  10,  figs.  19-24. 

1858  Infulaster  rostatus  (Forbes);  Desor,  p.  348. 

1881  Infulaster  rostratus  (Forbes);  Wright,  p.  307,  pi.  70,  figs.  2,  3. 

1889  Hagenowia  rostrata  (Forbes);  Duncan,  p.  210. 

1942  Martinosigra  rostrata  (Forbes);  Nielsen,  p.  163. 

1949  Hagenowia  rostrata  (Forbes);  Wright  and  Wright,  p.  462,  figs.  7,  8,  1 1,  12,  non  9,  10. 

1971  Hagenowia  rostrata  (Forbes);  Ernst  and  Schulz,  p.  138,  pi.  13,  fig.  5;  pi.  14,  fig.  1;  text-fig.  7. 

Lectotype.  A specimen  figured  by  Forbes  (1852,  pi.  10,  figs  19-21;  IGS  38656)  was  taken  as  lectotype  by 
Wright  and  Wright  (1949).  This  came  from  the  ‘Ctialk  with  Flints  of  Bostal  Heath,  near  Plumstead’  (Forbes 
1852,  p.  3),  in  south-east  London,  presumably  from  the  upper  part  of  the  coranguinum  Zone. 

Diagnosis.  Rostrum  short,  sharply  tapering;  plates  of  ambulacral  rows  Ha,  IVb  contiguous,  not 
reduced;  posterior  side  of  rostrum  broad,  evenly  rounded;  sulcus  deep;  subanal  protruberance 
double,  asymmetrical. 

Remarks.  The  plating  arrangements  of  the  rostrum  is  completely  known  only  in  this  species  of 
Hagenowia  (text-figs.  3,  4).  Oculars  I and  V are  small,  and  positioned  at  the  base  of  the  rostrum. 
Together  with  ambulacra  I and  V,  and  interambulacrum  5,  they  are  separated  from  the  rest  of  the 
apical  system  along  the  dorsal  margin  of  the  rostrum  by  interambulacra  1 and  4 (text-fig.  3).  The 


a 


b 


c 


text-fig.  4.  Plating  structure  of  the  rostrum  tip  in  Hagenowia  rostrata  (Forbes):  a-c.  Dorsal,  lateral,  and  frontal 
views  of  large  specimen  from  3 m below  Whitaker’s  3-inch  band.  North  Foreland,  near  Broadstairs,  Kent. 
BMNH  E76838,  A.  S.  Gale  Coll. 


18 


PALAEONTOLOGY,  VOLUME  25 


first  two  plates  of  interambulacral  rows  la  and  4b  meet  along  the  dorsal  margin  of  the  rostrum 
above  oculars  I and  V.  The  second  plates  are  often  slightly  swollen  in  lateral  profile  and  form  a 
distinct,  low  protuberance  at  the  base  of  the  rostrum  (e.g.  PI.  3,  fig.  2b).  la  and  4b  are  occluded 
from  the  rostrum  above  this  level  by  two  or  three  plates  of  the  outer  rows,  lb  and  4a. 

Variations  in  test  shape  of  H.  rostrata  is  shown  by  a group  of  specimens  from  the  upper 
coranguinum  Zone  of  the  Kent  coast  (PI.  3,  figs.  1-5).  Low,  depressed  forms  with  relatively  high 
anterior  angles  (PI.  3,  fig.  1)  intergrade  continuously  through  to  tall  individuals  with  steep  sides 
and  more  upright  rostra  (PI.  3,  fig.  2).  No  stratigraphical  separation  of  these  forms  is  known  to 
occur.  Small  individuals  of  H.  rostrata  have  shorter,  less  well  demarcated  rostra  (PI.  3,  fig.  4). 

Ernst  and  Schulz  (1971,  p.  138,  text-fig.  7,  fig.  1)  recorded  a stratigraphically  low  ‘early  form’ 
of  H.  rostrata  from  the  (?  Lower)  Coniacian  of  Lagerdorf,  north-west  Germany.  This  has  a short 
posterior  slope,  a short  rostrum,  and  a flat  profile  to  the  base.  A single  specimen  (PI.  3,  fig.  5) 
from  the  lowest  horizon  at  which  Hagenowia  occurs  at  Dover,  some  2 m above  the  East  Cliff 
Semitabular  flint  (text-fig.  1)  compares  quite  well  in  lateral  profile  with  the  figured  Lagerdorf 
individual.  The  Dover  specimen  differs  from  other  English  examples  of  H.  rostrata  in  having  a 
broad,  relatively  shallow  sulcus  without  narrow  margins.  In  anterior  profile,  the  sides  of  the  test 
do  not  inflect  sharply  as  the  body  passes  into  the  rostrum. 

Occurrence.  On  the  Kent  coast  (text-fig.  1)  H.  rostrata  is  found  in  a succession  of  discrete  bands 
in  the  coranguinum  Zone,  from  the  East  Cliff  Semitabular  to  Whitaker’s  3-inch  band.  It  occurs  in 
the  same  zone  throughout  its  outcrop  in  southern  England,  although  precise  details  of  horizon 
are  seldom  recorded.  In  northern  France  a solitary  example  was  found  3 m below  the  equivalent 
of  Whitaker’s  3-inch  band  at  Coquelles,  near  Calais  (A.  S.  G.  Coll.).  In  Yorkshire,  Rowe  (1904) 
used  H.  rostrata  as  a local  index  for  the  M.  coranguinum  Zone,  which  was  subsequently  formalized 
by  Wright  and  Wright  (1942).  Recent  study  has  shown  that  the  species  of  Hagenowia  common  in 
the  flintless  chalk  of  the  Yorkshire  coast  below  the  entry  of  Uintacrinus  is,  in  fact,  H.  anterior. 
The  only  true  H.  rostrata  we  have  seen  from  Y orkshire  are  from  the  flinty  chalk  with  Inoceramus 
involutus  of  Little  Weighton  (Wrights’  Coll.).  In  north-west  Germany  H.  rostrata  occurs  in  the 
Coniacian  Chalk  at  Lagerdorf  (Ernst  and  Schulz  1971,  text-fig.  5;  1974,  text-fig.  4 a). 


EXPLANATION  OF  PLATE  3 

Figs.  1 -5.  Hagenowia  rostrata  (Forbes).  Micraster  coranguinum  Zone,  Kent  coast.  1 a-c,  oral,  dorsal,  and  lateral 
views  of  depressed  specimen  from  3 m beneath  Whitaker’s  3-inch  band,  North  Foreland  near  Broadstairs. 
BMNH  E76838,  A.  S.  Gale  Coll.  2a,  b,  oral  and  lateral  views  of  high  individual,  with  more  vertical  rostrum, 
retaining  radioles  on  the  base  from  2 m above  Bedwell’s  Columnar  Band,  same  locality.  BMNH  E76848,  A.  S. 
Gale  Coll.  3,  lateral  view  of  specimen,  same  horizon  and  locality.  BMNH  E76849,  A.  S.  Gale  Coll.  4,  small 
individual  with  short  rostrum,  1 m above  Bedwell’s  Columnar  Band,  East  Cliff,  Dover.  BMNH  E76835,  A.  S. 
Gale  Coll.  5,  lateral  view  of  specimen  with  poorly  demarcated  rostrum.  2 m above  East  Cliff  Semitabular,  East 
Cliff,  Dover.  BMNH  E76850,  A.  S.  Gale  Coll. 

Fig.  6a,  b.  H.  anterior  Ernst  and  Schulz.  Oral  and  lateral  views  of  body.  The  rostrum  has  broken  off,  and  its  base 
is  bored.  Uintacrinus  Zone,  Harding’s  Whiting  Pits,  Devizes  Road,  Salisbury  Wilts.  BMNH  E35788, 
Blackmore  Coll. 

Figs.  7,  8.  H.  blackmorei  Wright  and  Wright,  la-c,  lateral,  frontal,  and  oral  views  of  the  holotype  (body  only). 
The  rostrum  belongs  to  H.  anterior  and  has  been  artificially  attached.  Lower  Campanian,  probably  lower  G.  | 
quadrata  Zone,  West  Harnham,  near  Salisbury,  Wilts.  BMNH  E33916,  Blackmore  Coll.  8a,  b,  lateral  and 
frontal  views  of  rostrum  from  Hagenowia  horizon,  lower  G.  quadrata  Zone,  pit  no.  3 of  Gaster  (1924),  North 
Lancing,  near  Worthing,  Sussex.  I.G.S.  no.  Zm  2907. 

All  specimens  x 2.  Coated  with  ammonium  chloride. 


PLATE  3 


1 

5 IIEP* 

II 

| -mm 

k 8 , 

GALE  and  SMITH,  Cretaceous  irregular  echinoids 


20 


PALAEONTOLOGY,  VOLUME  25 


Hagenowia  anterior  Ernst  and  Schulz,  1971 
Plate  3,  fig.  6;  Plate  4,  figs.  5,  10,  12;  Plate  5,  figs.  6,  8;  Plate  6,  figs.  2,  6,  8;  text-fig.  5 
1949  Hagenowia  rostrata  (Forbes);  Wright  and  Wright,  p.  462,  figs.  9,  13. 

1971  Hagenowia  blackmorei  Wright  and  Wright,  anterior  Ernst  and  Schulz,  p.  140,  pi.  13,  fig.  6; 
pi.  14,  figs.  2,  3;  text-fig.  8. 

Types.  The  holotype  is  from  the  Middle  Santonian  rogalae-westfalica  Zone  of  Lagerdorf,  north-west  Germany 
(Ernst  and  Schulz  1971,  text-fig.  8,  fig.  1;  pi.  14,  figs.  2,  3).  The  paratypes  are  from  the  same  horizon  and  locality. 

Diagnosis.  Rostrum  long  and  slender,  cross-section  triangular,  posterior  side  evenly  rounded; 
individual  plates  of  ambulacral  rows  Ha,  IVb  small,  separated  by  interambulacral  rows  la,  4b  and 
ambulacra!  rows  lib,  IVa  in  the  rostrum;  subanal  protruberance  single,  narrow. 

Remarks.  The  plating  arrangement  of  all  but  the  basal  rostrum  is  known  in  this  species  (text-fig. 
5).  Individual  plate  rows  are  proportionately  longer  and  narrower  than  in  H.  rostrata.  At  least  six 
plates  of  interambulacral  rows  la,  4b  meet  along  the  dorsal  margin  of  the  rostrum  (text-fig.  5b); 
these  are  variable  in  width.  Plates  of  ambulacral  rows  Ha  and  IVb  are  small,  and  although  the 
first  few  are  usually  in  contact  with  each  other,  most  are  separated  by  la  and  4b,  and  lib,  IVa 
(text-fig.  5a-c,  e).  The  size  and  shape  of  these  occluded  plates  is  variable.  In  some  individuals  (e.g. 
text-fig.  5e)  plates  of  Ila  and  IVb  are  locally  separated  by  interambulacra  la,  4b  and  2a,  3b. 
Genitals  1 and  4 are  absent. 


text-fig.  5.  Plating  structure  and  cross-sectional  shape  of  the  rostrum  in  Hagenowia  anterior  Ernst  and  Schulz: 
a , dorsal  view  of  rostrum  from  5 m above  Whitaker’s  3-inch  band,  Kingsgate,  Kent.  BMNH  E76839,  A.  S.  Gale 
Coll,  b,  dorsal  view  of  rostrum  from  basal  1 m.  of  Marsupites  Zone,  Minnis  Bay,  Kent.  BMNH  E76840,  A. 
S.  Gale  Coll,  c,  lateral  view  of  rostrum  tip,  mid  Uintacrinus  Zone,  near  Margate,  Kent.  BMNH  E76841,  A. 
S.  Gale  Coll,  d,  cross-section  of  rostrum,  taken  at  2nd/3rd  plate  of  interambulacral  row  1 b,  mid  Uintacrinus 
Zone,  near  Margate,  e,  lateral  view  of  rostrum,  base  of  Uintacrinus  anglicus  band,  Foreness  Point,  near 
Margate.  BMNH  E76842,  A.  S.  Gale  Coll. 


GALE  AND  SMITH:  CRETACEOUS  ECHINOIDS 


2 


Some  variation  in  the  attitude  of  the  rostrum  occurs  in  M.  anterior , although  the  body  shape 
is  quite  consistent.  Lateral  furrows  on  the  sides  of  the  test,  just  anterior  to  the  periproct  (Ernst 
and  Schulz  1971,  text-fig.  8,  figs.  2,  3)  are  sometimes  present.  H.  anterior  is  raised  to  specific  rank 
on  account  of  significant  differences  in  rostral  structure  and  body  shape  from  H.  blackmorei. 

Occurrence.  In  east  Kent  the  species  first  appears  at  3-5  m above  Whitaker’s  3-inch  band  in  the 
coranguinum  Zone  and  ranges  up  into  the  Uintacrinus  anglicus  band  in  the  lower  Offaster  pilula 
Zone  (text-fig.  1).  Levels  of  abundance  occur  in  the  mid  U.  socialis  Zone  and  in  the  basal  Marsupites 
Zone.  Scattered  records  of  this  species  exist  from  Dorset,  Wiltshire,  Hampshire,  and  the  Isle  of 
Wight  ( Uintacrinus  Zone).  In  Yorkshire  H.  anterior  occurs  commonly  in  the  flintless  upper  part 
of  the  coranguinum  Zone  and  ranges  up  into  the  lower  Campanian  I.  lingua  Zone.  At  Lagerdorf 
the  species  ranges  from  the  mid  rogalae-westfalica  Zone  to  the  Marsupites  Zone,  with  one  doubtful 
earlier  record  from  the  coranguinum-westfalica  zone  (Ernst  and  Schulz  1971,  1974). 

Hagenowia  blackmorei  Wright  and  Wright,  1949 
Plate  3,  figs.  7,  8;  Plate  4,  fig.  6;  Plate  6,  figs.  4,  7,  9;  text-fig.  6 
1949  Hagenowia  blackmorei  Wright  and  Wright,  pp.  467-70,  figs.  14-16. 

1971  Hagenowia  blackmorei  blackmorei  Wright  and  Wright;  Ernst  and  Schulz,  p.  140,  text-fig.  2. 

Types.  The  specimen  chosen  as  holotype  by  Wright  and  Wright  (1949,  fig.  14)  is  from  the  Lower  Campanian 
Chalk  of  East  Harnham,  near  Salisbury,  Wiltshire  (BMNH  E33916,  Blackmore  Coll.),  probably  from  the  G. 
quadrata  Zone.  However,  as  pointed  out  by  C.  J.  Wood  (in  Ernst  and  Schulz  1971),  the  body  and  rostrum 
belong  to  two  specimens  of  separate  species,  and  have  been  artificially  joined.  The  body  belongs  to  the  species 


text-fig.  6.  Plating  structure  and  cross-sectional  shape  of  the  rostrum  in 
Hagenowia  blackmorei  Wright  and  Wright:  a,  b,  dorsal  and  lateral  views  of 
rostrum,  based  on  specimen  from  10  m.  above planoconvexa  bed,  G.  quadrata 
Zone,  cliffs  west  of  Newhaven,  Sussex.  BMNH  E76843,  A.  S.  Gale  Coll. 
c,  cross-section  of  rostrum  3 mm  below  tip,  same  horizon  and  locality. 


22 


PALAEONTOLOGY,  VOLUME  25 


characteristic  of  the  lower  quadrata  Zone  of  southern  England,  and  possesses  the  features  of  H.  blackmorei 
as  described  by  Wright  and  Wright.  The  rostrum  is  from  an  individual  of  H.  anterior  from  the  Santonian. 
The  body  alone  is  therefore  selected  as  lectotype.  The  paratype  is  an  internal  and  external  mould  in  flint  from 
the  Haldon  Gravel,  Devon  (Wright  and  Wright  1949,  figs.  14,  15;  BMNH  E8404,  Vicary  Coll.). 

Diagnosis.  Rostrum  slender,  upright,  well  demarcated  from  body,  equilaterally  triangular  in 
cross-section;  interambulacral  rows  la,  4b  form  narrow  dorsal  ridge  on  rostrum,  surface  of 
ambulacral  rows  lib,  IVa  slightly  inset;  sulcus  shallow,  with  a gently  concave  floor;  Ha,  IVb  not 
present  in  rostrum;  body  squarish  in  lateral  profile,  anterior  part  of  base  vertically  below  peristome; 
subanal  protruberance  single,  small;  posterior  slope  steep;  fasciole  noticeably  oblique. 

Remarks.  The  structure  of  the  rostrum  in  this  species  (text-fig.  6)  is  similar  to  that  in  H.  elongata 
(Nielsen)  (Schmid  1972)  but  individual  plates  are  less  elongated  and  broader  than  in  the 
Maastrichtian  species.  The  dorsal  ridge  on  the  rostrum  is  broader  and  less  sharply  defined  than 
in  H.  elongata,  and  genitals  2,  3 are  not  usually  separated  from  oculars  II,  IV  in  H.  blackmorei, 
as  is  invariably  so  in  H.  elongata.  The  latter  species  has  only  two,  vertically  situated,  madreporic 
pores,  whereas  the  number  and  arrangement  of  these  is  variable  in  H.  blackmorei. 

A few  Hagenowia  rostra  are  known  from  the  Upper  Campanian  Belemnitella  mucronata  Zone 
of  southern  England  (Wright  and  Wright  1949).  These  show  affinities  with  both  H.  blackmorei 
and  H.  elongata,  but  lack  the  specialized  madreporic  pores  of  the  latter  and  are  referred  to  H.  cf. 
blackmorei. 

Occurrence.  H.  blackmorei  appears  to  be  restricted  to  the  lower  part  of  the  G.  quadrata  Zone  in 
southern  England.  Gaster  (1924)  used  the  term  Hagenowia  horizon  for  this  level  in  Sussex.  Collecting 
from  the  cliffs  west  of  Newhaven  in  Sussex  suggests  that  the  species  occurs  only  within  the  12  m 
of  chalk  above  the  marl  pair  (planoconvexa  bed)  taken  by  Brydone  (1914)  as  the  base  of  the  G. 
quadrata  Zone.  The  species  is  only  abundant  in  the  interval  from  9-11  m above  this  bed,  where 
approximately  fifty  specimens  have  been  found.  Blackmore’s  material  from  East  Harnham  lacks 
detailed  stratigraphical  location,  but  probably  came  from  the  correlative  horizon. 


Hagenowia  elongata  (Nielsen,  1942) 

Plate  4,  figs.  4,  7-9;  Plate  5,  figs.  7,  9;  Plate  6,  figs.  1,  3,  10 

1942  Martinosigra  elongata  Nielsen,  p.  163,  fig.  2. 

1949  Hagenowia  rostrata  (Forbes);  Wright  and  Wright,  p.  462. 

1950  Hagenowia  elongata  (Nielsen);  Mortensen,  p.  97,  fig.  100. 

1971  Hagenowia  elongata  (Nielsen);  Schmid,  in  Ernst  and  Schulz,  p.  141,  text-fig.  2,  el. 

1972  Hagenowia  elongata  (Nielsen);  Schmid,  p.  179,  pis.  1-4;  text-figs.  1,  2. 

Diagnosis.  Plate  rows  in  rostrum  very  long  and  narrow;  genitals  2 and  3 separated  from  oculars 
II  and  IV  by  interambulacral  rows  2a,  3b;  madreporite  with  only  two,  vertically  situated,  pores; 
sulcus  shallow,  flat. 

Remarks.  Now  that  the  rostral  structures  of  preceding  species  of  Hagenowia  are  known,  it  is 
necessary  to  modify  the  interpretation  of  this  species  from  Schmid  (1972).  Instead  of  representing 
both  plate  rows  of  ambulacra  II  and  IV,  the  four  rows  of  plates  on  the  dorsal  side  of  the  rostrum 
are  lib,  IVa  (outer  rows)  and  lb,  4a  (inner  rows).  There  is  little  else  to  add  to  Schmid’s  excellent 
description  of  the  species.  The  outline  of  a body  stated  to  belong  to  this  species  was  shown 
by  Ernst  and  Seibertz  (1977,  text-fig.  6),  but  this  new  material  awaits  detailed  description  and 
figuring. 

Occurrence.  Upper  Lower  Maastrichtian,  Zone  of  Belemnella  occidentalis,  Denmark  and  north-west 
Germany. 


GALE  AND  SMITH:  CRETACEOUS  ECHINOIDS 


23 


PHYLOGENY 

Wright  and  Wright  (1949)  suggested  that  I.  infulasteroides  arose  from  the  ‘small  I excentricus ’ of 
the  cortestudinarium  Zone  (here  I.  tuberculatus ) and  subsequently  gave  rise  to  H.  rostrata.  They 
regarded  H.  blackmorei  as  a short-lived  offshoot  from  H.  rostrata,  which  itself  survived  into  the 
Upper  Campanian.  Ernst  and  Schulz  (1971,  p.  141,  text-fig.  2)  modified  this  story  only  slightly, 
by  placing  the  origin  of  H.  rostrata  near  the  point  at  which  an  infulasteroides  lineage  diverged 
from  the  main  Infulaster  stock.  They  retained  separate  rostrata  and  blackmorei  lineages  throughout 
the  Santonian  and  Campanian.  More  recently,  Ernst  and  Seibertz  (1977,  p.  563,  text-fig.  6)  showed 
the  H.  rostrata  lineage  as  terminating  within  the  Middle  Santonian. 

It  is  suggested  here  that  only  two  lineages  are  recognizable  in  this  group  of  echinoids. 

1.  Infulaster  lineage:  I.  excentricus  ( sensu  lato)  I.  tuberculatus— I.  infulasteroides 

2.  Hagenowia  lineage:  H.  rostrata — H.  anterior — H.  blackmorei — H.  elongata. 

Since  Hagenowia  appeared  before  I.  tuberculatus  gave  rise  to  I.  infulasteroides,  it  is  likely  that  the 
origins  of  this  genus  lay  in  the  former  species,  probably  during  the  mid-Coniacian.  In  this  case, 
I.  infulasteroides  parallels  rather  than  antecedes  Hagenowia  in  the  development  of  an  acutely  angled 
apex  and  an  interrupted  plastron. 

Although  the  over-all  changes  in  each  of  the  lineages  are  progressive,  individual  species  are  in 
general  well  demarcated  from  both  ancestor  and  descendant.  No  two  species  in  either  lineage  are 
known  definitely  to  overlap  in  stratigraphical  order. 


EVOLUTIONARY  TRENDS 

1.  Size.  In  the  Infulaster  lineage  there  is  a dramatic  decrease  in  size  from  I.  excentricus,  which  is 
characteristically  between  40  and  50  mm  in  length,  to  I.  tuberculatus  (15-25  mm).  With  this  change, 
the  test  became  proportionately  shorter  and  deeper.  In  late  I.  tuberculatus  and  in  I.  infulasteroides 
the  apex  of  the  test  became  more  acutely  angled  resulting  from  an  increase  in  the  length  of  plate 
rows  on  the  sides  of  the  test. 

2.  Rostrum.  In  Hagenowia  the  apical  system  is  separated  into  two  parts  to  form  a trivium  and 
bivium.  This  was  a direct  consequence  of  the  considerable  elongation  of  the  apical  region. 

In  the  earliest  species,  H.  rostrata,  oculars  I and  V are  separated  from  genitals  1 and  4 by 
interambulacral  rows  1 and  4 along  the  dorsal  side  of  the  rostrum.  Plate  rows  la  and  4b  are 
occluded  from  the  distal  rostrum  by  rows  lb  and  4a,  which  meet  along  the  dorsal  mid-line.  Rostral 
plates  are  elongated,  the  sulcus  is  deep,  and  the  body  cavity  within  the  rostrum  is  large.  In  profile, 
the  rostral  tip  is  only  slightly  enlarged. 

In  H.  anterior  the  rostrum  is  better  demarcated  from  the  body  and  is  relatively  narrow.  Plate 
rows  are  proportionally  narrower  and  more  plates  per  row  are  present  in  the  rostrum.  Plates  of 
ambulacral  rows  Ila  and  IVb  are  typically  diminutive  while  genitals  1 and  4 have  been  lost.  In 
comparison  with  H.  rostrata,  the  rostrum  of  H.  anterior  is  more  equilaterally  triangular  in 
cross-section  and  the  rims  of  the  sulcus,  formed  by  interambulacral  rows  2b  and  3a,  are  distinctly 
thickened.  The  walls  of  the  rostrum  are  also  proportionally  thicker  and  the  body  cavity  smaller. 
The  sulcus  is  significantly  shallower. 

The  trend  towards  increasing  slenderness  and  sharper  demarcation  of  the  rostrum  is  continued 
in  H.  blackmorei.  Ambulacral  rows  Ila  and  IVb  are  lost  completely  from  the  part  of  the  rostrum 
where  the  plating  arrangement  is  known.  Rostral  plates  are  more  elongate  than  in  previous  species. 
In  cross-section,  the  rostrum  is  triangular  and  thickened  at  each  corner  while  the  body  cavity  is 
further  reduced.  Ambulacral  plate  rows  Ila  and  IVb  are  slightly  depressed.  The  sulcus  is  shallower 
than  in  H.  anterior  and  its  rims  are  parallel.  The  tip  of  the  rostrum  is  broader  than  the  shaft  and 
the  rostrum  is  more  or  less  vertical  in  attitude. 

Rostral  development  is  most  pronounced  in  H.  elongata.  Rostral  plates  are  narrower  and  more 
elongate  than  in  any  of  its  predecessors.  Interambulacral  plate  rows  2a  and  3b  always  separate 


24  PALAEONTOLOGY,  VOLUME  25 

genitals  2 and  3 from  oculars  II  and  IV.  Only  two  vertically  placed  madreporic  pores  are 
present. 

Summarizing  these  trends,  there  is  a progressive  increase  in  the  slenderness  and  clearer 
demarcation  of  the  rostrum.  This  is  achieved  through  elongation  and  narrowing  of  plate  rows, 
increasing  the  number  of  plates  in  each  row  present,  and  loss  of  plate  rows.  The  rostrum  adopts 
a more  vertical  attitude  and  is  strengthened  by  the  development  of  buttresses  and  by  thickening  of 
the  walls.  The  body  cavity  within  the  rostrum  is  progressively  reduced  and  genitals  1 and  4 are 
lost  during  evolution. 

3.  Body.  The  body  of  H.  rostrata  differs  little  from  that  of  the  ancestral  Infulaster.  During  the 
evolution  of  Hagenowia  there  were  five  important  changes  to  the  structure  of  the  body.  Firstly, 
the  plastron  became  interrupted  in  I.  infulasteroides  and  in  all  species  of  Hagenowia.  Secondly,  the 
base  became  progressively  more  convex.  Thirdly,  the  peristome  moved  from  a basal  to  an  anterior 
position  with  a vertical  attitude.  Fourthly,  the  posterior  slope  became  more  vertical  and  the  fasciole 
more  oblique.  Finally  there  was  a change  from  a double,  asymmetrical  sub-anal  protruberance  in 
H.  rostrata  to  a single  centrally  placed  protruberance  in  later  species. 

4.  Pore  morphology.  The  more  important  changes  in  pore  morphology  during  the  evolution  of 
Infulaster  and  Hagenowia  include  the  progressive  loss  of  phyllode  and  dorso-lateral  pores  and  the 
increasing  differentiation  of  pores  at  the  rostral  tip.  Three  distinct  regions  of  pores  are  recognizable 
and  will  be  dealt  with  separately.  Pore  terminology  is  taken  from  Smith  (1980a). 

(a)  Phyllode  pores.  In  Infulaster,  phyllode  pores  are  relatively  large  isopores,  rounded  in  outline, 
with  an  axially  positioned  neural  canal  in  an  adoral  position.  They  are  300-350  in  length  in  I. 
excentricus  but  only  180-250  /z m in  length  in  later  species.  Thirty  such  pores  lie  around  the  peristome 
in  I.  excentricus  but  only  20  to  22  are  present  in  I.  tuberculatus  and  I.  infulasteroides.  Similar 
phyllode  pores  are  found  in  H.  rostrata  and  H.  anterior.  H.  rostrata  has  between  14  and  18  such 
pores  while  H.  anterior  has  slightly  fewer  (the  exact  number  cannot  be  ascertained  in  any  of  the 
specimens  examined).  H.  blackmorei  has  only  two  small,  circular  unipores  lacking  any  periporal 
ornament.  These  lie  on  either  side  of  the  anterior  sulcus  immediately  above  the  peristome.  There 
is  no  sign  of  pores  in  lateral  and  posterior  ambulacra  adjacent  to  the  peristome. 

( b ) Latero-dorsal  pores.  In  Infulaster,  isopores  of  the  posterior  columns  of  ambulacra  II  and  IV 
are  twice  as  long  as  other  latero-dorsal  pores.  These  elongate  isopores  have  no  obvious  neural 


EXPLANATION  OF  PLATE  4 
Bottom  of  micrograph  adoral  unless  otherwise  stated. 

Fig.  1.  Infulaster  infulasteroides:  latero-dorsal  partitioned  isopore  from  interambulacrum  5.  BMNH  E76851. 
Fig.  2.  I.  tuberculatus:  partitioned  isopore  with  axially  positioned  neural  canal  near  the  apex  of  ambulacrum  III. 
BMNH  E76852. 

Fig.  3.  Hagenowia  rostrata:  partitioned  isopore  with  axially  positioned  neural  canal  near  the  apex  of 
ambulacrum  III.  BMNH  E7684. 

Fig.  4.  H.  elongata:  simple  unipore  of  ambulacrum  III  within  the  rostral  sulcus.  MM  12820. 

Fig.  5.  H.  anterior:  unipore  with  broad  periporal  area  in  ambulacrum  III  at  the  rostral  head.  Adoral  to  the  right. 
BMNH  E76844. 

Fig.  6.  H.  blackmorei:  unipore  with  broad  periporal  area  in  ambulacrum  III  at  the  rostral  head.  BMNH  E76846. 
Fig.  7.  H.  elongata:  unipore  in  ambulacrum  III  at  the  rostral  head. 

Fig.  8.  H.  elongata:  tuberculation  on  the  lateral  face  of  the  rostrum.  Apex  to  the  left,  dorsal  ridge  at  top. 
Fig.  9.  H.  elongata:  enlargement  of  fig.  8 showing  one  tubercle. 

Fig.  10.  H.  anterior:  dorsal  tubercle  on  rostrum  showing  enlarged  areole  and  radially  symmetrical  crenulation. 
Fig.  11.  /.  tuberculatus:  latero-dorsal  tubercles. 

Fig.  12.  H.  anterior:  latero-dorsal  tubercles  at  the  base  of  the  rostrum. 

Scale  bar  in  figs.  1-4,  6,  7,  9,  12  = 100  ^m;  figs.  5,  8,  10,  1 1 = 200  Mm. 


PLATE  4 


GALE  and  SMITH,  Cretaceous  irregular  echinoids 


26 


PALAEONTOLOGY,  VOLUME  25 


canal  and  the  interporal  partition  is  more  or  less  flush  with  the  test  surface  and  may  be  very  broad. 
Other  dorso-lateral  pores  are  partitioned  isopores  with  a small,  laterally  positioned  neural  canal 
(PI.  4,  fig.  1).  All  isopores  are  widely  spaced  and  have  no  recognizable  attachment  area.  Pores  of 
I.  excentricus  are  approximately  twice  the  size  of  pores  in  later  species. 

Latero-dorsal  pores  vary  tremendously  in  specimens  of  H.  rostrata.  In  some  of  the  larger  specimens 
there  are  isopores,  arranged  as  in  Infulaster.  One  of  these  has  elongate  isopores  of  ambulacra  lib 
and  IVa  which  are  even  larger  than  those  in  I.  excentricus.  The  majority  of  specimens  lack  such 
elongate  isopores  and  have  small  partitioned  isopores  in  all  columns.  In  other  specimens,  all 
latero-dorsal  pores  are  reduced  to  unipores.  Finally,  one  specimen  has  unipores  in  ambulacral 
columns  lib  and  IVa  and  small  partitioned  isopores  in  columns  Ila  and  IVb.  In  H.  anterior  there 
are  no  pores  in  rostral  plates  of  ambulacra  II  and  IV  but  there  are  between  three  and  five  partitioned 
isopores  in  each  column  on  the  body  above  the  fasciole.  There  are  no  latero-dorsal  pores  in  H. 
blackmorei. 

( c ) Ambulacrum  III  pores.  Pores  in  the  frontal  sulcus  are  similar  in  size  and  shape  in  the  three 
species  of  Infulaster  and  in  H.  rostrata.  These  are  partitioned  isopores  with  axially  aligned  neural 
canals  (PI.  4,  figs.  2,  3).  Pores  near  the  apex  are  identical  with  those  within  the  sulcus.  In  later 
species  of  Hagenowia,  pores  in  the  sulcus  are  smaller  than  those  at  the  apex.  The  12  to  14  most 
adapical  pores  in  H.  anterior  and  H.  blackmorei  are  unipores,  circular  in  outline,  and  with  a narrow 
central  pore  surrounded  by  a clear  periporal  area  (PI.  4 figs.  5,  6).  Unipores  in  the  sulcus  are 
smaller  with  a reduced  periporal  area.  In  the  apical  unipores  of  H.  elongata,  the  central  pore  is 
much  larger  and  most  of  the  periporal  area  is  lost  (PI.  4,  fig.  7).  Unipores  in  the  sulcus  are  minute 
and  lack  periporal  ornament  (PI.  4,  fig.  4). 


text-fig.  7.  Tubercle  arrangement  in  Infulaster  excentricus.  a — anterior:  b — lateral:  c — posterior:  d — aboral: 
e— oral.  Arrows  on  the  left-hand  side  of  diagrams  show  the  direction  of  areole  enlargement.  Circles  indicate 
radially  symmetrical  tubercles.  Stippling  on  the  right-hand  side  of  diagrams  indicates  the  size  and  arrangement 
of  tubercles.  The  marginal  fasciole  appears  as  a black  band. 


GALE  AND  SMITH:  CRETACEOUS  ECHINOIDS 


27 


5.  Tubercle  arrangement.  Although  some  marked  changes  occur  in  the  detailed  arrangement  of 
tubercles  during  the  evolution  of  Hagenowia,  all  functionally  distinct  groups  of  tubercles  were 
already  developed  in  I.  ex centricus.  The  more  important  changes  are  outlined  below. 

(i)  There  is  a slight  reduction  in  the  area  of  plastron  tubercles  in  later  species  of  Hagenowia. 
This  is  accompanied  by  a corresponding  increase  in  the  tubercle-free  ambulacral  zones  on  either 
side  of  the  plastron.  The  plastron  is  extremely  narrow  in  H.  blackmorei  (text-fig.  8f). 

(ii)  There  is  a relative  increase  in  the  area  of  ventro-lateral  tubercles,  which  are  best  developed 
in  H.  blackmorei.  This  is  a consequence  of  the  increasing  obliquity  of  the  fasciole  (which,  in  turn, 
arose  from  the  progressive  shift  of  the  peristome). 

(iii)  Tubercles  in  the  sub-anal  region  are  radially  arranged  around  two  points  in  Infulaster  and 
H.  rostrata  (text-figs.  7,  8).  In  H.  anterior  and  H.  blackmorei , sub-anal  tubercles  are  radially  arranged 
around  a single  point  (text-fig.  8i). 


text-fig.  8.  Tubercle  arrangement  in  Hagenowia.  a-e,  H.  rostrata-.  a— oral;  b— anterior; 
c — posterior;  d— aboral;  e— lateral,  f-j,  H.  blackmorei.  f — oral;  G — anterior;  h— lateral;  i— posterior; 
J — arrangement  at  the  base  of  the  sulcus  and  surrounding  the  peristome.  For  explanation  see 

text-fig.  7. 


28 


PALAEONTOLOGY,  VOLUME  25 


(iv)  Latero-dorsal  tubercles  in  Infulaster  have  an  areole  which  is  radially  symmetrical  or  only 
weakly  enlarged  adapical  (PI.  4,  fig.  11).  In  Hagenowia  the  areole  becomes  increasingly  enlarged 
in  an  adapical  direction,  whilst  crenulation  is  enlarged  on  the  opposite  side  (PI.  4,  figs.  8-10,  12;  text- 
fig.  8).  Latero-dorsal  tubercles  are  least  dense  immediately  above  the  marginal  fasciole  and  become 
progressively  denser  towards  the  dorsal  ridge.  Tubercle  density  over  the  dorsal  surface  increases 
progressively  during  evolution.  There  are  between  1 and  3 tubercles  per  mm2  in  7.  excentricus  and 
I.  tuber culatus  and  this  increases  to  4 or  5 tubercles  per  mm2  in  7.  infulasteroides.  A similar  tubercle 
density  is  found  laterally  at  the  base  of  the  rostrum  in  species  of  Hagenowia.  On  the  rostrum  itself, 
tubercle  density  increases  to  8 to  10  tubercles  per  mm2  in  77.  blackmorei  and  as  much  as  15  tubercles 
per  mm2  in  77.  elongata  where  there  are  four  closely  packed  rows  of  tubercles. 

(v)  The  parts  of  interambulacra  2 and  3 which  form  the  outer  zone  of  lateral  walls  of  the  anterior 
sulcus  have  relatively  small  and  radially  symmetrical  tubercles  (PI.  5,  figs.  3-9).  There  are  three 
rather  irregular  rows  of  these  tubercles  above  the  marginal  fasciole  in  7.  excentricus.  In  7.  tuberculatus 
and  7.  infulasteroides  the  interambulacral  zones  are  narrower  and  there  is  only  a single  rather 
irregular  row  of  tubercles  marginal  to  the  sulcus  (PI.  5,  fig.  3).  In  77.  rostrata  these  tubercles  are 
linearly  arranged  along  the  sulcus  lip  and  each  is  separated  from  its  neighbours  by  a single  row 
of  miliaries  (PI.  5,  fig.  5).  A similar  arrangement  is  found  in  later  species  of  Hagenowia  though 
without  the  intervening  miliaries  (PI.  6,  figs.  1-4). 

(vi)  Tubercle  density  progressively  increases  on  the  lateral  and  posterior  edges  of  the  peristome 
in  Hagenowia.  They  are  best  developed  in  77.  blackmorei  where  they  form  a dense  U-shaped  band 
at  the  base  of  the  sulcus  (text-fig.  8j). 


EXPLANATION  OF  PLATE  5 

Bottom  of  micrograph  adoral  unless  otherwise  stated. 

Fig.  1.  Hagenowia  rostrata : ventral  view  of  rostrum  apex  with  ambulacrum  III  on  the  right.  The  large 
interambulacral  tubercles  at  the  apex  face  outwards  and  their  spines  obviously  did  not  converge.  BMNH 
E76845. 

Fig.  2.  H.  rostrata:  dorsal  view  of  rostrum  apex.  The  large  interambulacral  tubercles  form  a collar  around  the 
frontal  sulcus.  BMNH  E76845. 

Fig.  3.  Infulaster  tuberculatus-.  ambulacrum  III  towards  the  apex.  Large  apical  interambulacral  tubercles  set 
perpendicular  to  the  anterior  ambulacrum  can  be  seen  in  the  top  left  corner.  Smaller  interambulacral 
tubercles,  adjacent  to  ambulacrum  III,  are  irregularly  arranged  and  their  spines  would  not  have  formed  a 
protective  arch  above  the  ambulacrum.  BMNH  E76852. 

Fig.  4.  7.  tuberculatus-.  enlargement  of  fig.  3 showing  the  relatively  dense  arrangement  of  variably  sized  tubercles 
in  ambulacrum  III. 

Fig.  5.  H.  rostrata:  stereo  view  of  the  rostral  sulcus  slightly  below  the  apex.  Ambulacrum  III  is  deeply  sunken 
and  has  denser  and  more  uniformly  sized  tuberculation  than  7.  tuberculatus.  Interambulacral  tubercles 
bordering  the  sulcus  are  arranged  in  two  rather  irregular  rows,  an  inner  row  of  small  tubercles,  that  would 
have  supported  spines  forming  a protective  arch,  and  an  outer  row  of  more  laterally  facing  tubercles  for 
excavating  spines.  BMNH  E76845. 

Fig.  6.  H.  anterior,  ventral  view  of  rostral  sulcus  about  mid-length,  apex  to  the  right.  Ambulacrum  III  is 
moderately  sunken  and  rather  sparsely  covered  in  uniformly  sized  miliary  tubercles.  There  are  two  well- 
defined  rows  of  tubercles  on  the  adjacent  interambulacra,  an  inner  row  of  small  tubercles  for  spines  of  the 
protective  arch  and  an  outer  row  of  larger,  laterally  facing  tubercles  for  excavating  spines.  BMNH  E76844. 

Fig.  7.  H.  elongata : oblique  ventral  view  of  rostral  sulcus,  apex  to  right.  Ambulacrum  III  is  only  slightly  sunken 
and  the  two  rows  of  adjacent  interambulacral  tubercles  are  well  developed  and  set  at  different  angles  to  the 
sulcus.  MM  12820. 

Fig.  8.  H.  anterior,  enlargement  of  fig.  6. 

Fig.  9.  77.  elongata-.  as  fig.  7 but  viewed  perpendicular  to  the  sulcus. 

Scale  bar  in  figs.  4,  7-9  = 200  ^ m;  figs.  1-3,  5,  6 = 400  ^m. 


PLATE  5 


GALE  and  SMITH,  Cretaceous  irregular  echinoids 


30 


PALAEONTOLOGY,  VOLUME  25 


(vii)  The  large  tubercles  of  the  anterior  interambulacra  found  on  either  side  of  the  frontal  sulcus 
are  arranged  in  two  rather  irregular  rows  with  interspersed  miliaries  in  species  of  Infulaster.  With 
the  development  of  the  rostrum  in  Hagenowia,  the  areas  of  interambulacra  which  face  towards 
the  anterior  are  reduced  and  there  is  only  a single  row  of  these  large  tubercles  on  either  side. 
Tubercles  are  rather  irregularly  arranged  in  H.  rostrata  but  in  later  species  they  abut  to  form  a 
well-defined  row  (PI.  5,  figs.  5-9;  PI.  6,  figs.  1,  2). 

(viii)  The  size  and  density  of  the  miliaries  which  cover  the  floor  and  walls  of  the  frontal  sulcus 
vary  amongst  species  (Table  1).  Miliaries  are  largest  in  I.  infulaster oides  and  H.  rostrata  and  become 
progressively  smaller  in  later  species.  Miliary  density  does  not  vary  along  the  length  of  the  sulcus 
in  I.  excentricus  and  I.  tuberculatus,  but  in  I.  infulasteroides  and  in  Hagenowia  the  miliaries  are 
denser  at  the  top  of  the  sulcus,  where  they  may  reach  densities  equivalent  to  that  found  in  the 
lateral  fasciole  (120  to  140  miliaries  per  mm* 1 2).  There  is  a corresponding  increase  in  the  density  of 
miliaries  lying  outside  the  sulcus  near  the  head  of  the  rostrum.  This  results  in  the  development  of 
a rather  diffuse  band  of  miliary  tubercles  that  fades  towards  the  dorsal  surface  of  the  rostrum. 


table  1.  Size  and  density  of  miliary  tubercles  in  ambulacrum  III 


Species 

Adapical  miliaries 

Median  miliaries 

Size 

Oun) 

Mean 

density 

(mm-2) 

Size 

(/am) 

Mean 

density 

(mm-2) 

Infulaster  excentricus 

50-60 

35 

50-60 

35 

I.  tuberculatus 

60-70 

50-55 

50 

50-55 

I.  infulasteroides 

50-75 

80 

80-100 

50-60 

Hagenowia  rostrata 

60-75 

75 

80-90 

40-50 

H.  anterior 

50 

100 

60-70 

60-70 

H.  blackmorei 

40-50 

130 

50-60 

70 

H.  elongata 

40-50 

140 

50-60 

60 

FUNCTIONAL  MORPHOLOGY 

Size  and  Shape 

1.  Size.  The  marked  reduction  in  the  size  of  Infulaster  in  the  early  Coniacian  could  have  had 
several  possible  advantages.  A reduction  in  size  increases  the  surface  area  to  volume  ratio.  This 
eases  the  animal’s  respiratory  demands  by  increasing  the  percentage  of  oxygen  that  can  be  obtained 
by  direct  diffusion.  By  reaching  sexual  maturity  at  an  earlier  stage,  smaller  species  can  have  a 
shorter  generation  time.  This  could  have  arisen  with  the  development  of  stable  and  highly  suitable 
conditions,  which  would  favour  those  species  able  to  reproduce  and  multiply  more  rapidly.  The 
reduction  in  size  may  also  have  been  brought  about  by  predatory  pressures.  Small  animals  with 
apical  elongation  would  be  much  less  obvious  from  the  surface,  when  buried,  than  larger, 
semi-infaunal  species. 

2.  The  anterior  ambulacrum.  The  anterior  ambulacrum  provides  the  main,  if  not  sole,  passageway 
for  transferring  sediment  adorally  in  Recent  spatangoids  (Smith  19806).  In  certain  species,  spines 
of  the  anterior  ambulacrum  are  specialized  for  mucous  string  feeding  (Chesher  1963;  Buchanan 
1966).  The  sunken  anterior  ambulacrum  in  Infulaster  and  Hagenowia  must  also  have  provided  an 
important  pathway  for  sediment  transportation.  In  Infulaster  the  anterior  sulcus  is  deep  but  in 
Hagenowia  it  shallows  progressively.  This  is  accompanied  by  increasing  development  of  the 
protective  arch  of  spines  covering  the  sulcus  (see  later,  p.  36).  The  reduction  in  the  depth  of  the 


GALE  AND  SMITH:  CRETACEOUS  ECHINOIDS 


31 


sulcus  is  linked  to  the  development  of  rostrum.  The  sulcus  decreases  in  depth  as  the  rostrum 
increases  in  slenderness  because  of  the  reduction  in  internal  volume.  In  Hagenoxvia  the  frontal 
channel  is  maintained  as  the  sulcus  shallows  by  the  development  of  a dense  grill  of  spines  arching 
across  ambulacrum  III. 

3.  The  rostrum.  The  development  and  elongation  of  the  rostrum  permitted  Hagenowia  to  live 
buried  within  the  sediment  while  maintaining  contact  with  surface  waters  (Nichols  1959;  Ernst  and 
Schulz  1971).  Via  the  rostrum,  oxygenated  water  could  be  drawn  into  the  burrow  and  the  surface, 
organic-rich  layer  of  sediment  passed  to  the  mouth.  Because  of  its  small  size,  Hagenowia  cannot 
be  said  to  have  burrowed  deeply.  The  oral  surface  of  H.  blackmorei  could  only  have  been  about 
3 cm  below  the  surface,  approximately  the  same  level  as  postulated  for  I.  excentricus.  However, 
the  rostrum  presumably  made  Hagenowia  less  obvious  from  the  surface  since  only  the  rostral  tip 
would  have  disturbed  the  surface  sediment  and  not  the  whole  dorsal  surface  (text-fig.  11).  Predation 
may  have  been  an  important  selection  pressure. 

The  rostrum,  which  is  oblique  in  H.  rostrata,  becomes  progressively  more  vertical  in  later  species. 
Simultaneously  it  also  develops  a cross-section  with  thicker  plates  and  buttressing.  Early  Hagenowia 
presumably  needed  an  oblique  rostrum  to  streamline  their  movement  through  the  sediment  and 
thus  minimize  the  stresses  applied  to  the  thin-walled  rostrum.  In  later  species,  as  the  rostrum 
became  more  robust,  it  could  be  held  in  a more  vertical  position. 

4.  The  anterior  movement  of  the  peristome.  In  Infulaster  the  peristome  lies  on  the  oral  surface 
near  the  anterior  margin  (text-fig.  7)  but  in  H.  rostrata  it  lies  marginally  at  the  base  of  the  anterior 
sulcus  and  in  H.  blackmorei  it  has  shifted  to  a frontal  position  (PI.  3,  fig.  lb).  This  change  is 
thought  to  reflect  an  increasing  dependence  on  sediment  coming  down  the  anterior  sulcus.  In 
Recent  spatangoids,  phyllode  tube-feet  are  used  to  collect  sediment  from  the  floor  of  the  burrow 
and  transfer  it  to  the  mouth  (Nichols  1959;  Chesher  1969;  Smith  1980o).  This  is  also  likely  to 
have  been  true  in  Infulaster.  Particles  coming  down  the  anterior  sulcus  would  have  landed  on  the 
floor  of  the  burrow  just  in  front  of  the  peristome  and  within  reach  of  the  tube-feet.  In  H.  rostrata 
the  mouth  lies  at  the  base  of  the  sulcus  so  that  tube-feet  would  have  been  able  to  collect  and 
transfer  particles  direct  from  the  sulcus.  In  H.  blackmorei  the  mouth  is  anterior  and  raised  well 
above  the  floor  of  the  burrow.  Peristomial  tube-feet  are  reduced  and,  because  of  their  position, 
could  only  have  been  involved  with  sediment  coming  down  the  sulcus. 

5.  The  plastron.  The  oral  surface  is  relatively  flat  in  Infulaster  but  becomes  more  or  less  keeled 
in  Hagenowia.  An  arched  or  keeled  plastron  is  found  in  Recent  spatangoids  which  burrow  in 
muddy  or  sandy  substrata.  Surface-dwelling  spatangoids  or  spatangoids  that  burrow  shallowly  in 
sands  or  gravels  have  a flat  plastron.  A keeled  or  arched  plastron  may  provide  a better  arrangement 
of  tubercles  and  spines  for  efficient  forward  thrust. 

6.  Sub-anal  protruberances.  Infulaster  has  two  clear  bulges  in  the  sub-anal  region  whereas  H. 
anterior  and  H.  blackmorei  have  only  a single  protruberance.  In  H.  rostrata  there  are  two  sub-anal 
protruberances  which  are  asymmetrical.  Usually  the  right-hand  bulge  is  larger  and  more  prominent 
than  the  other.  Devries  (1953)  showed  that  asymmetry  in  the  test  of  spatangoids  relates  to  the  way 
in  which  the  gut  is  coiled.  This  probably  accounts  for  the  asymmetry  of  the  sub-anal  region  in  H. 
rostrata. 

The  change  from  a double  to  a single  sub-anal  protruberance  is  probably  linked  with  the 
development  of  asymmetry.  Each  bulge  bore  a tuft  of  spines  (see  p.  34).  Whereas  the  test  of  I. 
excentricus  was  broad  enough  to  have  had  two  distinct  tufts  of  sub-anal  spines,  later  species  had 
progressively  narrower  bodies.  Presumably,  as  internal  volume  reduced,  the  gut  became  more 
tightly  packed  and  any  asymmetry  this  gave  to  the  test  was  enhanced.  In  Hagenowia  this  made 
one  of  the  tufts  of  spines  larger  and  more  posterior  in  position.  The  tuft  of  spines  on  the  smaller 
bulge  would  therefore  have  become  less  important  and  was  lost  in  later  species. 

Pore  morphology 

1.  Latero-dorsal  pores  and  their  tube-feet.  Pore  morphology  suggests  that  specialized  respiratory 
tube-feet  were  present  only  in  species  of  Infulaster  and  in  some  large  forms  of  H.  rostrata.  By 


32 


PALAEONTOLOGY,  VOLUME  25 


comparison  with  extant  species,  these  tube-feet  were  probably  rather  elongate  with  a central, 
partitioned  region  (see  Smith  1980a).  Respiratory  tube-feet  were  best  developed  in  the  posterior  I 
columns  of  ambulacra  II  and  IV.  Other  columns  either  had  less  elongate  respiratory  tube-feet  or  | 
had  thin-walled  cylindrical  tube-feet.  The  positioning  of  respiratory  tube-feet  presumably  reflects 
the  water-circulation  pattern  over  the  aboral  surface. 

The  unipores  and  partitioned  isopores  in  H.  rostrata  would  have  unspecialized  sensory  tube-feet. 

In  later  species  of  Hagenowia,  aboral  tube-feet  are  reduced  and  finally  lost.  This,  at  first  glance, 
appears  to  be  rather  a strange  adaptation  for  an  infaunal  echinoid.  However,  holasteroids,  in 
general,  show  no  modifications  for  protecting  aboral  tube-feet  during  burial.  Respiratory  tube-feet 
can  function  efficiently  in  infaunal  spatangoids  because  of  their  sunken  ambulacra  and  the  curved 
arch  of  protective  spines  above  them.  Respiratory  tube-feet  in  clypeasteroids  and  cassidulids  function 
efficiently  because  of  their  extreme  elongation  (Smith  1980a).  In  Infulaster,  respiratory  tube-feet 
were  neither  extremely  elongate,  nor  associated  with  sunken  ambulacra  nor  a protective  arch  of 
spines.  The  tube-feet  were  able  to  play  an  effective  part  in  gaseous  exchange  only  because  much 
of  the  dorsal  surface  remained  uncovered.  In  Hagenowia,  all  but  the  extreme  tip  of  the  rostrum 
was  covered,  thus  it  became  impossible  for  the  tube-feet  to  function  efficiently  and  they  were  quickly 
lost. 

The  change  in  efficiency  of  the  respiratory  tube-feet  may  be  linked  with  the  over-all  reduction 
in  body  size  that  occurred.  As  more  of  the  dorsal  surface  became  covered  by  sediment,  with  apical 
elongation,  respiratory  tube-feet  presumably  became  less  efficient.  To  compensate  for  this,  a 
reduction  in  body  size  took  place  so  that  a larger  proportion  of  the  total  oxygen  consumed  could 
come  from  direct  diffusion.  In  Hagenowia,  body  size  is  further  reduced  with  the  complete  loss  of 
latero-dorsal  tube  feet. 

2.  Pores  and  tube-feet  of  the  anterior  ambulacrum.  Small  isopores  are  found  along  the  length  of 

the  anterior  ambulacrum  in  species  of  Infulaster  and  in  H.  rostrata.  Their  similarity  with  the  ambital 
isopores  in  other  ambulacra  and  the  fact  that  they  are  uniform  in  shape  and  size  along  the  whole 
length  of  the  ambulacrum  (excluding  phyllode  pores)  indicates  that  the  associated  tube-feet  were 
sensory  in  function  and  terminated  in  a sensory  pad  rather  than  a disc.  The  pores  diverge  little  as 
they  pass  inwards,  and  the  accompanying  ampullae  were  probably  cylindrical  (Smith  1980a).  \ 

These  tube-feet  are  likely  to  have  actively  probed  the  sediment  in  front  of  them  as  well  as  sensing 
the  particles  passing  down  the  sulcus. 

In  later  species  of  Hagenowia,  isopores  are  replaced  by  unipores.  Those  in  the  sulcus  are  extremely 
small  and,  by  comparison  with  Recent  spatangoids,  are  likely  to  have  borne  small,  non-extensible 
epithelial  knobs  which  were  chemosensory  in  function.  Apical  unipores  differ  both  in  size  and  ; 
shape.  Similar  broad-rimmed  unipores  in  Recent  holasteroids  and  spatangoids  bear  relatively  large 
and  extensible,  sensory  tube-feet  (Smith  1980a).  Such  tube-feet  are  likely  to  have  been  important 
chemical  and  tactile  sense  organs  and  no  doubt  would  have  extended  out  of  the  funnel  to  probe 
the  surrounding  sediment. 

The  increase  in  pore  size  in  apical  unipores  of  H.  elongata  suggests  that  the  associated  tube-feet 
were  more  extensible  than  those  of  previous  species.  A larger  pore  means  that  a larger  volume  of 
coelomic  fluid  can  pass  in  and  out  of  the  tube-feet  rapidly. 

3.  Phyllode  pores  and  their  tube-feet.  The  isopores  surrounding  the  peristome  are  much  larger 
than  isopores  of  other  non-respiratory  tube-feet  in  both  Infulaster  and  Hagenowia.  Phyllode  tube-feet 
of  Recent  spatangoids  and  holasteroids  are  penicillate  and  work  by  mucous  adhesion  (Smith  1980a). 
Their  broad  disc  requires  the  support  of  a large  diameter  stem.  They  are  therefore  associated  with 
large,  oval  isopores  or  unipores.  The  arrangement  of  a few  large  pores  around  the  peristome  in 
Infulaster  and  Hagenowia  indicates  that  their  tube-feet  also  collected  sediment  by  means  of  a 
mucous  adhesive  disc.  However,  the  disc  could  not  have  been  particularly  broad,  judging  from 
the  size  of  the  isopores,  and  may  not  have  been  penicillate. 

There  are  only  two  small  unipores  around  the  peristome  of  H.  blackmorei.  These  are  likely  to 
have  been  associated  with  sensory  tube-feet. 


GALE  AND  SMITH:  CRETACEOUS  ECHINOIDS 


33 


Spine  morphology 

A specimen  of  /.  infulasteroides  (BMNH  E35766)  and  a specimen  of  H.  rostrata  (BMNH  E76848) 
both  retain  a large  number  of  different  spines.  Camera  lucida  drawings  of  these  spines  are  given 
in  text-fig.  9. 

A large  number  of  latero-dorsal  spines  lie  near  the  dorsal  ridge  in  the  specimen  of  I.  infulasteroides. 
These  are  about  1-75  mm  in  length  and  have  a pronounced  spatulate  tip.  The  spatulate  tip  is  set 
oblique  to  the  straight  shaft.  Amongst  these  spatulate  spines  is  a shaft,  about  5 mm  in  length,  with 
neither  tip  nor  base  preserved.  The  thinner,  distal  end  tapers  gently  and  is  unspatulate.  This  gently 
curved  spine  almost  certainly  attached  to  one  of  the  large  interambulacral  tubercles  at  the  apex 
of  the  anterior  sulcus. 

Ventro-lateral  spines,  found  below  the  marginal  fasciole  of  I.  infulasteroides,  are  spatulate  and 
2 to  2-5  mm  in  length.  The  spatulate  tip  is  also  set  oblique  to  the  shaft.  On  the  plastron  are  three 
spatulate  spines,  none  of  which  have  their  base  preserved.  Entire  spines  must  be  greater  than  1-75 
mm  in  length.  On  the  oral  surface  there  are  two  long  and  slender  spines,  one  of  which  is  broken 


1 . /.  excentricus  (BMNH  E40770):  a,  latero-ventral  spine;  b,  latero-dorsal  spine. 

2.  I.  infulasteroides  (BMNH  E35766):  a,  sub-anal  spine;  b,  latero-ventral  spine;  c,  latero-dorsal  spine; 
d,  apical  spine;  e,  plastron  spine;  /,  spine  from  the  protective  arch  across  the  frontal  sulcus; 
g,  anterior  excavatory  spine. 

3.  H.  anterior  (BMNH  E40158):  spine  from  the  protective  arch  across  the  frontal  sulcus. 

4.  H.  rostrata  (BMNH  E76848):  a,  latero-ventral  spine;  b,  anal  spine;  c,  plastron  spine. 


34 


PALAEONTOLOGY,  VOLUME  25 


just  below  the  milled  ring,  the  other  having  no  remnant  of  the  spine  base.  The  proximal  end  lies 
near  the  large  tubercles  of  the  sub-anal  protruberance.  Both  spines  are  gently  curved  and  are  about 
2-75  mm  in  length.  The  shaft  becomes  slightly  broader  distally,  but  does  not  appear  to  be  spatulate. 

In  the  anterior  sulcus  there  are  a number  of  non-spatulate  spines.  Most  of  those  lying  on  the 
floor  of  the  sulcus  are  small,  slender,  and  strongly  curved.  They  are  about  T5  mm  in  length  and 
do  not  flatten  distally.  These  are  derived  from  the  small  interambulacral  tubercules  lining  the  outer 
margin  of  the  sulcus,  judging  from  the  size  and  shape  of  the  spine  base.  Stouter  spines,  about  1-75 
mm  in  length,  are  found  within  the  sulcus  and  on  the  adjacent  interambulacra.  The  shaft  is 
slightly  bent  a little  above  the  milled  ring,  and  tapers  distally  (the  extreme  tip  is  unfortunately  not 
preserved).  These  spines  are  associated  with  the  large  interambulacral  tubercles  adjacent  to  the 
sulcus. 

Plastron  and  latero-ventral  spines  of  H.  rostrata  are  similar  to  those  of  I.  infulasteroides.  In  the 
specimen  of  H.  rostrata,  a number  of  stout,  straight  spines,  up  to  2-5  mm  in  length,  are  preserved 
within  the  periproct.  Their  tip  is  slightly  flattened  and  is  set  slightly  oblique  to  the  shaft.  These 
spines  are  probably  associated  with  the  large  interambulacral  tubercles  which  surround  the  periproct.  , 

The  spines  in  all  species  of  Infulaster  and  Hagenowia  are  likely  to  be  similar,  judging  from  the 
similarity  in  tubercle  structure  and  arrangement.  The  function  of  each  group  of  spines  is  best 
discussed  in  conjunction  with  the  structure  of  their  tubercles.  A reconstruction  of  the  spine  coverage 
in  I.  infulasteroides  and  Hagenowia  is  given  in  text-fig.  10.  Spine  posture  is  inferred  from  tubercle 
structure. 

Tubercle  structure  and  arrangement 

The  shape  and  arrangement  of  tubercles  can  give  information  on  spine  posture  and  movement 
(Smith  19806).  There  are  a number  of  distinct  areas  of  tubercles  in  Infulaster  and  Hagenowia,  each 
of  which  bore  spines  with  a specific  function.  Tubercle  arrangement  in  I.  excentricus,  H.  rostrata, 
and  H.  blackmorei  is  depicted  in  text-figs.  7 and  8.  The  following  groups  of  tubercles  can  be 
distinguished. 

1.  Plastron  tubercles.  Plastron  tubercles  in  both  Infulaster  and  Hagenowia  are  closely  spaced 
with  only  a few  interspersed  miliary  tubercles.  Largest  tubercles  lie  laterally  with  smallest  tubercles 
lying  along  the  central  ridge.  Their  areole  is  slightly  enlarged  on  the  meso-posterior  side  (text-figs. 

7,  8)  and  platform  crenulation  is  radially  symmetrical.  As  in  Recent  spatangoids,  the  power  stroke  | 
would  have  been  towards  the  mid  posterior,  and  plastron  spines,  with  their  spatulate  tips,  must 
have  provided  the  main  thrust  for  locomotion.  However,  plastron  tubercles  of  Recent  spatangoids 
have  asymmetric  crenulation  and  pronounced  areole  enlargement  (Smith  19806).  The  movement 
of  plastron  spines  was  probably  more  restricted  and  less  effective  in  Infulaster  and  Hagenowia 
compared  with  Recent  spatangoids.  Radially  symmetrical  crenulation  indicates  that  plastron  spines 
were  held  more  or  less  perpendicular  to  the  test  (Smith  19806). 

2.  Latero-ventral  tubercles.  Tubercles  found  below  the  marginal  fasciole  in  interambulacra  1-4 
have  their  areole  enlarged  adambitally  and  towards  the  fasciole.  Platform  crenulation  is  either 
radially  symmetrical  or  is  slightly  larger  on  the  adoral  side  (opposite  the  direction  of  areole 
enlargement).  These  tubercles  are  denser  on  the  anterior  half  but  become  more  uniformly  distributed 
in  later  species.  They  decrease  in  size  and  become  more  densely  packed  towards  the  fasciole. 

This  arrangement  differs  from  that  found  in  Recent  spatangoids,  where  the  areole  is  enlarged 
in  a latero-posterior  direction  and  crenulation  is  asymmetrical.  (Smith  19806).  In  spatangoids  the 
associated  spines  lie  oblique  to  the  test  and  excavate  sediment  from  beneath  the  animal  with  an 
oar-like  action.  Tubercle  structure,  in  Infulaster  and  Hagenowia,  indicates  that,  whereas  oral  spines 
were  more  or  less  perpendicular,  adambital  spines  were  inclined  slightly  downwards  away  from 
the  fasciole.  Areole  enlargement  suggests  that  spines  pushed  sediment  outwards  and  upwards  and 
were  principally  used  in  excavating  and  burrowing. 

3.  Sub-anal  tubercles.  Tubercles  are  radially  arranged  on  each  sub-anal  bulge  with  large,  radially 
symmetrical  tubercles  centrally,  and  smaller  tubercles  marginally  (text-figs.  7 and  8).  Marginal 
tubercles  often  have  a slight  areole  enlargement  on  the  side  away  from  the  centre.  A similar  tubercle 


GALE  AND  SMITH:  CRETACEOUS  ECHINOIDS 


35 


EXCAVATORY 
SPINES 


SEDIMENT-MOVING 
SPINES 


SUB-ANAL  TUFT 


LOCOMOTORY  SPINES 


text-fig.  10.  Spine  posture  and  arrangement  in  Infulaster  and  Hagenowia.  A,  Infulaster  infulasteroides, 
lateral  view,  b,  Hagenowia  blackmorei,  cut-away  section  through  the  rostrum  showing  the  anterior  sulcus. 
The  relative  size  of  spines  is  based  on  specimen  BMNH  E35766,  and  spine  posture  and  function  is 
interpreted  from  tubercle  structure. 


arrangement  in  Recent  spatangoids  is  associated  with  tufts  of  spines  (Smith  19806).  Infulaster  and 
H.  rostrata  must  have  had  two  sub-anal  tufts  of  spines  whereas  later  species  of  Hagenowia  had 
only  a single  tuft.  The  upper  spines  of  each  tuft  abut  against  the  marginal  fasciole.  Mucus  from 
the  fasciole  could  easily  have  been  transferred  to  the  sub-anal  spines.  The  sub-anal  tuft  of  spines 
is  used  in  constructing  a tunnel,  in  spatangoids,  with  the  aid  of  specialized  tube-feet  (Nichols  1959; 
Chesher  1968).  Infulaster  and  Hagenowia  lacked  sub-anal,  funnel-building  tube-feet  and  were 
unlikely  to  have  built  an  extensive  tunnel.  However,  the  spines  alone  may  have  been  able  to 
maintain  a short  sub-anal  tunnel  with  the  help  of  mucus  from  the  fasciole.  In  spatangoids  the 
sub-anal  tunnel  is  used  to  remove  water  from  the  burrow  (Chesher  1968). 

4.  Anal  tubercles.  Large  tubercles  are  found  around  the  dorsal  and  lateral  edges  of  the  periproct. 
They  decrease  in  size  away  from  the  periproct  and  are  most  numerous  dorsally  and  on  either  side 
immediately  below  the  periproct.  The  crenulate  platform  is  radially  symmetrical.  There  may  be 
slight  areole  enlargement  on  the  anterior  side.  Beneath  the  periproct  is  a broad  band  of  miliaries, 
extending  to  the  fasciole.  A similar  arrangement  of  tubercles  is  found  in  Recent  spatangoids,  where 
a semicircle  of  longer  spines  surrounds  the  periproct.  These  maintain  a space  within  the  burrow 
for  defaecation  and  prevent  fouling  of  the  aboral  surface. 


36 


PALAEONTOLOGY,  VOLUME  25 


5.  Latero-dorsal  tubercles.  Tubercles  above  the  marginal  fasciole  have  only  a weak  bilateral 
symmetry  in  I.  excentricus  and  I.  tuberculatus.  This  becomes  much  more  pronounced  in  Hagenowia 
due  to  a marked  enlargement  of  the  areole  on  the  anterior  or  adapical  side  (see  text-fig.  8).  Numerous 
miliaries  occur  between  the  tubercles.  These  tubercles  support  short,  spatulate  spines  in  /. 
infulasteroides.  Similar  spatulate  spines,  attaching  to  bilaterally  symmetrical  tubercles,  are  found 
in  Recent  spatangoids,  where  they  are  used  for  compacting  the  burrow  walls,  transporting  sediment 
posteriorly,  and  moving  the  dorsal  mucous  coat  (Smith  19806).  I.  excentricus  has  a low  aboral 
tubercle  density  but  later  species  of  Infulaster  and  Hagenowia  have  progressively  denser  tubercula- 
tion.  The  increased  protection  provided  by  a denser  spine  coverage  is  an  adaptation  for  living 
infaunally.  Tubercles  are  especially  dense  along  the  rostrum,  presumably  for  compacting  and 
maintaining  an  open  passageway  down  into  the  burrow. 

In  H.  blackmorei,  tubercles  just  above  the  marginal  fasciole  have  an  adoral  areole  enlargement 
indicating  that  the  spatulate  tips  of  spines  were  tilted  in  opposite  directions  on  either  side  of  the 
fasciole,  a situation  common  in  Recent  spatangoids  for  distributing  mucus. 

6.  Tubercles  associated  with  the  anterior  sulcus.  There  are  two  sets  of  interambulacral  tubercles 
lying  adjacent  to  the  anterior  sulcus.  Firstly,  there  are  small,  radially  symmetrical  tubercles  which 
form  an  inward-facing  row  along  the  lip  of  the  sulcus  (PI.  4,  figs.  5-9;  PI.  5,  figs.  1-4).  They 
are  rather  irregularly  arranged  in  early  species  but  become  organized  into  a well-defined  line  in 
Hagenowia.  The  associated  curved  and  non-spatulate  spines  would  have  formed  an  arched  roofing 
to  the  anterior  sulcus  (text-fig.  10b).  The  increasing  density  and  organization  of  these  spines  was 
partially  due  to  the  decreasing  depth  of  the  anterior  sulcus  and  partially  due  to  the  increasing 
importance  of  the  sulcus  as  a pathway  for  transporting  sediment  to  the  mouth.  In  Hagenowia  the 
anterior  shift  of  the  peristome  shows  that  more  reliance  was  placed  on  sediment  coming  down  the 
sulcus.  The  arch  of  spines  would  have  maintained  a clear  frontal  pathway  and  would  have  prevented 
sediment  from  the  frontal  wall  of  the  burrow  clogging  or  contaminating  the  stream  of  sediment 
passing  down  the  anterior  sulcus.  They  ensured  that  sediment  could  only  enter  the  sulcus  at  the  apex. 


EXPLANATION  OF  PLATE  6 

Fig.  1.  Hagenowia  elongata:  large  interambulacral  tubercles  adjacent  to  the  rostral  sulcus  that  support 
excavating  spines.  Crenulation  is  radially  symmetrical  but  the  areole  is  enlarged  laterally  away  from 
ambulacrum  III.  Smaller  tubercles  for  protective  arch  spines  can  be  seen  towards  the  bottom.  MM  12820. 

Fig.  2.  H.  anterior,  interambulacral  tubercles,  as  above  but  with  intervening  miliaries.  BMNH  E76844. 

Fig.  3.  H.  elongata\  row  of  small  interambulacral  tubercles  immediately  adjacent  to  the  frontal  sulcus  (bottom  of 
micrograph)  for  spines  forming  a protective  arch  to  the  sulcus.  MM  12820. 

Fig.  4.  H.  elongata:  enlargement  of  fig.  3 showing  the  radially  symmetrical  areole  and  the  lack  of  crenulation. 

Fig.  5.  Infulaster  tuberculatus : large  apical  interambulacral  tubercles,  ambulacrum  III  to  the  right.  These  have 
radially  symmetrical  areoles  and  crenulation  and  are  not  tilted.  Their  spines  must  have  been  held  pointing 
away  from  ambulacrum  III.  BMNH  E76852. 

Fig.  6.  H.  anterior:  miliary  tubercles  near  the  apex  of  the  rostral  sulcus,  apex  to  the  sulcus.  These  are  tilted 
adapically  and  laterally,  away  from  the  mid-line,  suggesting  that  their  spines  had  a meso-adoral  power  stroke. 
BMNH  E76844. 

Fig.  7.  H.  blackmorei:  adapical  part  of  ambulacrum  III  showing  dense,  irregular  arrangement  of  varyingly  sized 
tubercles  at  the  top  of  the  sulcus.  BMNH  E76846. 

Fig.  8.  H.  anterior:  large  ambulacral  tubercle  adjacent  to  an  apical  ambulacrum  III  unipore.  Crenulation  is 
slightly  stronger  laterally  (towards  the  pore)  and  the  protective  spine  must  have  been  gently  curved.  BMNH 
E76844. 

Fig.  9.  H.  blackmorei:  large  apical  interambulacral  tubercle  with  a radially  symmetrical  areole  and  slightly 
stronger  crenulation  towards  ambulacrum  III  (top  of  micrograph).  BMNH  E76846. 

Fig.  10.  H.  elongata:  large  apical  interambulacral  tubercles.  The  areole  is  slightly  border  to  the  posterior 
(bottom  of  micrograph),  crenulation  is  radially  symmetrical.  MM  12820. 

Scale  bar  in  figs.  4,  8 = 40  ^m;  figs.  1-3,  6,  9,  10  = 100  /urn;  figs.  5,  7 = 200  /u m . 


PLATE  6 


GALE  and  SMITH,  Cretaceous  irregular  echinoids 


38 


PALAEONTOLOGY,  VOLUME  25 


The  other  set  of  interambulacral  tubercles  are  large  and  outward-facing  (PI.  5,  figs.  5-9;  PI.  6, 
figs.  1-3).  They  form  a single,  well-defined  row  in  Hagenowia.  Platform  crenulation  is  either  radially 
symmetrical  or  is  enlarged  slightly  on  the  side  facing  the  sulcus.  Their  areole  is  enlarged  in  the 
opposite  direction  (text-fig.  8;  PI.  6,  fig.  1)  showing  that  the  power  stroke  of  the  spine  was  directed 
away  from  the  sulcus.  These  stout,  pointed  spines  were  probably  used  to  loosen  and  excavate 
sediment  of  the  anterior  wall  of  the  burrow.  Recent  spatangoids,  such  as  Echinocardium  cordatum 
(Pennant),  have  both  an  anterior  arch  of  protective  spines  and  adjacent  excavating  spines,  and 
their  tubercles  and  spines  are  rather  similar. 

The  apical  interambulacral  tubercles  and  spines  of  Infulaster  and  Hagenowia  are  the  largest  they 
possess.  There  are  between  16  and  20  such  tubercles  on  the  adapical  plates  of  the  anterior 
interambulacra.  Their  areoles  are  either  radially  symmetrical  or  are  enlarged  slightly  to  the  posterior 
(PI.  5,  figs.  1-3;  PI.  6,  figs.  5,  9,  10).  Platform  crenulation  is  more  or  less  radially  symmetrical 
though  crenulation  may  be  slightly  enlarged  on  the  anterior  side  (PI.  6,  fig.  9).  Tubercles  are  so 
positioned  that,  were  the  spines  to  converge  adapically  to  form  a tuft,  they  would  have  to  lie  at 
a considerable  angle  to  the  tubercles  and  the  plate  surface.  The  structure  of  the  tubercles,  however, 
shows  that  spines  were  not  set  obliquely,  forming  an  apical  tuft,  but  probably  formed  a fan-shaped 
array.  This  fan  of  spines  would  have  formed  a collar  to  the  anterior  sulcus  (text-fig.  10a).  This 
arrangement  is  common  to  all  species  of  Infulaster  and  Hagenowia  and  is  also  found  in  certain 
species  of  Cardiaster.  As  the  apex  is  the  most  vulnerable  part  of  the  test  in  infaunal  or  semi-infaunal 
species,  these  apical  spines  may  have  acted  as  a deterrent  to  predators.  The  fan-shaped  arrangement 
around  the  apical  opening  to  the  frontal  sulcus  suggests  that  apical  spines  were  also  involved  in 
collecting  sediment,  either  acting  as  a funnel  or  helping  to  cascade  surface  sediment  into  the  sulcus. 

Within  ambulacrum  III,  small  tubercles  are  interspersed  with  miliaries  near  the  apex  (PI.  6, 
fig.  7),  but  in  the  sulcus  ambulacral  plates  are  covered  almost  entirely  by  uniformly  sized  miliaries 
(PI.  6,  fig.  6).  A similar  arrangement  is  found  in  Micr aster  but  has  not  yet  been  reported  in  any 
Recent  species  (Smith,  19806).  In  many  Recent  spatangoids,  miliary  tubercles  of  ambulacrum  III 
bear  short,  straight  spines  with  fleshy  tips,  usually  well  endowed  with  cilia  and  mucous  glands.  A 
similar  type  of  spine  was  probably  present  in  the  sulcus  of  Infulaster  and  Hagenowia. 

Several  lines  of  evidence  suggest  that  the  anterior  sulcus  became  increasingly  important  as  a 
pathway  for  transferring  sediment  to  the  mouth.  Evolutionary  changes  in  the  size  and  density  of 
miliaries  in  the  sulcus  may  therefore  be  related  to  an  improved  feeding  mechanism.  The  progressive 
differentiation  of  a distinct  band  of  dense  miliaries  at  the  apex  of  the  sulcus  may  mark  the  increasing 
importance  of  ciliary  currents  or  mucus  binding  for  particle  transportation.  By  concentrating 
miliary  spines  at  the  apex,  where  sediment  entered  the  sulcus,  it  is  possible  that  copious  amounts 
of  mucus  could  have  been  produced  to  enmesh  particles  loosely.  The  relatively  few  spines  within 
the  sulcus  could  then  have  transported  a stream  of  mucus-bound  sediment  with  relative  ease. 
Tubercle  and  spine  arrangement  is  highly  distinctive  in  Recent,  mucous  string  feeding  spatangoids 
(Chesher  1963;  Buchanan  1966;  Smith  19806)  and  very  different  from  that  in  Hagenowia.  Hagenowia , 
then,  did  not  feed  by  means  of  a compact  mucous  string  but  may  have  had  a much  looser  flow 
of  weakly  bound  particles. 

7.  Fascioles.  The  marginal  fasciole  shows  little  change  except  that  it  becomes  more  sharply 
defined  across  the  anterior  interambulacra  in  later  species  of  Hagenowia.  The  increasing  obliquity 
of  the  fasciole  is  a direct  consequence  of  the  changing  position  of  the  peristome.  The  fasciole  does 
not  cross  the  anterior  ambulacrum,  as  stated  by  Nichols  (1959),  but  becomes  more  diffuse  and 
peters  out  towards  the  sulcus. 

Fascioles  act  as  pumps,  drawing  oxygenated  water  through  the  burrow,  and  as  a source  of 
mucus.  A dorsal  mucous  coat  prevents  finer  particles  falling  between  spines  and  clogging  the 
burrow.  Both  Infulaster  and  Hagenowia  are  likely  to  have  lived  at  least  partially  buried  and  the 
fasciole  probably  drew  water  down  from  the  apex  and  over  the  central  and  posterior  parts  of 
the  test. 


GALE  AND  SMITH:  CRETACEOUS  ECHINOIDS 


39 


PRESERVATION 

Increased  differentiation  and  strengthening  of  the  rostrum  in  successive  species  of  Hagenowia  was 
accompanied  by  relative  thinning  of  the  body  wall.  This  is  reflected  in  the  increased  rarity  of 
preservation  of  Hagenowia  bodies  in  successively  later  species  and  the  greater  frequency  of  isolated 
rostra. 

Bodies  of  Hagenowia  are  rarely  bored  or  encrusted  by  the  organisms  which  colonized  much  of 
the  skeletal  debris  littering  the  chalk  sea  floor,  and  it  is  likely  that  once  exhumed,  the  fragile  tests 
fragmented  rapidly  leaving  only  the  more  durable  rostra.  The  rostra  often  contain  sponge  crypts, 
and  characteristic  oval  borings  (500-1000  /xm  in  length)  with  irregularly  bevelled  rims.  Encrusters 
are  extremely  rare  on  H.  rostrata,  probably  because  they  were  too  small  to  provide  suitable  substrata. 

One  preservational  style  occurs  consistently  in  Hagenowia,  in  which  the  rostrum  is  lost,  having 
broken  away  along  plate  sutures  near  its  base  (e.g.  PI.  3,  fig.  2).  The  resulting  irregular  stump  has, 
in  some  specimens  been  attacked  by  boring  organisms  (e.g.  PI.  3,  figs.  6b,  la).  Such  specimens 
frequently  retain  scattered  spines  on  the  sides  and  base  of  the  body.  Presumably  these  individuals 
remained  in  life  orientation  within  the  sediment  after  death,  with  the  rostrum  protruding  above 
the  sediment-water  interface,  which  was  stable  for  long  enough  to  allow  colonization  by  borers. 

Two  specimens  of  Hagenowia  have  repaired  injury  to  the  rostrum.  In  one  specimen  (BMNH 
E76849)  the  obliquely  broken  tip  of  the  rostrum  was  sealed  across,  and  new  tubercles  and  genital 
pores  developed.  Such  damage  was  probably  the  result  of  predation. 

SELECTION  PRESSURES  AND  MODE  OF  LIFE 
The  majority  of  structural  changes  in  the  Infulaster  I Hagenowia  lineage  were  caused,  either  directly 
or  indirectly,  by  elongation  of  the  apical  region  of  the  test,  and  to  a lesser  extent  by  the  reduction 
in  size.  Apical  elongation  necessitated  disruption  of  the  apical  system,  plate  elongation,  loss  of 
plate  columns  from  the  rostrum,  and  an  increased  slenderness  of  the  rostrum.  This  increased 
slenderness  in  turn  reduced  the  internal  volume  of  the  rostrum,  which  caused  a decrease  in  depth 
of  the  sulcus,  loss  of  two  of  the  four  gonads,  and  the  change  from  extensible  tube-feet  with  large 
ampullae,  to  small  tube-feet  with  reduced  ampullae.  The  slender  rostrum  required  buttressing  at 
the  anterior  and  posterior  margins,  and  an  increasingly  better  organization  of  tubercles.  The  increase 
in  obliquity  of  the  fasciole  was  a consequence  of  the  anterior  and  adapical  movement  of  the 
peristome.  The  size  reduction  resulted  in  the  change  from  a double  to  a single  subanal  bulge  and 
in  the  interruption  of  the  plastron. 

Various  features  displayed  by  Infulaster  and  Hagenowia  are  interpreted  as  adaptations  to  an 
infaunal  mode  of  life.  The  latero-ventral  spines  were  specialized  for  excavating  sediment  from 
beneath  the  animal.  In  order  for  the  fasciole  and  subanal  tuft  of  spines  to  have  functioned,  these 
echinoids  must  have  lived  at  least  partially  buried.  Latero-dorsal  spines  modified  for  compacting 
sediment  of  the  burrow  walls  are  developed  to  immediately  below  the  apex,  as  are  the  excavating 
spines  and  the  protective  arch  of  spines  across  the  sulcus  (text-fig.  10).  This  clearly  implies  that 
both  Infulaster  and  Hagenowia  lived  within  the  sediment,  with  their  apices  at  or  just  below  the 
surface. 

It  is  highly  improbable  that  either  genus  could  have  burrowed  more  deeply,  as  funnel-building 
tube-feet  are  lacking,  and  the  apical  spines  were  held  in  a broad  fan  and  not  a tuft.  Living  within 
fine-grained  sediments,  they  had  to  maintain  direct  contact  with  surface  waters.  This  could  only 
have  been  achieved  by  keeping  the  apex  of  the  test  at  the  surface. 

I.  excentricus  has  an  almost  flat  dorsal  surface,  most  of  which  must  have  remained  exposed  on 
the  sea  floor.  Later  species  evolved  to  become  less  conspicuous,  by  apical  elongation.  In  this  way 
the  bulk  of  the  animal  could  be  removed  from  the  surface,  with  only  a small  apical  part  of  the 
test  visible.  This  apical  elongation  brought  about  the  structural  changes  enumerated  above.  The 
loss  of  aboral  respiratory  tube-feet  was  probably  an  important  factor  in  influencing  the  size  reduction 
which  occurred. 


40 


PALAEONTOLOGY,  VOLUME  25 


Other  morphological  changes  accompany  a change  in  feeding  strategy.  Infulaster  used  its  phyllode 
tube-feet  to  pick  up  sediment  around  the  mouth.  Sediment  coming  down  the  frontal  sulcus  was 
not  collected  directly  and  probably  formed  only  a small  part  of  the  animal’s  diet.  Hagenowia  came 
to  rely  much  more  heavily  on  sediment  from  the  anterior  sulcus  and  may  have  developed  some 
sort  of  mucus  binding  to  cope  with  this  increased  volume  of  sediment.  Adaptations  linked  with 
this  change  in  feeding  strategy  include  the  anterior  movement  of  the  peristome,  loss  of  large 
phyllode  tube-feet,  and  increasingly  denser  arch  of  spines  across  the  sulcus  and  their  extension  to 
cover  the  peristome,  the  more  extensible  sensory  tube-feet  at  the  apex,  and  the  increasing 
differentiation  of  miliaries  within  the  sulcus. 

All  the  morphological  trends  and  specializations  discussed  are  directly  or  indirectly  concerned 
with  adaptations  for  feeding  or  burial.  The  predominant  pressures  acting  on  the  Infulaster  - 
Hagenowia  lineage  were  for  increased  efficiency  in  feeding  and  for  increased  protection  by  becoming 
less  obvious  at  the  surface.  I.  excentricus  lived  semi-infaunally  with  much  of  the  dorsal  surface 
exposed,  and  progressed  leaving  a clear  furrow  in  its  wake  (text-fig.  1 1).  In  later  species,  as  the 
dorsal  surface  became  more  oblique  with  apical  elongation,  less  and  less  of  the  test  remained 
exposed  at  the  surface.  In  H.  blackmorei  only  the  very  small  apical  head  remained  near  the  surface, 
the  remainder  of  the  body  being  fully  covered  (text-fig.  1 1).  Only  a small  circular  hole  at  the  surface 
would  have  marked  the  position  of  this  species. 

It  is  not  entirely  correct  to  speak  of  an  increasing  depth  of  burial  in  the  Infulaster-Hagenowia 
lineage,  since  the  apex  stays  at  the  same  level  and  the  oral  surface  of  I.  excentricus  lay  at  considerably 
greater  depth  than  in  many  later  species,  because  of  the  reduction  in  size.  However,  with  apical 
elongation  and  the  more  vertical  orientation  of  the  rostrum,  the  oral  surface  was  removed  further 
from  the  sea  floor.  The  most  likely  cause  of  this  change  is  increasing  predation,  with  selection 


text-fig.  1 1 . Mode  of  life  of  Infulaster  and  Hagenowia.  Cut-away  block  sections  showing  the  depth  of  burial  of 
Infulaster  excentricus  (left-hand  side)  and  Hagenowia  blackmorei  (right-hand  side).  Further  back  a second 
individual  is  illustrated  to  highlight  the  change  in  surface  appearance  with  the  development  of  apical  elongation. 


GALE  AND  SMITH:  CRETACEOUS  ECHINOIDS 


41 


against  those  more  obvious  at  the  surface.  Predation  was  a threat  to  Hagenowia,  since  regenerated 
rostra  are  known. 

It  is  not  clear  whether  the  change  in  feeding  habit  resulted  from  a change  in  the  nature  of  the 
sediment  (e.g.  the  development  of  a surface  layer  richer  in  organic  material  or,  conversely,  an 
over-all  reduction  in  organic  content)  or  reflected  a progressive  adaptation  to  a constant  sediment 
type.  Infulaster  and  Hagenowia  are  both  found  in  bands  in  apparently  uniform  pelagic  chalk 
sequences.  However,  original  subtle  differences  in  the  nature  of  the  sediment  may  well  have  been 
obscured  by  bioturbation  and  diagenesis. 

Hagenowia  provides  some  indirect  evidence  as  to  the  nature  of  the  chalk  sea  floor.  The  presence 
of  cutting  and  compacting  spines  on  the  rostrum  implies  that  the  sediment  in  which  it  lived  was 
sufficiently  cohesive  to  both  require  and  allow  such  treatment.  This  is  in  accord  with  the  suggestion 
made  from  other  lines  of  evidence  by  Surlyk  and  Birkelund  (1977)  that  the  bottom  was  relatively 
stable,  with  only  a superficial  layer  (a  few  millimetres  thick)  of  easily  resuspendible  material. 

Acknowledgements.  R.  P.  S.  Jefferies  (Natural  History  Museum,  London),  C.  J.  Wood  (Geological  Museum, 
London),  S.  Floris  (Geological  Museum,  Copenhagen),  and  C.  W.  Wright  kindly  made  specimens  in  their 
care  available  for  study.  A.  S.  G.  would  like  to  thank  J.  M.  Hancock,  K.  Holdaway,  C.  W.  Wright,  and  U. 
Asgaard  for  valuable  discussion  and  criticism.  P.  Howard  (King’s  College  London)  helped  with  the 
photography.  Part  of  this  work  was  carried  out  with  the  support  of  an  S.R.C.  grant. 


REFERENCES 

brydone,  r.  m.  1914.  The  Zone  of  Off aster  pilula  in  the  South  English  Chalk.  Part  1 -4.  Geol.  Mag.  (6)  1, 395-469, 
405-11,  449-57,  509-73. 

buchanan,  i.  b.  1966.  The  biology  of  Echinocardium  cor  datum  (Echinodermata:  Spatangoida)  from  different 
habitats.  J.  mar.  biol.  Ass.  U.K.,  46,  97-114. 

chesher,  R.  H.  1963.  The  morphology  and  function  of  the  frontal  ambulacrum  of  Moira  atropos  (Spatangoida). 
Bull  mar.  Sci.  Gulf  Carrib.  13,  549-573. 

— 1968.  The  systematics  of  sympatric  species  in  West  Indian  spatangoids:  a revision  of  the  genera  Brissopsis, 
Plethotaenia,  Paleopneustes  and  Saviniaster.  Stud.  trop.  Oceanogr.  Miami , 7,  viii+168  pp. 

— 1969.  Contributions  to  the  biology  of  Meoma  ventricosa  (Echinoidea:  Spatangoida).  Bull.  mar.  Sci.  19, 
72-110. 

desor,  e.  1858.  Synopsis  des  Echinides  Fossiles,  vi,  321-490.  Paris. 

devries,  a.  1953.  Contribution  a l’etude  de  la  symetrie  bilaterale  chez  les  Spatangues.  Bull.  Serv.  Carte  geol. 
Algerie  (1)  Paleont,  15,  37-69. 

duncan,  p.  M.  1889.  A revision  of  the  genera  and  great  groups  of  the  Echinoidea.  J.  Linn.  Soc.  23,  1-311. 
ernst,  G.  and  schulz,  M.-G.  1971.  Die  Entwicklungsgeschichte  der  hochspezialisierten  Echiniden-Reihe 
Infulaster— Hagenowia  in  der  Borealen  Oberkreide.  Palaont.  Z.  45,  120-143. 

— 1974.  Stratigraphie  und  Fauna  des  Coniac  und  Santon  im  Schreibkreide-Richtprofil  von  Lagerdorf 
(Holstein).  Mitt.  geol. — palaont.  Inst.  Univ.  Hamburg,  43,  5-60. 
ernst,  G.  and  seibertz,  E.  1977.  Concepts  and  methods  of  echinoid  biostratigraphy.  Pp.  541-63.  In  Kauffman,  E. 
G.,  and  Hazel,  J.  E.  (eds.).  Concepts  and  methods  in  biostratigraphy.  Dowden,  Hutchinson  & Ross,  Inc., 
Stroudsburg. 

forbes,  e.  1852.  Figures  and  descriptions  illustrative  of  British  organic  remains.  Mem.  geol.  Surv.  U.K.,  Dec.  iv, 
1-4. 

gaster,  c.  t.  a.  1924.  The  Chalk  of  the  Worthing  District,  Sussex.  Proc.  Geol.  Tss.  35,  89-110. 
mortensen,  t.  1950.  A monograph  of  the  Echinoidea  V.  I.  Spatangoida  1.  Copenhagen. 
moskveena,  m.  m.,  1959.  Atlas  of  the  Upper  Cretaceous  fauna  of  the  northern  Caucasus  and  Crimea.  Trudy , 
Glav.  Uprav.  Gaz.  Promyshal.  Sov.  Min.  USSR.  (VNIGAZ),  310  pp.  [In  Russian.] 
nichols,  d.  1959.  Changes  in  the  Chalk  heart-urchin  Micraster  interpreted  in  relation  to  living  forms.  Phil. 
Trans.  R.  Soc.,  Lond.  B,  242,  347-437. 

nielsen,  k.  b.  1942.  Martinosigra  elongata  n.g.  et  n.  sp.,  A new  echinoid  from  the  White  Chalk  of  Denmark. 
Meddr  Dansk  Geol.  Foren.  10,  2,  pp.  159-166. 

nietsch,  H.  1921.  Die  irregularen  Echiniden  der  pommerschen  Kreide.  Abh.  geol. — palaont.  Inst.  Greifswald,  2, 
1-47. 


42 


PALAEONTOLOGY,  VOLUME  25 


peake,  N.  b.  and  Hancock,  J.  M.  1961.  The  Upper  Cretaceous  of  Norfolk.  Pp.  293-339.  In  Larwood,  G.  P.  and 
Funnell,  B.  M.  (eds.).  The  geology  of  Norfolk.  Trans.  Norfolk  Norwich  Nat.  Soc.  19,  (6),  pp.  269-375. 
rawson,  p.  f.  et  al.  1978.  A correlation  of  Cretaceous  rocks  in  the  British  Isles.  Geol.  Soc.  Lond.,  Special  Report 
No.  9,  70  pp. 

reid,  r.  e.  h.  1976.  Late  Cretaceous  climatic  trends,  faunas  and  hydrography  in  Britain  and  Ireland.  Geol.  Mag. 
113,  115-128. 

rowe,  a.  w.  1 899.  An  analysis  of  the  genus  Micraster,  as  determined  by  rigid  zonal  collecting  from  the  Zone  of 
Rhynchonella  Cuvieri  to  that  of  Micraster  coranguinum.  Q.  Jl  geol.  Soc.  Lond.  55,  494-547,  pis.  35-39. 

— 1904.  The  White  Chalk  of  the  English  coast:  Part  IV.  Yorkshire.  Proc.  Geol.  Ass.  18,  193-296. 
schmid,  f.  1972.  Hagenowia  elongata  (Nielsen),  ein  hochspezialisierter  Echinide  aus  dem  hoheren 
Untermaastricht  NW-Deutschlands.  Geol.  Jber.  4,  177-195. 
smith,  a.  b.  1980a.  The  structure,  function  and  evolution  of  ambulacral  tube  feet  and  pores  of  irregular 
echinoids.  Palaeontology , 23  (1),  39-83. 

— 19806.  The  structure  and  arrangement  of  echinoid  tubercles.  Phil.  Trans.  R.  Soc.,  Lond.  B,  289,  1-54. 
surlyk,  f.  and  birkelund,  t.  1977.  An  integrated  stratigraphical  study  of  fossil  assemblages  from  the 
Maastrichtian  White  Chalk  of  northwestern  Europe.  Pp.  257-281.  In  Kauffman,  E.  G.,  and  Hazel,  J.  E.  (eds.). 
Concepts  and  methods  of  biostratigraphy.  Dowden,  Hutchinson  & Ross  Inc.,  Stroudsburg. 
wright,  t.  w.  1881.  Monograph  of  the  British  Fossil  Echinodermata  from  the  Cretaceous  Formations.  Vol.  1. 

The  Echinoidea.  Part  9,  pp.  301-324,  Palaeontogr.  Soc.  ( Monogr .) 
wright,  c.  w.  and  wright,  e.  v.  1942.  The  Chalk  of  the  Yorkshire  Wolds.  Proc.  Geol.  Ass.  53,  112-127. 

1949.  The  Cretaceous  echinoid  genera  Infulaster  Desor  and  Hagenowia  Duncan.  Ann.  Mag.  nat.  Hist. 

(12)  2,  454-474. 


ANDREW  S.  GALE 

Department  of  Geology 
King’s  College,  London 
Strand 

London  WC2R  2LS 

ANDREW  B.  SMITH 

Department  of  Geology 
University  of  Liverpool 
P.O.  Box  147 
Liverpool  L69  38X 


Typescript  received  28  January  1980 


A REVISION  OF  LATE  ORDOVICIAN  BIVALVES 
FROM  POMEROY,  CO.  TYRONE,  IRELAND 

by  S.  P.  TUNNICLIFF 


Abstract.  A history  of  research  on  Ordovician  bivalves  from  Pomeroy  is  given,  and  their  stratigraphic 
distribution  in  the  Caradoc-Ashgill  rocks  is  outlined.  More  than  thirty  taxa  are  described  including  five  new 
species:  Praenucula  dispersa,  P.  infirma,  P.  praetermissa,  Ambonychia  arundinea,  and  Cleionychia  incognita. 
The  rostroconch  Hippocardia  praepristis  (Reed)  is  illustrated. 


Since  Portlock  (1837,  1843)  first  described  the  faunas  of  the  Caradoc-Ashgill  rocks  of  the  Pomeroy 
area  of  County  Tyrone  (text-fig.  1),  some  elements  have  been  studied  thoroughly,  especially  the 
trilobites  and  brachiopods,  which  have  helped  to  show  the  relationship  of  the  Pomeroy  strata  to 
the  Ordovician  of  North  America  and  southern  Scotland  (Mitchell,  1977,  and  others).  Although 
molluscs  are  abundant  in  the  rocks  of  Pomeroy,  they  have  received  scant  attention  in  recent  years. 
The  orthoconic  nautiloids  were  redescribed  by  Blake  (1882)  but  have  remained  unrevised  since; 
the  gastropods  have  been  described  by  Longstaff  ( = Donald)  (1902,  1924)  and  Reed  (1952).  The 
bivalves  were  largely  ignored  until  Reed  (1952)  redescribed  Portlock’s  species  and  added  a number 
of  new  taxa.  Reed  was  unaware  of  Portlock’s  many  duplicate  specimens,  including  syntypes,  now 
housed  in  the  Ulster  Museum  (Tunnicliff  1980),  and  did  not  have  the  benefit  of  collections  made 
in  recent  years  by  the  Geological  Survey  of  Northern  Ireland,  by  Dr.  W.  I.  Mitchell  and  by  Mr. 
R.  P.  Tripp  and  others  of  the  British  Museum  (Natural  History).  With  the  aid  of  this  new  material 
a critical  re-examination  of  the  bivalve  fauna  has  been  carried  out  and  several  new  forms  have 
been  recognized. 


History  of  Pomeroy  bivalve  research 


Portlock  (1843)  mentioned  or  described  in  varying  detail  the  following  species,  arranged  here  according  to 
the  formation  from  which  they  can  now  be  recognized  to  have  come. 


Bardahessiagh  Formation  (Caradoc): 
Avicula  obliqua  Murchison 
Avicula  orbicularis  Murchison 
Modiola  Brycei  Portlock 
Modiola  expansa  Portlock 
Modiola  Nerei  (Munster) 


Modiola  securiformis  Portlock 
Inoceramus  priscus  Portlock 
Inoceramus  transversus  Portlock 
Inoceramus  trigonus  (Munster) 
Nucula  radiata  Portlock 


Killey  Bridge  Formation  (Ashgill,  Cautleyan): 


Inoceramus  contortus  Portlock 
Cypricardia  simplex  Portlock 
Area  cylindrica  Portlock 
Area  dissimilis  Portlock 
Area  Eastnori  (Murchison) 
Area  obliqua  Portlock 
Area  regularis  Portlock 


Area  subtruncata  Portlock 

Area  transversa  Portlock 

Pectunculus  ambiguus  Portlock 

Pectunculus  Apjohnni  Portlock 

Pectunculus  semitruncatus  Portlock 

Nucula  acutal  (Sowerby)  var.  imbricata  Portlock 

Uncites  gryphus  Schlotheim  var. 


In  addition  to  these,  Portlock  had  a number  of  specimens  to  which  he  did  not  refer  in  print  but  to  which 
he  fixed  labels  with  suggested  identifications  and  observations.  Notes  on  how  Portlock  labelled  his  specimens 
are  given  by  Tunnicliff  (1980).  His  Mytilus  cinctus  Portlock  and  Posidonomyal  venusta  (Munster)  are  from 
rocks  of  Llandovery  age  as  is  his  Mytilus?  semi-rugatus  Portlock  from  Lisbellaw,  Co.  Fermanagh.  Portlock’s 


[Palaeontology,  Vol.  25,  Part  1,  1982,  pp.  43-88,  pis.  7-13.] 


44 


PALAEONTOLOGY,  VOLUME  25 


text-fig.  1.  The  area  around  Pomeroy,  Co.  Tyrone,  based  on  Mitchell  1977  (taken  from  Tunnicliff  1980),  with 
Ireland  inset  to  show  the  position  of  Pomeroy. 


figures,  although  generally  of  natural  size,  often  lack  the  detail  which  might  be  expected  in  modern  illustrations. 
Many  of  his  descriptions  contain  useful  detail  such  as  height  and  length  of  valves  and  the  number  of  teeth 
visible,  but  others  are  less  helpful. 

Describing  specimens  from  the  collection  of  Sir  R.  Griffith,  McCoy  (1846)  added  the  following  species 
from  the  Bardahessiagh  Formation,  Pullastra  speciosa  McCoy  and  from  the  Killey  Bridge  Formation,  Nucula 
protei  Munster,  N.  subacuta  McCoy,  Area  quadrata  McCoy. 

In  1878  (p.  28)  Baily  published  a list  of  Portlock  bivalve  species  giving  amended  generic  names  and  gave 
a list  (p.  25)  of  the  species  collected  by  officers  of  the  Geological  Survey  of  Ireland  during  the  survey  of  the 
1870s. 

Apart  from  some  brief  discussion  by  Hind  (1910),  Baily’s  was  the  last  contribution  to  the  study  of  Pomeroy 
Ordovician  bivalves  before  the  publication  of  Reed’s  posthumous  paper  in  1952.  In  this  Reed  redescribed 
most  of  Portlock’s  species  and  added  the  following: 


Ctenodonta  perangulata  Reed 
Ctenodonta  deserta  Reed 
Ctenodonta  cf.  nitida  (Ulrich) 
Clidophorus  occultus  Reed 
Vanuxemia?  suspecta  Reed 
Modiolopsis  concentrica  simulans  Reed 
Orthodesma  tyronense  Reed 
Conocardium  praepristis  Reed 


Paramodiolal  sp. 

Whiteavesia  subexpansa  Reed 
Ambonychinia  cf.  volvens  Isberg 
Ambonychinia  cf.  amygdalina  (Hall) 
Ambonychinia  cf.  intermedia  Isberg 
Clionychia  subovalis  Reed 
Clionychia  subquadrata  Reed 


TUNNICLIFF:  LATE  ORDOVICIAN  BIVALVES 


45 


Horizons  and  localities  (Text-figs.  1 and  2) 

The  Ordovician  stratigraphy  of  Pomeroy  was  discussed  by  Mitchell  (1977,  pp.  4-18).  Bivalves  are  recorded 
from  the  following  strata: 

Tirnaskea  Formation  (Ashgill,  Hirnantian) 

Killey  Bridge  Formation  (Ashgill,  Cautleyan) 

Bardahessiagh  Formation  (Caradoc). 

No  bivalves  are  recognized  as  being  from  Mitchell’s  Junction  beds. 

The  horizon  of  Portlock  specimens  and  other  material  from  old  collections  is  inferred  from  the  lithologies 
and  from  the  known  locality  information  (see  Tunnicliff  1980). 

The  localities  of  many  of  the  specimens  in  the  old  collections  are  uncertain,  often  given  only  as  ‘Pomeroy’. 
However,  by  comparison  with  material  in  the  recent  well-localized  collections  and  with  the  small  number  of 
likely  sites,  it  is  possible  to  suggest  localities  for  most  specimens,  usually  with  reference  to  those  listed  by 
Mitchell  (1977,  pp.  5-7).  Portlock’s  localities  are  discussed  by  Tunnicliff  (1980). 

Stratigraphic  distribution  (Table  1) 

The  bivalve  fauna  of  the  Bardahessiagh  Formation  is  dominated  by  cyrtodontids  and  ambonychiids,  the 
former  probably  infaunal  or  semi-infaunal,  the  latter  epifaunal  and  byssate.  Cyrtodontal  often  occurs  as 
conjoined  valves  suggestive  of  little  or  no  transport  after  death.  The  other  infaunal  genera  represented  in  the 
Bardahessiagh  Formation  are  Praenucula,  Deceptrix,  Similodontal , and  Lyrodesma  all  of  which  are  uncommon 
compared  with  Cyrtodontal . Similodontal  and  Lyrodesma  both  occur  as  conjoined  or  closely  associated  valves 


Series  Graptolite 

& Stage  zone 


ASHGILL 

Hirnantian 


Rawtheyan 


Cautleyan 


D. 


anceps 


Pusgillian 


CARADOC 

Onnian 

Actonian 

Marshbrookian 


D.  complanatus 
P.  linearis 


Woolstonian  D.  clingani 


Longvi  Ilian 


Soudleyan  C.  wilsoni 


Harnagian  C.  peltijer 


Costonian 

N.  gracilis 


Pomeroy  Girvan  Bohemia  North 

America 


Tirnaskea 

Formation 


High  Mains 


Kosov 

Formation 


Killey 

Bridge 

Formation 


Drummuck 

Group 


Kraluv  d5 

Dvur 

Formation 


Richmondian 


Shalloch 

Formation 


Maysvillian 


Whitehouse 

Group  Edenian 

_ _ Bohdalec 

Formation 


Junction  Ardwell 

Beds  Group 

Bardahessiagh 

Formation 


Balclatchie 


Zahorany  F.  y 
Vinice  F.  ,33 

Letna  ^2 

Formation 


Shermanian 


Wildernessian 


text-fig.  2.  Approximate  correlation  of  the  Ordovician  rocks  of  Pomeroy  based  on  Williams  et  al.  (1972), 
Mitchell  (1977),  and  R.  P.  Tripp  and  J.  K.  Ingham  (pers.  comm.  1981).  The  approximate  stratigraphic 
equivalents  of  Barrande’s  stages  d2,  d3,  d4,  and  d5  in  Bohemia  are  shown  (based  on  Krtz  and  Pojeta  1974). 


46 


PALAEONTOLOGY,  VOLUME  25 


table  1.  The  stratigraphic  distribution  of  bivalves  from  the  Ordovician  rocks  of  Pomeroy,  Co.  Tyrone 


Bardahessiagh 

Formation 

Killey  Bridge 
Formation 

Tirnaskea 

Formation 

Praenucula  dispersa  sp.  nov. 

X 

P.  praetermissa  sp.  nov. 

X 

P.  aff.  praetermissa  sp.  nov. 

X 

P.  infirma  sp.  nov. 

X 

Deceptrix  sp. 

X 

D.  apjohnni  (Portlock) 

X 

D.  regular  is  (Portlock) 

X 

D.  semitruncata  (Portlock) 

X 

D.  subtruncata  (Portlock) 

X 

Similodonta?  sp. 

X 

Concavodonta  imbricata  (Portlock) 

X 

Nuculites  cylindricus  (Portlock) 

X 

Nuculoid  gen.  and  sp.  nov. 

X 

Cyrtodonta?  expansa  (Portlock) 

X 

C.?  securiformis  (Portlock) 

X 

C.7  sp. 

X 

Vanuxemia  contorta  (Portlock) 

X 

Ambonychia  arundinea  sp.  nov. 

X 

Cleionychia  transversa  (Portlock) 

X 

C.  prisca  (Portlock) 

X 

C.  incognita  sp.  nov. 

X 

Ambonychiopsis  suspecta  (Reed) 

X 

pterineid?  gen.  and  sp.  indet. 

X 

? bivalve  gen.  and  sp.  indet. 

X 

Modiolopsis  sp. 

X 

Corallidomus  concentrica  (Hall  and  Whitfield) 

X 

Corallidomus?  sp. 

X 

Goniophora  sp. 

X 

Colpomya  simplex  (Portlock) 

X 

Semicorallidomus?  sp. 

X 

Cycloconcha?  speciosa  (McCoy) 

X 

Lyrodesma  radiatum  (Portlock) 

X 

Hippocardia  praepristis  (Reed) 

? 

but,  like  Cyrtodonta?,  are  never  suggestive  of  life  position  but  rather  of  having  been  winnowed  out  from  the 
substrate.  Of  the  epifaunal  ambonychiids,  Cleionychia  is  best  represented,  and  some  also  occur  as  conjoined 
valves.  The  nature  of  the  present-day  Bardahessiagh  Formation  exposure  (Mitchell  1977,  p.  5)  precludes 
identifying  the  exact  localities  for  most  specimens. 

In  the  Killey  Bridge  Formation  the  principal  elements  of  the  bivalve  fauna  are  infaunal  nuculoids,  Praenucula, 
Deceptrix,  Concavodonta,  and  Nuculites.  These  are  all  numerous  compared  with  other  genera,  most  of  which 
are  represented  in  the  collections  by  only  one  or  two  specimens.  Most  specimens  from  the  Killey  Bridge 
Formation  are  single  valves,  but  conjoined  valves  of  all  nuculoid  genera  are  present.  Bivalves  are  recorded 
from  all  the  localities  where  the  Killey  Bridge  Formation  is  represented  by  dark-grey  calcareous  mudstone 
(e.g.  localities  1-3  of  Mitchell  1977)  but  determinable  bivalves  are  not  known  from  the  rottenstone  and 
sandstone  lens  recorded  by  Mitchell  (localities  4a,  b of  Mitchell  1977)  in  which  brachiopods  are  abundant 
and  varied. 

Determinable  bivalves  are  also  almost  unknown  from  the  Tirnaskea  Formation;  only  Nuculoid  gen.  nov. 
and  the  rostroconch  Hippocardia  (by  Reed  1952)  are  recorded  and  the  latter  is  more  probably  from  the  Killey 
Bridge  Formation.  Mitchell  (1977,  p.  17)  compared  the  Tirnaskea  fauna  with  the  Hirnantian  Hirnantia  fauna. 


TUNNICLIFF:  LATE  ORDOVICIAN  BIVALVES 


47 


The  differences  between  the  Bardahessiagh  and  Killey  Bridge  bivalve  faunas  are  a reflection  of  their  differing 
palaeoenvironments;  the  generally  sandy  Bardahessiagh  Formation  probably  represents  a shallow-shelf 
environment  favouring  the  modes  of  life  of  cyrtodontids  and  ambonychiids,  while  the  mudstones  of  the  Killey 
Bridge  Formation  were  probably  the  product  of  a more  stable  deeper-shelf  environment.  The  Bardahessiagh 
Formation  bivalve  fauna  has  a distinctly  eastern  American  aspect  which  is  less  obvious,  but  still  present  in 
the  Killey  Bridge  Formation. 


SYSTEMATIC  PALAEONTOLOGY 

In  general,  the  classification  used  in  the  Treatise  of  Invertebrate  Palaeontology,  N1  and  2 (Moore  1969)  is 
adopted.  Where  this  is  deviated  from  is  stated  in  this  text,  and  the  author  whose  precedent  is  followed  is  given. 

Measurements.  In  some  cases  tables  of  measurements  are  given  in  which  the  terms  mean  and  median  are 
used.  Mean  is  used  to  show  the  sum  of  the  items  divided  by  the  number  of  items;  median  is  used  to  convey 
the  mid-point  between  the  maximum  and  minimum  values  of  a group  of  items.  All  linear  measurements  are 
given  in  millimetres  (mm),  measured  with  a vernier  caliper  or  under  a binocular  microscope  with  a graticule. 
Abbreviations  used  are  L for  length,  H for  height,  AL  for  preumbonal  length.  In  tables  of  measurements, 
the  term  ‘angle’  is  used  for  the  angle  between  the  anterior  and  posterior  hinge-plates  in  nuculoids. 

Preservation.  Two  terms  used  require  a note  of  explanation.  Composite  mould  is  used  in  the  sense  of  McAlester 
1962.  External  cast  is  used  to  describe  a naturally  occurring  convex  specimen  which  shows  only  external 
features  but  which  has  probably  been  produced  in  a similar  manner  to  composite  moulds. 

This  study  is  based  on  material  in  the  following  museums:  Institute  of  Geological  Sciences  (IGS),  British 
Museum  (Natural  History)  (BM),  Sedgwick  Museum,  Cambridge  (SM),  Ulster  Museum,  Belfast  (UM), 
National  Museum  of  Ireland,  Dublin  (NMI). 

Class  bivalvia  Linnaeus,  1758 
Subclass  palaeotaxodonta  Korobkov,  1954 
Order  nuculoidea  Dali,  1 889 
Superfamily  nuculacea  Gray,  1824 
Family  praenuculidae  Pfab,  1934 

Remarks.  Three  genera  of  Praenuculids  are  certainly  recognized  from  Pomeroy:  Praenucula  and 
Deceptrix  from  the  Bardahessiagh  and  Killey  Bridge  Formations  and  Concavodonta  from  the  Killey 
Bridge  Formation.  Concavodonta  is  distinct  from  the  other  two  genera  (see  discussion  of 
Concavodonta)  but  Praenucula  and  Deceptrix  are  less  easily  distinguished  at  first  sight.  Pojeta  (1971 , 
p.  16)  implied  that  they  were  synonymous  (‘. . . some  authors  recognize  Praenucula  Pfab  as  being 
distinct  from  Deceptrix. . .’).  Pojeta,  Kriz,  and  Berdan  (1976)  treated  Praenucula  as  a subgenus  of 
Deceptrix.  McAlester  (1969,  p.  229),  placed  Pfab’s  Praeleda  in  synonomy  with  Deceptrix  and  both 
have  been  used  for  species  which  show  the  characters  of  Praenucula , for  example  Praeleda  ciae 
(Sharpe)  of  Bradshaw  1970,  and  Deceptrix  cf.  ciae  (Sharpe)  of  Babin  and  Robardet  1973.  Praenucula 
and  Deceptrix  have  remained  virtually  unused  in  the  British  Ordovician  and  I can  find  no  reference 
to  their  occurrence  in  the  British  Isles,  except  Praenucula  from  the  Arenig  of  Ramsey  Island  (Carter 
1971,  p.  251).  Specimens  from  British  localities  which  could  now  be  placed  in  Praenucula  or 
Deceptrix  have  usually  been  recorded  as  Ctenodonta,  often  under  open  nomenclature. 

In  the  species  discussed  here,  distinction  is  made  between  Praenucula  and  Deceptrix  principally 
on  the  differences  indicated  by  McAlester  (1969,  p.  229);  that  is  that  the  posterior  teeth  in  Deceptrix 
are  smaller  and  more  numerous  than  the  anterior  teeth,  while  in  Praenucula  the  anterior  and 
posterior  teeth  are  similar  in  size  and  number.  To  this  we  may  add  that  the  umbones  in  Praenucula 
lie  in  the  posterior  half  while  in  Deceptrix  they  generally  lie  in  the  anterior  half.  Corresponding 
to  the  greater  posterior  length  in  Deceptrix  is  the  relatively  greater  length  (equal  to  or  longer  than 
the  anterior)  of  its  posterior  hinge-plate.  In  Praenucula  the  posterior  hinge-plate  is  equal  to  or 
shorter  in  length  than  the  anterior  hinge-plate.  In  both  genera  the  anterior  hinge-plate  is  arched 
towards  the  body  of  the  shell  and  the  posterior  hinge-plate  is  straight  or  gently  arched  away  from 
I the  body  of  the  shell.  The  adductor  muscle  scars  in  both  genera  are  equal  or  subequal  but  their 


PALAEONTOLOGY,  VOLUME  25 


relative  sizes  and  positions  differ,  being  larger  and  more  ventral  in  Deceptrix  (Pis.  7-9).  In  Praenucula 
the  adductor  muscle  scars  are  close  to,  and  may  extend  beyond,  the  ends  of  the  hinge-plates,  while 
in  Deceptrix  the  posterior  scar  lies  closer  to  the  umbo,  only  extending  as  far  as  the  end  of  the 
hinge-plate,  and  the  anterior  scar  lies  in  the  most  anterior  part  of  the  shell,  beyond  the  end  of  the 
hinge-plate. 

A fourth  genus,  Similodonta,  may  be  represented  by  specimens  from  the  Bardahessiagh 
Formation. 


Genus  praenucula  Pfab,  1934 

Type  species.  By  original  designation  of  Pfab  1934,  pp.  234-235,  Praenucula  dispar  expansa  Pfab,  1934,  from 
the  Sarka  Formation  (Llanvirn)  of  Bohemia  (Havlicek  and  Vanek  1966)  and  described  and  figured  by 
McAlester  (1968,  p.  46,  pi.  8,  figs.  3-9).  See  discussion  under  Praenuculidae  above. 


Praenucula  dispersa  sp.  nov. 

Plate  7,  figs.  1-5;  text-fig.  3 a 

Derivation  of  name.  From  Latin  dispersus  scattered,  referring  to  the  occurrence  in  loose  blocks. 

Type  specimens.  Holotype  IGS  GU  1893,  internal  mould  of  right  valve,  and  paratypes  IGS  GU  1354,  1894, 
1897,  1906,  3265,  UM  K4203,  SM  A 16296,  3 left  and  4 right  valves;  one  a composite  mould,  the  remainder 
internal  moulds. 

Horizon  and  locality.  All  specimens  are  from  the  Bardahessiagh  Formation;  GU  specimens  from  the  area  of 
Mitchell’s  Bardahessiagh  Formation  collecting  (Mitchell  1977),  the  others  from  uncertain  localities  to  the 
south  of  Craigbardahessiagh. 


Measurements 

L 

H 

H/L 

AL/L 

Angle 

Max. 

13-2  mm 

110  mm 

0-87 

0-65 

155° 

Min. 

7-8  mm 

6-8  mm 

0 61 

0-50 

130° 

Mean 

10-62  mm 

8-15  mm 

0-77 

0-50 

145° 

Median 

10-5  mm 

8-9  mm 

0-74 

0-58 

142-5‘ 

Description.  Ovate  Praenucula  with  height  about  0-7  of  the  length  but  variable.  Umbo  slightly  posterior  of 
centre,  between  the  posterior  two-fifths  and  mid-point,  and  extending  slightly  above  the  hinge-line.  Convexity 
greatest  slightly  anterior  to  the  umbo,  maximum  inflation  of  a single  valve  2-4  mm  seen  in  a valve  1 1 mm 


a 


b 


c 


text-fig.  3.  Praenucula  spp.  based  on  internal  moulds,  c.  x 5.  ( a ) P.  dispersa  sp.  nov.,  IGS  GU  1906,  right 
valve.  ( b ) P.  infirma  sp.  nov.,  IGS  NIL  8935,  right  valve,  (c)  P.  praetermissa  sp.  nov.,  IGS  NIL  8757,  latex  of 
left  valve.  Abbreviations:  aa— anterior  adductor  muscle  scar,  pa— posterior  adductor  muscle  scar,  a— accessory 
muscle  scar.  Compare  with  PI.  7,  figs.  5,  12,  21.  Note  the  differences  in  size  of  the  adductor  muscle  scars  and  the 
differences  in  the  size  of  the  teeth  and  in  their  distribution  on  either  side  of  the  umbo.  In  each  case,  the  finest  teeth 
below  the  umbo  are  too  small  to  be  represented  clearly. 


TUNNICLIFF:  LATE  ORDOVICIAN  BIVALVES 


49 


long  and  8 mm  high.  Angle  between  anterior  and  posterior  hinge-plates  about  140  to  145°.  Posterior  hinge-plate 
straight  with  12+  uniformly  sized  teeth.  Anterior  hinge-plate  concave  towards  the  body  of  the  shell,  with 
13+  teeth,  increasing  in  size  forwards.  Slightly  chevron-shaped  teeth  passing  beneath  umbo  in  a continuous 
series,  chevrons  directed  towards  the  umbo.  Anterior  and  posterior  muscle  scars  clearly  distinguished,  equal 
in  size  the  anterior  more  deeply  impressed  than  the  posterior  and  most  deeply  impressed  at  its  posterior  edge. 
Accessory  muscles  known  (text-fig.  3a).  Ligament  and  shell  material  unknown.  Sculpture  of  fine  concentric 
lines  (c.  4 per  mm)  (PI.  7,  fig.  3). 

Discussion.  Nuculoid  bivalves  are  poorly  represented  in  the  Bardahessiagh  Formation,  and  those 
that  are  found  are  not  well  preserved.  In  particular,  the  very  soft  nature  of  some  of  the  rottenstone 
has  led  to  damage  to,  or  loss  of,  the  hinge-plate  and  teeth.  The  deformation  apparent  in  other 
species  from  the  Bardahessiagh  Formation  (e.g.  Cyrtodonta!  spp.)  appears  to  have  produced  some 
variation  among  the  Praenucula  specimens;  however,  it  is  reasonable  to  regard  them  as  belonging 
to  a single  species. 

Praenucula  dispersa  is  distinguished  from  the  Killey  Bridge  Formation  species,  Praenucula 
praetermissa  and  P.  infirma,  by  having  smaller,  less  strongly  chevron-shaped  teeth.  In  the  same 
way  it  can  be  differentiated  from  P.  costae  (Sharpe)  as  described  by  Bradshaw  (1970,  pp.  630-633) 
from  the  Llandeilo  of  Finistere.  P.  dispersa  has  more  teeth  then  P.  ciae  (Sharpe)  (of  Bradshaw 
1970,  pp.  633-636)  and  Deceptrix  pulchra  armoricana  Babin  and  Melou  (1972,  pp.  81-82,  pi.  8, 
figs.  4-7),  and  it  lacks  the  posterior  lobe  seen  in  Hind’s  Drummuck  species  Nuculana  lobata  (1910, 
p.  519,  pi.  4,  figs.  1-3)  and  is  less  elongate  than  his  N.  curta  (1910,  p.  521,  pi.  4,  figs.  10-14)  which 
is  also  from  the  Drummuck  Group  (Ashgill). 

Ctenodonta  albertina  Ulrich  and  C.  filistriata  Ulrich  (1894,  pp.  598-599,  pi.  42,  figs.  76-82, 
text-fig.  44),  both  from  the  Cincinnatian,  have  a pit  beneath  the  umbo  according  to  Ulrich’s 
descriptions.  Recent  illustrations  (Pojeta  1971,  pi.  5,  figs.  14-16;  1978,  pi.  2,  figs.  1-2)  show  a 
discontinuity  in  the  dentition  beneath  the  umbo  not  apparent  in  P.  dispersa.  Nucula  dispar  Barrande 
(1881,  pi.  273,  VII,  figs.  1-14)  is  more  rounded  than  P.  dispersa. 

Praenucula  praetermissa  sp.  nov. 

Plate  7,  figs.  16,  21,  24;  text-fig.  3c 

Derivation  of  name.  From  the  Latin  praetermittere,  to  overlook,  referring  to  the  fact  that  the  species  has  been 
overlooked  by  previous  authors. 

Type  specimens.  Holotype,  IGS  NIL  8757,  composite  mould  of  a left  valve  from  the  Killey  Bridge  Formation 
of  the  Crocknagargan  Stream  section  (locality  1 of  Mitchell  1977),  Pomeroy,  Co.  Tyrone.  Paratypes:  6 left 
and  6 right  valves,  internal,  external,  and  composite  moulds:  IGS  NIL  8687,  8747-8748,  8755,  8768,  8893, 
8917,  8986,  8987  all  from  the  same  locality  as  the  holotype;  IGS  NIG  482,  Zs  2758  from  locality  3 of  Mitchell 
1977,  and  IGS  GU  1451  from  his  locality  2 (Warren  Wood  River).  All  from  the  Killey  Bridge  Formation. 


Measurements 

L 

H 

H/L 

AL/L 

Angle 

Max. 

110  mm 

9-2  mm 

0-83 

0-70 

160° 

Min. 

5-4  mm 

3-6  mm 

0-50 

0-50 

130° 

Mean 

8-6  mm 

5-71  mm 

0-72 

0-60 

143° 

Median 

8-2  mm 

6-40  mm 

0-66 

0-60 

145° 

Description.  Subovate  to  circular  Praenucula  in  which  the  height  is  about  0-7  of  the  length.  Maximum  inflation 
of  a single  valve  1-4  mm  seen  in  a valve  110  mm  long  and  9-2  mm  high.  The  umbo  is  at  about  the  posterior 
two-fifths.  The  shell  is  truncate  behind  and  rounded  in  front.  The  anterior  hinge-plate  is  broad  and  strongly 
arched  towards  the  body  of  the  shell.  The  anterior  and  posterior  hinge-plates  meet  at  an  angle  of  about  145°, 
and  usually  bear  the  same  number  of  teeth,  up  to  nine.  The  teeth  of  the  anterior  hinge-plate  are  the  larger 
I and  are  strongly  chevron-shaped;  those  of  the  posterior  hinge-plate  are  markedly  less  so.  The  adductor  muscle 
scars  are  subequal  in  size,  the  anterior  scar  being  the  more  strongly  impressed  and  lying  below  and  slightly 
anterior  to  the  anterior  teeth  with  a diameter  about  one-quarter  of  the  valve  height;  the  posterior  scar  lies 
immediately  below  the  end  of  the  posterior  dental  series.  Accessory  muscle  scars  unknown,  except  for  one 
anterior  scar.  Sculpture  is  of  fine  concentric  lines,  c.  16  per  mm. 


50 


PALAEONTOLOGY,  VOLUME  25 


Discussion.  This  is  a most  distinctive  form  within  the  Pomeroy  fauna,  which,  surprisingly,  does 
not  seem  to  have  been  seen  by  previous  workers  (Portlock,  McCoy,  Reed)  and  yet  is  well  represented 
in  recent  collections,  especially  as  fragments.  Praenucula  praetermissa  is  distinguished  from  P. 
infirma,  described  herein  from  the  Killey  Bridge  Formation,  by  its  larger  adductor  muscle  scars 
and  its  more  numerous  posterior  teeth,  and  from  P.  dispersa  from  the  Bardahessiagh  Formation 
by  its  larger  and  strongly  chevron-shaped  teeth.  P.  praetermissa  lacks  the  lobate  posterior  of  Hind’s 
Nuculana  lobata  (1910,  p.  519,  pi.  4,  figs.  1-3)  and  is  less  elongate  than  his  N.  curta  (1910,  p.  521, 
pi.  4,  figs.  10-14),  both  from  the  Drummuck  Group  (Ashgill).  It  also  lacks  the  pit  beneath  the 
umbo  recorded  in  N.  curta  and  in  the  American  species  Ctenodonta  albertina  Ulrich  and  C.filistriata 
Ulrich  (1894,  pp.  598-599,  pi.  42,  figs.  76-82,  text-fig.  44). 

It  may  be  compared  with  other  Ordovician  Praenucula  described  from  Europe:  Deceptrix  pulchra 
armoricana  Babin  and  Melou,  1972  ( = D . cf.  ciae  (Sharpe)  of  Babin  and  Robardet  1973)  from  the 
Caradocian  rocks  of  Normandy,  has  a coarser  sculpture  than  P.  praetermissa,  c.  6-8  lines  per  mm, 
grouped  into  ‘faisceaux’  (Babin  and  Melou  1972,  p.  82,  pi.  7,  fig.  7).  Praeleda  pulchra  Pfab,  1934, 
is  recorded  by  Havlicek  and  Vanek  1966,  from  the  Llanvirn,  Caradoc,  and  Ashgill  of  Bohemia, 
and  in  shape  resembles  Praenucula  praetermissa  (Pfab  1934,  p.  234,  pi.  3,  fig.  6)  but  Pfab  gave  no 
detail  of  dentition  or  hinge,  and  no  sculpture  is  evident  in  his  figure.  Nucula  costae  (Sharpe)  from 
Portugal,  according  to  Sharpe’s  description  (1853,  pp.  148-149,  pi.  9,  fig.  4a,  b ) has  more  teeth 
than  P.  praetermissa,  about  20  (given  as  25-30  at  one  point  and  as  a total  of  about  18  elsewhere) 
divided  unequally  between  posterior  and  anterior,  and  shows  the  grouping  of  concentric  striae 
also  described  in  D.  armoricana.  The  forms  described  by  Bradshaw  (1970,  pp.  630-633)  as  Praeleda 
costae  (Sharpe)  from  the  Llandeilo  of  Finistere  are  closer  in  most  respects  to  Praenucula  praetermissa 


EXPLANATION  OF  PLATE  7 

Figs.  1-5.  Praenucula  dispersa  sp.  nov.  1,  2,  oblique  dorsal  view  and  lateral  view,  internal  mould  of  right 
valve,  holotype,  IGS  GU  1893.  3,  ?composite  mould  of  left  valve,  UM  K4203.  4,  internal  mould  of  left 

valve,  paratype,  IGS  GU  3265.  5,  internal  mould  of  right  valve,  paratype  IGS  GU  1906.  All  from  the 
Bardahessiagh  Formation  (Caradoc),  south  of  Craigbardahessiagh,  Pomeroy,  x 3. 

Figs.  6-15,  17-19.  Praenucula  infirma  sp.  nov.  6,  7,  oblique  dorsal  and  lateral  views,  internal  mould  of  left 
valve,  IGS  NIL  8824,  Crocknagargan  Stream  section,  Pomeroy  (IGR  C.H721737).  8, 15,  internal  mould  and 

latex  cast  of  external  mould  of  right  valve,  IGS  NIL  8682,  8682  pars,  ditch  exposure  south  of  Craigbarda- 
hessiagh, Pomeroy  (IGR  H7195  7385).  9,  10,  dorsal  view  of  external  mould  and  dorsal  view  of  internal 
mould,  conjoined  valves,  IGS  NIL  8938,  Crocknagargan  Stream  section,  Pomeroy  (IGR  C.H721737).  11, 

12,  13,  dorsal  and  lateral  views  and  latex  cast  of  internal  mould  of  right  valve,  holotype,  IGS  NIL  8935, 
Crocknagargan  Stream  section,  Pomeroy  (IGR  C.H721737).  14,  internal  mould  of  left  valve,  IGS  NIL  8754, 

Crocknagargan  Stream  section,  Pomeroy  (IGR  C.H721737),  x 3.  17,  18,  19,  internal  mould  of  left  valve, 

IGS  NIL  8444,  Tirnaskea  Stream  section,  Pomeroy  (IGR  C.H727725);  17,  latex  cast,  x 6;  18,  oblique  dorsal  j 
view,  x3;  19,  lateral  view,  x3.  All  Killey  Bridge  Formation  (Ashgill,  Cautleyan),  x3. 

Figs.  16,  21,  24.  Praenucula  praetermissa  sp.  nov.  16,  internal  mould  of  right  valve,  IGS  NIL  8917, 
Crocknagargan  Stream  section,  Pomeroy  (IGRc.H721737),  x 3.  21, 24,  left  valve  internal  mould,  holotype, 
IGS  NIL  8757,  Crocknagargan  Stream  section,  Pomeroy  (IGR  C.H721737);  21,  latex  cast,  x 6;  24,  lateral 
view,  x 3.  All  Killey  Bridge  Formation  (Ashgill,  Cautleyan). 

Fig.  20.  Praenucula  aff.  praetermissa  sp.  nov.  internal  mould  of  left  valve,  IGS  NIL  8972,  Killey  Bridge 
Formation  (Ashgill,  Cautleyan),  Crocknagargan  Stream  section,  Pomeroy  (IGR  C.H721737),  x3. 

Figs.  22,  23.  Praenucula  aff.  praetermissa  sp.  nov.  Oblique  dorsal  view  and  lateral  view  of  crushed  internal 
mould  of  left  valve,  IGS  NIL  8934,  Killey  Bridge  Formation  (Ashgill,  Cautleyan),  Crocknagargan  Stream 
section,  Pomeroy  (IGR  C.H721737),  x 3. 

Fig.  25.  Similodonta?  sp.,  composite  mould  of  left  valve  showing  fine  concentric  ornament  and  anterior 
adductor  muscle  scar,  UM  K4241,  Bardahessiagh  Formation  (Caradoc),  south  of  Craigbardahessiagh, 
Pomeroy,  x 3. 


PLATE  7 


TUNNICLIFF,  Ordovician  bivalves 


52 


PALAEONTOLOGY,  VOLUME  25 


but  also  tend  to  have  more  teeth  (17-22)  and  no  detail  of  the  sculpture  is  given.  N.  ciae  Sharpe 
(1853,  p.  149,  pi.  9 fig.  5 a-c)  from  Portugal  has  a similar  number  of  teeth  (c.  19)  and  is  described 
as  having  fine  concentric  lines  like  P.  praetermissa.  However,  P.  praetermissa  differs  from  Bradshaw’s 
(1970,  pp.  633-636)  Praeleda  ciae  (Sharpe)  in  the  posterior  teeth:  in  Bradshaw’s  P.  ciae  and  in 
Babin’s  Ctenodonta  ciae  (Sharpe)  (Babin  1966,  pp.  49-52,  text-fig.  1 1)  the  posterior  chevron-shaped 
teeth  seem  to  point  away  from  the  umbo,  in  Praenucula  praetermissa  the  chevrons  point  towards 
the  umbo.  In  this  respect  P.  praetermissa  resembles  Sharpe’s  N.  ciae. 

A single  internal  mould  of  a large  left  valve  (IGS  NIL  8972;  PI.  7,  fig.  20)  from  the  Killey  Bridge 
Formation  of  the  Crocknagargan  Stream  Section  is  like  P.  praetermissa  in  shape,  but  is  16  mm 
long  and  13  mm  high  with  a maximum  inflation  of  4 mm.  The  hinge-plates  meet  at  an  angle  of 
145°  with  13  anterior  and  15  posterior  teeth  meeting  below  the  umbo.  Its  musculature  is  indistinct 
and  sculpture  unknown.  Because  of  its  large  size  and  more  central  umbo,  it  is  here  recorded  as  P. 
aff.  praetermissa. 

A second  specimen  (IGS  NIL  8934;  PI.  7,  figs.  22,  23)  from  the  same  horizon  and  locality  is 
also  recorded  as  P.  aff.  praetermissa.  It  is  large,  1 1-8  mm  long,  but  crushed  dorso-ventrally.  The 
umbo  is  at  the  mid-point  and  the  hinge-plates  meet  at  an  angle  of  165°  with  12  anterior  and  14  + 
posterior  chevron-shaped  teeth  meeting  below  the  umbo.  The  adductor  muscle  scars  are  round 
and  equal  in  size,  about  one-quarter  of  the  valve  height  in  diameter  and  each  is  close  below  and 
slightly  anterior  to  the  end  of  the  corresponding  dental  series.  The  sculpture  is  unknown.  It  differs 
from  P.  praetermissa  in  its  more  central  umbo  and  in  having  more  teeth  corresponding  to  its  large 
size. 

The  collections  at  IGS  contain  a specimen  (IGS  PT  9049)  from  the  Corona  Beds,  Dufton 
Shales  (Caradoc,  Longvillian)  which  is  comparable  with  both  P.  praetermissa  and  P.  pulchra 
armoricana. 


Praenucula  infirma  sp.  nov. 

Plate  7,  figs.  6-15,  17-19;  text-fig.  3 b 

Derivation  of  name.  From  the  Latin  infirma,  weak,  referring  to  the  small  size  of  the  adductor  muscle 
scars. 

Type  specimens.  Holotype,  IGS  NIL  8935,  an  internal  mould  of  a right  valve,  from  the  Killey  Bridge  Formation 
of  the  Crocknagargan  Stream  section  (locality  1 of  Mitchell  1977)  Pomeroy,  Co.  Tyrone.  Paratypes:  9 left  and 
5 right  valves,  internal,  external,  and  composite  moulds:  IGS  NIL  8754,  8772,  8824,  8828,  8938,  8952  all 
from  the  same  locality  as  the  holotype;  IGS  NIL  8444-8445  from  the  Tirnaskea  Stream  section;  IGS  Zs 
2729,  2737-2738  from  locality  3 of  Mitchell  1977;  IGS  NIL  8682,  8682  pars  from  a ditch  exposure  at  IGR 
H 7195  7385.  All  from  the  Killey  Bridge  Formation. 


Measurements 

L 

H 

H/L 

AL/L 

Angle 

Max. 

12-4  mm 

90  mm 

0-81 

0-64 

160° 

Min. 

8-6  mm 

4-2  mm 

0-49 

0-54 

120° 

Mean 

10-4  mm 

6-2  mm 

0-68 

0-59 

145-5° 

Median 

10-5  mm 

6-6  mm 

0-65 

0-59 

140° 

Description.  Transversely  subovate  Praenucula  in  which  the  posterior  end  is  slightly  rounded  or  subtruncate 
and  the  front  is  rounded  or  somewhat  angular.  The  height  is  usually  between  0-6  and  0-7  of  the  length.  The 
umbo  is  at  about  the  posterior  two-fifths.  Maximum  inflation  of  a single  valve  2-2  mm  seen  in  a valve  11-6  mm 
long  and  9-0  mm  high.  The  highest  part  of  the  shell  is  a little  anterior  of  the  umbo.  The  hinge-plates  meet 
at  an  angle  of  about  140-145°;  the  anterior  hinge-plate  is  broad  and  arched  ventrally  and  bears  up  to  eleven 
large  teeth,  the  posterior  hinge-plate  bearing  up  to  nine  which  are  smaller  and  less  markedly  chevron-shaped. 
The  dental  series  meet  below  the  umbo.  The  adductor  muscle  scars  are  of  equal  size,  in  diameter  about 
one-eighth  of  the  valve  height;  the  anterior  scar  is  anterior  to  the  end  of  the  anterior  hinge-plate  and  lies  at 
the  same  height  as  the  ventral  edge  of  the  plate.  The  posterior  scar  lies  immediately  below  the  end  of  the 
posterior  hinge-plate.  Both  scars  are  distinct  and  quite  strongly  impressed.  At  least  two  umbonal  muscle  scars 
present.  Sculpture  of  fine  concentric  lines,  c.  16  per  mm. 


TUNNICLIFF:  LATE  ORDOVICIAN  BIVALVES 


53 


Discussion.  The  small  size  of  the  adductor  muscle  scars  separates  Praenucula  infirma  from  other 
Praenucula  species.  In  addition  P.  infirma  is  distinguished  from  P.  praetermissa  by  its  generally 
more  elongate  shape,  and  relatively  shorter  posterior  hinge-plate  with  fewer  teeth,  and  from  P. 
dispersa  by  its  larger,  more  strongly  chevron-shaped  teeth  and  more  elongate  shape.  It  bears  some 
resemblance  in  shape  to  Hind’s  Nuculana  curta  (1910,  p.  521,  pi.  4,  figs.  10-14,  14a,  especially  figs. 
13,  14,  14a),  but  it  lacks  the  central  cartilage  pit  noted  by  Hind  in  that  species.  It  differs  in  the 
same  way  from  Ctenodonta  albertina  Ulrich  and  C.filistriata  Ulrich  (1894,  pp.  598-599,  pi.  42, 
figs.  76-82,  text-fig.  44).  It  lacks  the  posterior  lobe  seen  in  N.  lobata  Hind  (1910,  p.  519,  pi.  4,  figs. 
1-3).  P.  infirma  shows  no  grouping  of  the  growth  lines  as  seen  in  Deceptrix  pulchra  armoricana 
Babin  and  Melou  1972  ( = D . cf.  ciae  (Sharpe)  of  Babin  and  Robardet  1973)  from  the  Caradoc  of 
Normandy.  Like  P.  praetermissa , P.  infirma  differs  from  Praeleda  ciae  (Sharpe)  of  Bradshaw  (1970, 
pp.  633-636)  and  C.  ciae  (Sharpe)  of  Babin  (1966,  pp.  49-52,  pi.  1,  fig.  9)  in  the  posterior  teeth, 
the  chevrons  apparently  pointing  away  from  the  umbo  in  the  C.  ciae  specimens  illustrated,  but 
pointing  towards  it  in  infirma.  This  feature  is,  however,  less  marked  than  in  Praenucula  praetermissa. 
In  this  respect  P.  infirma,  like  P.  praetermissa,  resembles  Sharpe’s  original  N.  ciae  (1853,  p.  149, 
pi.  9,  figs.  5 a-c). 

Genus  deceptrix  Fuchs,  1919 

Type-species.  By  monotypy,  Deceptrix  carinata  Fuchs  1919,  p.  79,  pi.  7,  figs.  1,  2.  See  discussion  under 
Praenuculidae  above. 


Deceptrix  sp. 

Plate  8,  figs.  15,  18 

v pars.  1843  Area  regularis  Portlock,  p.  427,  pi.  34,  fig.  2 [see  Deceptrix  regularise  IGS  GSM  7805], 
Material.  Two  left  valves,  IGS  GU  1896  and  UM  K4202  and  one  right  valve,  IGS  GSM  23215. 

Horizon  and  locality.  Bardahessiagh  Formation;  south  of  Craigbardahessiagh,  Pomeroy,  exact  localities 
uncertain. 

Description.  Shell  subquadrate,  truncate  posteriorly  and  slightly  rounded  to  subtruncate  anteriorly.  The  height 
to  length  ratio  is  between  0-62  and  0-68  and  the  umbo  is  at  about  the  anterior  two-fifths.  Inflation  of  single 
valve  1-5  to  2-0  mm.  The  anterior  and  posterior  hinge-plates  meet  at  an  angle  of  140  to  150°.  4+  anterior 
and  10+  posterior  teeth  are  visible  in  the  specimens  despite  imperfect  preservation.  Anterior  and  posterior 
muscle  scars  indistinct,  lying  close  below  the  ends  of  the  hinge-plates.  Sculpture  of  fine  concentric  lines  of 
variable  strength. 

Discussion.  Like  other  nuculoids,  Deceptrix  is  uncommon  in  the  Bardahessiagh  Formation,  and 
they  can  all  be  described  under  one  heading.  Lacking  complete  information  on  dentition  and 
musculature,  no  specific  name  is  proposed. 

A specimen  in  the  Sedgwick  Museum  (A  16306;  a left  valve)  is  said  to  be  from  the  Bardahessiagh 
Formation  and  may  belong  to  the  same  species,  but  the  anterior  appears  deformed  and  is  very 
slender  compared  with  the  posterior. 

A specimen  collected  by  Fearnsides,  Elies  and  Smith,  SM  A 16295,  also  from  the  Bardahessiagh 
Formation,  was  described  by  Reed  (1952,  p.  65)  as  Ctenodonta  cf.  nitida  (Ulrich).  It  is  a badly 
preserved  left  valve  with  8 + posterior  and  3 + anterior  teeth,  the  dentition  being  obscured  in  part. 
Its  height  is  1 1 mm,  and  length  14  mm,  giving  height  to  length  ratio  0-79.  The  umbo  is  at  about 
the  anterior  third.  The  angle  between  the  anterior  and  posterior  hinge-plates  is  1 1 5°  and  the  inflation 
of  the  single  valve  is  2 mm.  The  musculature  is  unknown.  It  is  unlike  other  Deceptrix  from  the 
Bardahessiagh  Formation  in  having  a more  rounded  outline  and  the  umbo  closer  to  the  anterior. 
As  Reed  (1952,  p.  65)  observed,  it  bears  a strong  resemblance  in  form  to  Ctenodonta  nitida  (Ulrich 
1894,  p.  592,  pi.  42,  figs.  44-49),  but  he  was  clearly  mistaken  in  suggesting  that  this  specimen 
belongs  to  the  same  species  as  McCoy’s  Area  quadrata  (see  Deceptrix  apjohnni ) and  his  comparison 
with  Nuculites  dissimilis  (Portlock)  is  equally  mistaken  (see  N.  cylindricus). 


54 


PALAEONTOLOGY,  VOLUME  25 


Deceptrix  apjohnni  (Portlock  1843) 

Plate  8,  figs.  7,  8,  10,  11,  13,  14,  16,  17;  text-fig.  4 a 

v*  1843  Pectunculus  Apjohnni  Portlock,  p.  429,  pi.  34,  fig.  8. 
v.  1843  Pectunculus!  ambiguus  Portlock,  p.  430,  pi.  34,  fig.  11. 
v.  1846  Pectunculus  Apjohni  Portlock;  McCoy,  p.  19. 
v.  1846  Area  quadrata  McCoy,  p.  20,  pi.  2,  fig.  5 (reversed). 

. 1878  Ctenodonta  ambigua  Portlock;  Baily,  p.  28. 
v.  1952  Ctenodonta  apjohni  (Portlock);  Reed,  p.  60. 

Type-specimens : IGS  GSM  12409,  Portlock’s  (1843,  pi.  34,  fig.  8)  figured  specimen,  is  here  selected  as  lectotype 
of  Pectunculus  apjohnni ; paralectotype  UM  K4219.  The  only  known  syntype  of  PP.  ambiguus  Portlock,  IGS 
GSM  12410,  is  here  selected  as  lectotype  of  P.  ambiguus.  The  figured  specimen  of  Area  quadrata  McCoy  NMI 
G.3.  1979,  which  is  in  the  Griffith  Collection,  and  which  is  the  only  known  syntype,  is  here  selected  as  lectotype 
of  A.  quadrata.  All  are  from  the  Killey  Bridge  Formation,  Pomeroy,  exact  locality  uncertain,  but  possibly 
locality  3 of  Mitchell,  1977. 

Material'.  Sixteen  specimens  in  IGS,  BM,  NMI,  UM. 

Horizon  and  locality.  All  from  the  Killey  Bridge  Formation  (Ashgill,  Cautleyan),  with  well-localized  specimens 
recorded  from  the  Crocknagargen  Stream  section  (locality  1 of  Mitchell  1977)  and  the  Tirnaskea  Stream 
section.  Specimens  from  old  collections  are  less  well  localized. 


Measurements 

L 

H 

H/L 

AL/L 

Angle 

Max. 

14-0  mm 

14-0  mm 

1-17 

0-5 

160° 

Min. 

4-8  mm 

4-4  mm 

0 81 

0-32 

125° 

Mean 

11-78  mm 

10-23  mm 

0-93 

0-36 

145° 

Median 

9-4  mm 

9-2  mm 

0-99 

0-41 

142-5° 

Description.  Rounded  Deceptrix,  occasionally  subquadrate  (as  in  A.  quadrata),  usually  increasing  in  height 
towards  the  front.  The  postero-dorsal  margin  extends  a little  beyond  the  posterior  hinge-line  while  the  anterior 
hinge-plate  occupies  about  half  the  length  of  the  well-rounded  antero-dorsal  margin.  The  umbo  is  situated 
between  the  mid-point  and  the  anterior  third.  Fleight  occasionally  greater  than  length  but  generally  the  height 
to  length  ratio  is  about  0-9-10.  Maximum  inflation  of  a single  valve  4 mm  seen  in  the  lectotype,  14  mm  long 
and  13  mm  high.  The  posterior  hinge-line  is  straight  or  gently  curved,  generally  bearing  up  to  17  teeth 
(exceptionally  26  as  in  A.  quadrata)',  the  anterior  hinge-plate  is  nearly  straight  or  gently  arched  towards  the 
body  of  the  shell  with  10+  teeth.  The  anterior  and  posterior  hinge-plates  meet  below  the  umbo  at  an  angle 
of  140-150°.  The  anterior  adductor  muscle  scar  is  below  the  end  of  the  anterior  dental  series,  at  about  half 
the  height  of  the  shell  and  is  one-quarter  to  one-third  the  height  in  diameter.  The  posterior  adductor  muscle 
scar  is  similar  in  size  to  the  anterior  scar  and  is  close  below  the  posterior  dental  series,  a little  anterior  to  its 
end.  The  scars  of  three  umbonal  accessory  muscles  are  present  (PI.  8,  fig.  17).  Sculpture  is  of  fine  concentric 
lines,  often  of  variable  strength  towards  the  margin.  Pallial  line  not  seen. 

Discussion.  Portlock  (1843,  p.  429)  described  his  Pectunculus  apjohnni  as  a ‘very  well  marked  species’ 
and  its  well-rounded  or  subquadrate  appearance  is  certainly  distinctive  in  the  Killey  Bridge 
Formation  fauna.  It  can  only  be  confused,  if  fragmentary,  with  Deceptrix  semitruncata  but  the 
two  have  distinctive  forms,  D.  apjohnni  more  rounded  and  D.  semitruncata  more  obliquely  elongate 
and  do  not  overlap  in  the  range  of  their  height-length  ratios. 

McCoy’s  A.  quadrata  is  here  regarded  as  synonymous  with  D.  apjohnni.  A.  quadrata  was  described 
as  ‘rare’  by  McCoy  (1846,  p.  20)  and  only  the  type  specimen  is  known  and  although  it  differs  from 
Portlock’s  figured  specimen  of  D.  apjohnni  in  the  manner  outlined  by  McCoy,  in  particular  its 
apparently  square  form  (the  result  of  the  ventral  margin  being  obscured  by  matrix),  it  is  here 
considered  to  be  an  atypical  specimen  of  D.  apjohnni. 

Reed  (1952,  pp.  60-61)  likened  D.  apjohnni  to  several  North  American  species  which  would  now 
be  regarded  as  Similodonta,  and  in  this  he  was  clearly  misled  by  the  general  shape  of  the  shell,  for 
the  dentition  plainly  shows  D.  apjohnni  to  be  Deceptrix.  In  Similodonta  the  anterior  and  posterior 
teeth  are  equal  in  number  and  size.  Of  the  described  species  of  Deceptrix,  none  matches  D.  apjohnni 


TUNNICLIFF:  LATE  ORDOVICIAN  BIVALVES 


55 


a 


text-fig.  4.  Deceptrix  spp.  based  on  internal  moulds,  c.  x 5.  (a)  D.  apjohnni  (Portlock),  IGS  NIL  8910, 
right  valve.  ( b ) D.  semitruncata  (Portlock),  IGS  GSM  12421,  right  valve,  (c)  D.  subtruncata  (Portlock),  IGS 
NIL  8971,  right  valve,  (d)  D.  regularis  (Portlock),  IGS  NIL  8919,  left  valve.  Abbreviations  as  in  text-fig.  3. 
Note  the  differences  in  size  and  positions  of  the  muscle  scars  and  the  differences  in  number  and  distribution 
of  teeth.  Pedal  accessory  muscle  scars  are  shown  but  not  labelled  in  (a),  (c),  and  (d). 

so  closely  in  shape  as  the  type  species  D.  carinata  Fuchs,  1919,  which  has  fewer  anterior  and  more 
posterior  teeth. 


Deceptrix  semitruncata  (Portlock  1843) 

Plate  8,  figs.  9,  12;  text-fig.  4b 

v*  1843  Pectunculus  semi-truncatus  Portlock,  p.  429  (pars),  pi.  34,  fig.  7. 

. 1878  Ctenodonta  semi-truncatus  Portl.  (probably  transversa)',  Baily,  p.  28. 
v?  1910  Ctenodonta  semitruncatus  Portlock;  Hind,  p.  525,  pi.  3,  figs.  23-25. 
v.  1952  Ctenodonta  semitruncata  (Portlock);  Reed,  p.  60,  pi.  3,  fig.  3. 

Type-specimens.  Lectotype  selected  by  Reed  1952,  p.  60,  IGS  GSM  12421,  the  specimen  figured  by  Portlock; 
from  the  Killey  Bridge  Formation,  Pomeroy,  exact  locality  uncertain.  K4203,  a syntype  of  Pectunculus 
semitruncatus  is  now  identified  as  Praenucula  sp.  from  the  Bardahessiagh  Formation. 

Material,  horizon,  and  localities:  Seven  specimens  in  IGS  and  UM,  all  from  the  Killey  Bridge  Formation, 
recorded  from  the  Crocknagargan  Stream  section  (locality  1 of  Mitchell  1977)  and  in  the  Portlock  Collection, 
exact  locality  uncertain. 


56 


PALAEONTOLOGY,  VOLUME  25 


Measurements 

L 

H 

H/L 

AL/L 

Angle 

Max. 

17  0 mm 

13-0  mm 

0-8 

0-52 

160° 

Min. 

110  mm 

9.0  mm 

0-65 

0-36 

140° 

Mean 

15-29  mm 

11-14  mm 

0-73 

0-44 

150° 

Median 

14-0  mm 

1 1 -0  mm 

0-72 

0-44 

150° 

Description.  Obliquely  ovate  Deceptrix,  increasing  in  height  towards  the  anterior,  in  which  the  height  to  length 
ratio  is  between  0-6  and  0-8  and  maximum  inflation  of  a single  valve  seen  is  3 mm.  The  umbo  is  situated 
between  the  midpoint  and  the  anterior  third.  The  posterior  margin  is  truncate  and  the  anterior  margin  is 
produced  obliquely  and  is  sometimes  angular  antero-ventrally  where  it  is  met  by  a slight  umbonal  ridge.  The 
posterior  hinge-plate  is  straight  or  very  gently  curved,  bearing  1 7 + teeth.  The  anterior  hinge  plate  is  straight 
or  slightly  arched  towards  the  body  of  the  shell  with  1 1 + teeth.  The  anterior  and  posterior  dental  series  meet 
beneath  the  umbo.  The  anterior  and  posterior  hinge-plates  meet  at  an  angle  of  about  150°.  Musculature 
indistinct:  anterior  adductor  muscle  scar  is  at  about  half  the  height  of  the  shell,  below  and  a little  before  the  end  of 
the  anterior  dental  series;  posterior  adductor  muscle  scar  is  below  the  hind  third  of  the  posterior  dental  series. 
Sculpture  is  of  fine  concentric  lines  of  variable  strength. 


Discussion.  D.  semitruncata  is  compared  with  D.  apjohnni  under  that  species.  D.  semitruncata  is 
similar  in  shape  to  Ctenodonta  britannica  Babin  (1966,  p.  54,  pi.  1 , figs.  1,  2),  which  may  be  referable 
to  Deceptrix  but  which  has  more  numerous  anterior  teeth  on  a hinge-plate  apparently  arched  away 
from  the  body  of  the  shell.  C.  socialis  Ulrich  (1894,  p.  594,  pi.  37,  figs.  59,  60)  is  also  similar  in 
shape  to  D.  semitruncata  but  is  a very  small  species  with  relatively  fewer  anterior  teeth  (6  out  of 
19,  as  opposed  to  11  out  of  17  in  D.  semitruncata).  No  other  described  species  of  Deceptrix  is 
readily  comparable  with  D.  semitruncata. 


Deceptrix  regularis  (Portlock  1843) 
Plate  8,  figs.  1-6;  text-fig.  4 d 


x*  pars  1843 
. 1878 
v pars.  1910 
v.  1952 


Area  regularis  Portlock,  p.  427,  pi.  34,  fig.  2. 

Ctenodonta  regularis  Portlock;  Baily,  p.  28. 

Ctenodonta  regularis  Portlock;  Hind,  p.  524,  ?pl.  3,  figs.  15-17. 
Ctenodonta  regularis  (Portlock);  Reed,  p.  61,  pi.  3,  fig.  4. 


EXPLANATION  OF  PLATE  8 

Figs.  1-6.  Deceptrix  regularis  (Portlock  1843).  1,  4,  oblique  dorsal  view  and  lateral  view,  internal  mould  of 

left  valve,  IGS  GSM  12419,  Pomeroy,  exact  locality  uncertain.  2,  5,  lateral  view  and  latex  cast,  internal  mould 
of  right  valve,  IGS  NIL  8803,  Crocknagargan  Stream  section,  Pomeroy  (IGR  C.H721737).  3,  6,  lateral 
view  and  latex  cast,  internal  mould  of  left  valve,  IGS  NIL  8919,  Crocknagargan  Stream  section,  Pomeroy 
(IGR  C.H721737).  All  Killey  Bridge  Formation  (Ashgill,  Cautleyan),  x 3. 

Figs.  7,  8,  10,  11,  13,  14,  16,  17.  Deceptrix  apjohnni  (Portlock  1843).  7,  latex  cast  of  external  mould  of  right 

valve,  IGS  GSM  12410,  lectotype  of  Pectunculus  ambiguus  Portlock,  1 843,  Pomeroy,  exact  locality  uncertain, 
x 3.  8,  internal  mould  of  left  valve,  NMI  G.3.  1979,  lectotype  of  Area  quadrata  McCoy,  1846,  Pomeroy, 
exact  locality  uncertain,  x 2.  10,  internal  mould  of  left  valve,  lectotype,  IGS  GSM  12409,  Pomeroy,  exact 

locality  uncertain,  x3.  11,  incomplete  internal  mould  of  left  valve,  UM  K4253,  Pomeroy,  exact  locality 

uncertain,  x 3.  13,  16,  latex  cast  of  hinge  and  lateral  view,  internal  mould  of  left  valve  BM  LL40005,  locality 

3 of  Mitchell,  1977  (IGR  H72977268),  x 3.  14,  17,  latex  cast  of  hinge  and  lateral  view,  internal  mould  of 

right  valve,  IGS  NIL  8910,  Crocknagargan  Stream  section,  Pomeroy  (IGR  C.H721737),  x 3.  All  Killey  Bridge 
Formation  (Ashgill,  Cautleyan). 

Figs.  9,  12.  Deceptrix  semitruncata  (Portlock  1843),  oblique  dorsal  view  and  lateral  view,  internal  mould  of 
right  valve,  lectotype,  IGS  GSM  12421,  Killey  Bridge  Formation  (Ashgill,  Cautleyan),  Pomeroy,  exact 
locality  uncertain,  x 3. 

Figs.  15,  18.  Deceptrix  sp.  15,  composite  mould  of  right  valve,  IGS  GSM  23215.  18,  composite  mould  of 

left  valve,  UM  K4202.  Both  Bardahessiagh  Formation  (Caradoc),  south  of  Craigbardahessiagh,  Pomeroy, 
x 3. 


PLATE  8 


TUNNICLIFF,  Ordovician  bivalves 


58 


PALAEONTOLOGY,  VOLUME  25 


Type-specimens.  Lectotype  (as  ‘holotype’)  inadvertently  designated  by  Hind  (1910,  p.  524),  IGS  GSM  7805, 
probably  the  specimen  figured  by  Portlock  (1843,  pi.  34,  fig.  2);  remaining  syntypes  IGS  GSM  12419,  23215, 
UM  K4202.  GSM  23215  and  K4202  are  now  identified  as  coming  from  the  Bardahessiagh  Formation,  exact 
locality  uncertain,  and  are  referred  to  Deceptrix  sp.  The  lectotype  and  remaining  syntype  are  from  the  Killey 
Bridge  Formation,  Pomeroy,  exact  locality  uncertain  but  possibly  locality  3 of  Mitchell  1977. 

Material,  horizon,  and  localities.  Two  specimens  from  the  Portlock  Collection,  IGS  GSM  7805,  12419  and 
two  from  recent  collections,  IGS  NIL  8803,  8919.  All  are  from  the  Killey  Bridge  Formation;  the  locality  of 
the  Portlock  specimens  is  uncertain  but  the  remaining  specimens  are  from  the  Crocknagargan  Stream  section 
(locality  1 of  Mitchell  1977). 


Measurements 

L 

H 

H/L 

AL/L 

Angle 

Max. 

17-5  mm 

13  0 mm 

0-75 

0-46 

150° 

Min. 

10-5  mm 

7.0  mm 

0-64 

0-36 

145° 

Mean 

9-5  mm 

13-5  mm 

0-70 

0-42 

148-75' 

Median 

140  mm 

10  0 mm 

0-70 

0-41 

147-5° 

Description.  Transversely  ovate  Deceptrix,  rounded  at  both  anterior  and  posterior  ends  increasing  in  height 
slightly  towards  the  rear.  Height  to  length  ratio  about  0-7.  The  umbo  is  at  about  the  anterior  two-fifths.  The 
maximum  inflation  of  a single  valve  is  3 mm.  The  anterior  and  posterior  hinge-plates  meet  below  the  umbo 
at  an  angle  of  150°  with  a continuous  series  of  nearly  twice  as  many  smaller  posterior  teeth  than  anterior 
teeth,  up  to  15  anterior  and  25  posterior  in  a large  specimen  (PI.  8,  figs.  2,  5).  The  posterior  hinge-plate  is 
gently  curved,  the  anterior  is  slightly  arched  towards  the  body  of  the  shell.  Adductor  muscle  scars  vary  from 
quite  strongly  impressed  (PI.  8,  figs.  3,  6)  to  indistinct.  Anterior  scar  lies  below  and  a little  in  front  of  the 
end  of  the  anterior  teeth;  the  posterior  scar  lies  a little  anterior  to  the  end  of  the  posterior  teeth.  Three 
umbonal  accessory  muscle  scars  are  present  (PI.  8,  figs.  3,  6).  Sculpture  is  not  well  seen  in  the  specimens  but 
appears  to  be  of  fine  concentric  lines. 

Discussion.  The  lectotype  of  Deceptrix  regularis,  IGS  GSM  7805,  shows  little  other  than  the  shape 
and  is  not  illustrated  here.  It  is  perhaps  surprising  that  Hind  (1910,  p.  524)  regarded  this  as  the 
type  since  GSM  12419  shows  dentition,  and  there  is  no  clear  evidence  which  was  the  specimen 
figured  by  Portlock  [see  Tunnicliff  1980]. 


EXPLANATION  OF  PLATE  9 

Figs.  1-7,  9-11.  Deceptrix  subtruncata  (Portlock  1843).  1-6,  IGS  NIL  8971,  slab  with  internal  moulds  of 

both  valves,  Crocknagargan  Stream  section,  Pomeroy,  (IGR  C.H721737):  1,  2,  latex  cast  and  lateral  view 
of  right  valve;  5,  6,  latex  cast  and  lateral  view  of  left  valve  (in  fig.  6,  difficulty  in  lighting  the  subject  has 
resulted  in  a shortened  appearance  of  the  anterior);  3,  4,  dorsal  views  of  left  and  right  valves  juxtaposed. 
7,  internal  mould  left  valve,  IGS  GSM  12413,  holotype  of  Ctenodonta  deserta  Reed  1952,  Pomeroy,  exact 
locality  uncertain.  9,  oblique  dorsal  view  and  lateral  view,  damaged  internal  mould  of  left  valve, 
IGS  GSM  1 2423,  lectotype  of  Area  transversa  Portlock  1 843,  Pomeroy,  exact  locality  uncertain.  1 0,  internal 
mould  right  valve,  IGS  GSM  12422,  Pomeroy,  exact  locality  uncertain.  11,  internal  mould  right  valve, 
lectotype,  IGS  GSM  12424,  Pomeroy,  exact  locality  uncertain.  All  Killey  Bridge  Formation  (Ashgill, 
Cautleyan),  x 3. 

Fig.  8.  Hippocardia  praepristis  (Reed  1952),  holotype,  IGS  GSM  24147,  Killey  Bridge  Formation  (Ashgill, 
Cautleyan),  Pomeroy,  exact  locality  uncertain,  x 3. 

Figs.  12-20.  Concavodonta  imbricata  (Portlock,  1843).  12,  internal  mould  of  right  valve,  IGS  Zs  2790. 

13-15,  composite  mould  of  right  valve,  IGS  GU  1583;  15,  dorsal  view;  14,  latex  cast;  13,  lateral  view. 
1 6,  ?external  cast  of  left  valve,  NMI  G.  1 . 1 979,  lectotype  of  Nucula  subacuta  McCoy,  1 846.  1 7,  ?composite 

mould  of  left  valve,  NMI  G.2.  1979,  cited  by  McCoy  1846  as  Nucula  protei  Munster.  18-20,  composite 
mould  of  left  valve,  BM  LL40003;  18,  lateral  view;  19,  dorsal  view;  20,  latex  cast.  All  Killey  Bridge  Formation 
(Ashgill,  Cautleyan),  figs.  12-15,  18-20,  from  locality  3 of  Mitchell  1977  (IGR  H7297  7268),  figs.  16,  17, 
Pomeroy,  exact  locality  uncertain.  All  x4. 


PLATE  9 


TUNNICLIFF,  Ordovician  bivalves 


60 


PALAEONTOLOGY,  VOLUME  25 


Although  in  position  of  umbo  and  in  height-length  ratio  D.  regularis  closely  matches  the  range 
of  D.  semitruncata,  it  can  be  readily  distinguished  from  that  species  by  its  transverse  rather  than 
oblique  elongation  and  its  increase  in  height  towards  the  posterior  end,  while  D.  semitruncata  is 
highest  towards  the  anterior  end.  It  is  less  easy  to  distinguish  between  D.  regularis  and  D.  subtruncata, 
but  the  latter  is  always  more  transverse  with  the  highest  part  of  the  shell  at  the  umbo  rather  than 
towards  the  posterior  end,  and  has  relatively  larger  adductor  muscle  scars  and  fewer  teeth. 

Pojeta’s  illustrations  of  his  Deceptrix  ( D .)  n.  sp.  1 (1978,  pi.  1 figs.  1,  2)  show  a close  similarity 
to  D.  regularis  except  that  the  umbo  is  closer  to  the  mid-point  and  the  shell  is  highest  a little 
anterior  of  the  umbo  and  thus  more  closely  resembles  D.  subtruncata.  No  other  species  of  Deceptrix 
is  closely  comparable  with  D.  regularis. 


v*  1843 
vpars.  1843 
v.  1843 
v. 1846 
v. 1846 
. 1878 
v non.  1910 
1910 
71946 
v.  1952 
v.  1952 
v.  1952 


Deceptrix  subtruncata  (Portlock  1843) 

Plate  9,  figs.  1-7,  9-11;  text-fig.  4c 
Area  sub-truncata  Portlock,  p.  427,  pi.  34,  fig.  1. 

Area  transversa  Portlock,  p.  428,  pi.  34,  fig.  4 [non  IGS  GSM  12425  Nuculites  sp.] 
Area  Eastnori?  (Murchison);  Portlock,  p.  427,  pi.  34,  fig.  3. 

Area  subtruncata  Portlock;  McCoy,  p.  20. 

Area  transversa  Portlock;  McCoy,  p.  20. 

Ctenodonta  transversa  Portl.;  Baily,  p.  28. 

Ctenodonta  aff.  transversa  Portlock;  Hind,  p.  523,  pi.  3,  figs.  12-14. 

Ctenodonta  eastnori  Portlock;  Hind,  p.  525,  pi.  3,  fig.  20,  ?figs.  21,  22. 

Ctenodonta  transversa  (Portlock);  Reed,  p.  202. 

Ctenodonta  subtruncata  (Portlock);  Reed,  p.  62,  pi.  3,  fig.  6. 

Ctenodonta  transversa  (Portlock);  Reed,  p.  63. 

Ctenodonta  deserta  Reed,  p.  64,  pi.  3,  fig.  8. 


Type-specimens.  IGS  GSM  12424,  the  specimen  figured  by  Portlock  (1843,  pi.  34,  fig.  1)  is  here  selected  as 
lectotype  of  Area  subtruncata ; paralectotypes  IGS  GSM  12422,  TCD  14763.  IGS  GSM  12423,  the  specimen 
figured  by  Portlock  (1843,  pi.  34,  fig.  4)  was  selected  as  lectotype  of  A.  transversa  by  Reed  1952,  p.  68; 
remaining  syntypes  IGS  GSM  12425,  TCD  7854.  The  holotype  of  Ctenodonta  deserta  Reed  is  IGS  GSM 
12413  by  original  designation  (1952,  p.  64).  All  from  the  Killey  Bridge  Formation,  Pomeroy,  exact  localities 
uncertain  but  possibly  locality  3 of  Mitchell  1977. 


Material,  horizon,  and  localities.  Many  specimens  in  IGS,  UM,  SM,  TCD,  in  both  old  and  recent  collections, 
are  all  from  the  Killey  Bridge  Formation.  Recent  collections  contain  specimens  recorded  from  localities  1 
and  3 of  Mitchell  1977,  while  old  collection  material  is  less  well  localized. 


Measurements 

L 

H 

H/L 

AL/L 

Angle 

Max. 

19-4  mm 

11-4  mm 

0-64 

0-51 

170° 

Min. 

10  0 mm 

5.4  mm 

0-54 

0-37 

140° 

Mean 

17-3  mm 

1015  mm 

0-59 

0-48 

153-8° 

Median 

16-76  mm 

10-26  mm 

0-59 

0-44 

153-6° 

Description.  Transversely  elongate,  subelliptical  Deceptrix,  with  obliquely  subtruncate  anterior  and  rounded 
posterior  end.  Maximum  height  of  shell  below  umbo.  Height  to  length  ratio  about  0-6.  Umbo  nearly  central. 
Maximum  inflation  of  a single  valve  is  5-0  mm  in  a valve  17-5  mm  long  and  10-0  mm  high.  Hinge-line  straight 
or  slightly  curved;  anterior  and  posterior  hinge-plates  meet  at  an  angle  of  about  155-160°  in  most  specimens. 
The  posterior  dental  series  is  straight  with  up  to  twenty  teeth  seen  in  both  right  and  left  valves.  The  anterior 
dental  series  is  slightly  arched  towards  the  body  of  the  shell  with  up  to  twelve  teeth  in  the  right  valve  and 
thirteen  in  the  left  (a  single,  tiny  subumbonal  tooth  in  the  left  valve  is  here  counted  as  anterior).  The  dental 
series  meet  below  the  umbo.  Adductor  muscle  scars  are  large,  nearly  half  the  height  of  the  shell  in  diameter, 
and  distinct,  each  lying  slightly  in  front  of,  and  below  the  end  of,  its  corresponding  dental  series.  Four 
umbonal  accessory  muscle  scars  are  present  in  each  valve  (PI.  9,  figs.  1-6),  three  distinct,  one  less  so.  Sculpture 
of  fine  concentric  lines  of  variable  strength  especially  towards  the  margin.  Pallial  line  not  seen. 


TUNNICLIFF:  LATE  ORDOVICIAN  BIVALVES 


61 


Discussion.  There  is  no  basis  on  which  A.  transversa  Portlock  (IGS  GSM  12423)  or  the  specimen 
which  he  labelled  and  described  as  A.  Eastnori  can  be  separated  from  Deceptrix  subtruncata 
(Portlock).  Despite  Reed’s  observations  (1952,  p.  63)  to  the  contrary,  and  Portlock’s  original 
description  (1843,  p.  428),  the  proportions  of  the  shell  in  A.  transversa  are  not  greatly  different 
from  those  of  D.  subtruncata  although  the  type  specimen  of  A.  transversa  is  incomplete  and  its 
length  must  be  estimated.  Portlock’s  figures  of  the  three  species  (1843,  pi.  34,  figs.  1,  3,  4)  are  poor 
and  measurements  taken  from  them  are  unreliable. 

Reed  (1952,  p.  64)  quite  reasonably  dissociated  Portlock’s  specimen  of  A.  eastnori  (GSM  12413) 
from  Sowerby’s  species,  but  he  failed  to  recognize  its  affinities  with  D.  subtruncata  and  created  a 
new  species,  C.  deserta.  The  second  Portlock  specimen  of  A.  eastnori  (GSM  12414)  which  Reed 
compared  to  C.  cingulata  Ulrich  is  Nuculites  sp. 

The  differences  between  D.  subtruncata  and  D.  regu/aris,  with  which  it  might  be  confused,  are 
discussed  under  the  latter  species.  Of  the  published  species  of  Deceptrix  none  compares  so  closely 
with  D.  subtruncata  as  that  figured  by  Pojeta  (1978,  pi.  1 figs.  1,  2)  as  D.  (/).)  n.  sp.  1 which  has 
similar  numbers  of  teeth,  position  of  umbo,  and,  in  the  case  of  Pojeta’s  figure  2,  a closely  similar 
height  to  length  ratio  to  the  maximum  seen  in  D.  subtruncata. 


Genus  similodonta  Soot-Ryen,  1964 

Type-species.  Designated  by  Soot-Ryen  1964,  p.  498,  Tellinomya  similis  Ulrich,  1892. 


Similodonta?  sp. 

Plate  7,  fig.  25 

Material.  Left  valve,  UM  K4205  and  a slab,  UM  K4241,  bearing  a left  and  a right  valve;  all  composite  moulds. 
Horizon  and  locality.  Bardahessiagh  Formation,  locality  uncertain  but  south  of  Craigbardahessiagh,  Pomeroy. 
Measurements  H L H/L  AL/L  Angle 

K4241  (both  valves)  15  0 mm  16  0 mm  0-94  0-5  c.  110° 

K4205  15-6  mm  14  0 mm  Ml  0-5  95-100° 

Description.  Triangular,  equivalve,  about  as  high  as  long  with  the  umbo  at  the  centre.  Anterior  and  posterior 
hinge-plates  meeting  at  about  100°.  Details  of  hinge  unknown.  Anterior  adductor  muscle  scar  distinct,  lying 
below  anterior  end  of  the  anterior  hinge-plate  which  is  arched  towards  the  body  of  the  shell.  Posterior  adductor 
muscle  scar  not  known.  Auxiliary  musculature  unknown.  Sculpture  of  fine  concentric  striae  (about  7 per 
mm)  with  coarser  striae  developing  at  later  growth  stages. 

Discussion.  Reed’s  description  (1952,  p.  63)  of  Ctenodonta  perangulata  suggests  that  C.  perangulata 
should  be  placed  in  Similodonta,  but  examination  of  his  holotype  (SM  A 16454)  has  shown  it  to 
be  a Nuculites  and  it  is  here  considered  to  be  synonymous  with  Nuculites  cylindricus  (Portlock). 


Genus  concavodonta  Babin  and  Melou,  1972 

Type-species.  Originally  designated  by  Babin  and  Melou  1972,  p.  83,  Nucula  ponderata  Barrande,  1881,  from 
the  Vinice,  Zahorany,  and  Bohdalec  Formations  (Caradoc)  of  Bohemia  (Havlicek  and  Vanek  1966),  recorded 
again  from  Bohemia  by  Pfab  1934,  and  also  recorded  from  the  Upper  Ordovician  of  Normandy  (Babin  and 
Robardet  1973),  and,  with  doubt,  the  Caradoc  of  Brittany  (Babin  and  Melou  1972). 

Diagnosis  after  Babin  and  Melou  (1972,  p.  83).  Genus  of  Praenuculidae  of  fairly  rounded  shape, 
the  posterior  extremity  being  a little  more  slender  than  the  anterior.  Umbo  prosogyrate.  Imprints 
of  adductor  muscles  subequal,  the  anterior  more  strongly  impressed  in  the  shell  at  its  posterior 
edge.  Pallial  line  complete.  Teeth  concavodont,  that  is,  characterized  by  teeth  in  chevron  with  the 
concavities  towards  the  umbo.  Ornamentation  fairly  strong  and  regular,  concentric. 


62 


PALAEONTOLOGY,  VOLUME  25 


Concavodonta  imbricata  (Portlock  1 843) 


Plate  9,  figs.  12-20;  text-fig.  5 


* 1843 
v.  1846 
v. 1846 
v non.  1910 

v.  1952 
v. 1952 


Nucula  acutal  (Sowerby)  var.  imbricata  Portlock,  p.  430,  pi.  34,  fig.  10. 

Nucula  protei  Munster;  McCoy,  p.  19. 

Nucula  subacuta  McCoy,  p.  19,  pi.  2,  fig.  3 (reversed). 

Nuculana  imbricata  Portlock;  Hind,  p.  519,  pi.  4,  figs.  4-7,  7a  [none  of  these  specimens  is 
referable  to  Concavodonta  but  are  referable  to  IPraenucula.  Name  imbricata  re-established]. 
Nuculana ? imbricata  (Portlock);  Reed,  p.  66  [affirms  the  use  of  imbricata ]. 

Ctenodonta  cf.  gibberula  (Salter);  Reed,  p.  65. 


Type  specimens.  Portlock’s  type  specimen  remains  untraced  (Tunnicliff  1980).  A lectotype  for  Nucula  subacuta 
McCoy,  1846,  is  here  designated  NMI  G.l.  1979,  the  only  known  syntype. 


text-fig.  5.  Concavodonta  imbricata  (Portlock),  dentition  and  mus- 
culature based  on  composite  mould,  left  valve,  BM  LL40003,  c.  x 10. 
Compare  with  PI.  9,  fig.  20.  Abbreviations  as  in  text-fig.  3. 


Material.  Thirty-six  specimens  in  IGS,  UM,  BM,  SM,  mostly  internal  moulds  of  single  valves,  some  with 
counterparts,  many  apparently  composite  moulds  showing  the  concentric  ornament  on  the  internal  mould. 

Horizon  and  localities.  All  specimens  are  from  the  Killey  Bridge  Formation,  of  Pomeroy,  Co.  Tyrone.  Most 
specimens  are  from  locality  3 of  Mitchell  1977.  Others  are  from  localities  1 and  2 of  Mitchell  and  from  the  ji 
Tirnaskea  Stream  section.  Specimens  from  old  collections  are  less  precisely  localized. 


Measurements 

L 

H 

H/L 

AL/L 

Angle 

Max. 

9-4  mm 

6-6  mm 

0-923 

0-455 

155° 

Min. 

3-8  mm 

2-4  mm 

0-517 

0-250 

100° 

Mean 

6-6  mm 

4-5  mm 

0-685 

0-3367 

131-1° 

Median 

6-6  mm 

4-5  mm 

0-720 

0-3525 

127-5° 

Description.  Ovate,  equivalve  nuculoid.  Height  generally  about  0-7  of  the  length  but  variable.  Prosogyrate 
umbo  at  about  the  anterior  third  and  extending  a little  above  the  hinge-line.  Convexity  greatest  below  umbo, 
maximum  inflation  of  a single  valve  1-6  mm  seen  in  a valve  9-4  mm  long  and  5-0  mm  high.  Angle  between 
anterior  and  posterior  hinge-plates  about  130°.  Posterior  hinge-plate  straight.  Teeth  concavodont  (see  generic 
diagnosis),  anterior  larger  than  posterior,  apparently  passing  beneath  umbo  in  a continuous  series.  6 + anterior 
teeth  in  both  valves,  19+  posterior  teeth  seen  in  right  valve  and  16+  in  left  valve  (29  posterior  teeth  were 
seen  in  one  large  left  valve  (PI.  9,  fig.  20)).  Anterior  and  posterior  adductor  muscle  scars  clearly  distinguished; 
anterior  slightly  larger  and  more  deeply  impressed  than  posterior  and  most  deeply  impressed  on  its  posterior 


TUNNICLIFF:  LATE  ORDOVICIAN  BIVALVES 


63 


edge.  Accessory  muscles  unknown.  Pallial  line  probably  represented  by  a change  in  convexity  close  to  the 
ventral  margin.  Ligament  and  shell  material  unknown.  Two  forms  of  concentric  sculpture  observed;  coarse 
(3  to  6 striae  per  mm)  giving  the  imbricate  appearance  (PI.  9,  fig.  13),  and  fine  (12  to  20  striae  per  mm)  (PI. 
9,  fig.  12). 

Discussion.  This  is  the  first  record  of  Concavodonta  from  the  British  Isles  but  material  under  study 
from  Caradoc  and  Ashgill  rocks  in  north  Wales  includes  examples  of  Concavodonta  sp.  Specimens 
not  showing  the  dentition  may  be  distinguished  from  Deceptrix  and  Praenucula  by  the  relatively 
coarse  concentric  sculpture,  the  position  of  the  umbo  and  especially  the  straight  posterior  hinge-line. 
Ulrich’s  figures  (1894,  p.  589,  pi.  38,  figs.  25-28;  pi.  42,  figs.  38-40)  of  Ctenodonta  planodorsata 
(Ulrich),  from  the  Trenton  Shale,  show  similar  straight  sets  of  concavodont  teeth,  but  the  angle 
between  the  anterior  and  posterior  hinge-plates  is  more  acute  (105°)  and  the  position  of  the  posterior 
adductor  muscle  scar  differs,  being  in  line  with  the  posterior  hinge-plate  in  C.  planodorsata  but 
lying  internally  (i.e.  antero-ventrally)  to  the  teeth  in  Concavodonta  imbricata.  C.  ponderata,  as 
described  and  figured  by  Babin  and  Melou  (1972),  Babin  and  Robardet  (1973),  and  by  Pfab  (1934), 
differs  from  C.  imbricata  only  in  that  the  posterior  hinge-plate  appears  gently  curved. 


Superfamily  nuculanacea  H.  Adams  and  A.  Adams,  1858 
Family  malletiidae  H.  Adams  and  A.  Adams,  1858 
Genus  nuculites  Conrad,  1841 

Type-species.  Designated  by  Miller  1889,  p.  496,  Nuculites  oblongatus  Conrad,  1841,  from  the  probable  Middle 
Devonian  (McAlester  1968,  p.  37)  of  New  York. 


Nuculites  cylindricus  (Portlock  1843) 


Plate  10,  figs.  1-17 


71841 
v*  1843 
v.  1843 
v. 1843 
v pars.  1 843 
v.  1846 
v.  1846 
v.  1846 
1875 
1878 
1878 
v non.  1910 
v non.  1910 
v non.  1910 
? 1946 
v.  1952 
v.  1952 
v.  1952 
v.  1952 
v.  1952 
v pars.  1952 


Nuculites  planulatus  Conrad,  p.  50  [see  Bretsky  and  Bretsky  1977,  for  synonymy]. 

Area  cylindrica  Portlock,  pp.  428,  759,  pi.  34,  fig.  9. 

Area  dissimilis  Portlock,  pp.  428,  759,  pi.  34,  fig.  5. 

Area  obliqua  Portlock,  pp.  429,  759,  pi.  34,  fig.  6. 

Area  Eastnori  (Murchison);  Portlock,  pp.  427,  759,  non  pi.  34,  fig.  3. 

Area  cylindrica  Portlock;  McCoy,  p.  19. 

Area  dissimilis  Portlock;  McCoy,  p.  20. 

Area  obliqua  Portlock;  McCoy,  p.  20. 

Ctenodonta  obliqua  Portlock;  Baily,  p.  35,  pi.  12,  fig.  2. 

Ctenodonta  dissimilis  Portlock;  Baily,  p.  28. 

Ctenodonta  obliqua  Portlock;  Baily,  p.  28. 

Ctenodonta  dissimilis  Portlock;  Hind,  p.  522,  pi.  3,  figs.  5-7. 

Ctenodonta  eastnori  Portlock;  Hind,  p.  525,  pi.  3,  figs.  20-22. 

Ctenodonta  obliqua  Portlock;  Hind,  p.  524,  pi.  3,  figs.  18,  19. 

Clidophorus  diu  Lamont,  p.  365,  pi.  1,  fig.  3. 

Clidophorus  cylindricus  (Portlock);  Reed,  p.  66,  pi.  3,  fig.  9. 

Ctenodonta  dissimilis  (Portlock);  Reed,  p.  63. 

Ctenodonta  obliqua  (Portlock);  Reed,  p.  61,  pi.  3,  fig.  5. 

Clidophorus  occultus  Reed,  p.  67,  pi.  3,  fig.  10. 

Ctenodonta  perangulata  Reed,  p.  63,  pi.  3,  fig.  7. 

Ctenodonta  deserta  Reed,  p.  65,  non  pi.  3,  fig.  8 [IGS  GSM  12414,  non  holotype  IGS  GSM 
12413], 


Type-specimens.  Lectotype  of  Nuculites  cylindricus  (Portlock)  here  selected,  IGS  GSM  12416,  the  specimen 
figured  and  labelled  by  Portlock  and  described  by  Reed  (1952,  p.  66);  paralectotypes  are  TCD  7876, 
14761-14762,  UM  K4220,  4267.  Lectotype  of  Area  dissimilis  Portlock,  selected  Reed  (1952,  p.  63),  IGS  GSM 
12411;  paralectotype  is  IGS  GSM  12412.  For  A.  obliqua  Portlock,  Hind,  (1910,  p.  524)  selected  IGS  GSM 
12415  as  lectotype,  leaving  paralectotypes  IGS  GSM  12417,  12418.  The  holotype  of  Clidophorus  occultus 


64 


PALAEONTOLOGY,  VOLUME  25 


Reed  is  IGS  GSM  12415  by  original  designation  ( occultus  becomes  a junior  objective  synonym  of  obliqua ). 
The  holotype  of  Ctenodonta  perangulata  Reed  is  SM  A 16454  by  original  designation.  The  holotype  of 
Clidophorus  diu  Lamont  is  Lamont  Collection  No.  1 . 

Material,  localities,  and  horizon.  Many  specimens  in  IGS,  TCD,  UM,  NMI,  BM(NH),  and  SM,  both  in  old 
and  new  collections.  All  are  from  the  Killey  Bridge  Formation  of  Pomeroy.  Apart  from  recently  collected 


material  especially  from  localities  1 

l and  3 of  Mitchell  (1977), 

specimens  are  generally  poorly  localized. 

Measurements 

L 

H 

H/L 

AL/L 

Max. 

21-0  mm 

16-0  mm 

0-94 

0-46 

Min. 

8-4  mm 

3-8  mm 

0-43 

0 15 

Mean 

15-23  mm 

91  mm 

0 61 

0-33 

Median 

14-7  mm 

9-9  mm 

0-68 

0-30 

Description.  Nuculites  of  very  variable  shape,  from  posteriorly  elongate  transversely  or  obliquely,  or  rounded, 
to  more  or  less  truncate  posteriorly.  Umbo  between  the  mid-point  and  the  anterior  one-fifth.  The  height  to 
length  ratio  varies  between  about  0-4  and  0-95.  Many  specimens  show  a posterior  fold  running  from  the 
umbo  towards  the  postero-ventral  margin  (PI.  10,  fig.  5).  Occasionally  a weak  anterior  fold  is  developed 
running  from  the  umbo  towards  the  antero-ventral  margin  (PI.  10,  fig.  12),  anterior  to  the  position  of  the 
septum.  Septal  position  indicated  clearly  by  a strong  impression  on  internal  moulds  and  often  evident  on 
external  moulds.  The  septum  extends,  usually  vertically,  from  the  hinge-line  to  half-way  to  the  margin  (e.g. 
PI.  10,  figs.  2,  9)  but  may  appear  to  be  directed  slightly  posteriorly  (e.g.  PI.  10,  figs.  1,  5)  or  more  markedly 
towards  the  anterior  end  (e.g.  PI.  10,  fig.  17).  Posterior  hinge-plate  straight  or  slightly  curved  near  the  hind 
end  (PI.  10,  fig.  9):  anterior  hinge-plate  generally  at  an  angle  of  about  150°  to  the  posterior  hinge-plate,  but 
the  angle  is  variable  between  120  and  180°.  Dentition  taxodont:  3+  anterior  teeth  and  15+  posterior  teeth. 
The  anterior  teeth  are  simple  or  slightly  sigmoidal  and  are  larger  than  the  posterior  teeth  which  are 
chevron-shaped  pointing  towards  the  umbo  and  becoming  simpler  towards  the  posterior  end.  Dentition  below 
umbo  obscured  in  all  specimens.  Maximum  inflation  of  a single  valve  seen  3-5  mm  in  a valve  15-4  mm  long. 
Anterior  adductor  muscle  scar  distinct  and  occupying  much  of  the  portion  of  the  shell  anterior  to  the  septum: 
posterior  adductor  and  other  muscle  scars  unknown.  Sculpture  of  fine  ( c . 16-20  per  mm)  concentric  lines 
with  some  coarser  lines  interspersed,  especially  towards  the  margin  in  larger  specimens. 

Discussion.  Bretsky  and  Bretsky  (1977)  have  shown  that  N.  planulatus  Conrad,  1841,  from  the 
Upper  Ordovician  of  Quebec  is  a highly  variable  species;  Watkins  (1978,  p.  44)  has  done  the  same 
with  the  Silurian  species  N.  antiquus  (J.  de  C.  Sowerby).  This  also  appears  to  be  the  case  with  the 
Killey  Bridge  Formation  species.  As  isolated  specimens,  the  types  of  Portlock’s  Area  cylindrica,  A. 
dissimilis,  and  A.  obliqua  appear  to  be  different  species  but,  seen  in  combination  with  the  many 
other  specimens  of  Nuculites  both  in  old  and  recent  collections,  they  appear  to  be  extreme  variants 


I 


EXPLANATION  OF  PLATE  10 


Figs.  1-17.  Nuculites  cylindricus  (Portlock  1843).  All  preserved  as  internal  moulds.  All  Killey  Bridge 
Formation  (Ashgill,  Cautleyan).  1,  right  valve,  lectotype,  IGS  GSM  12416.  2,  left  valve,  UM  K4266. 
3,  4,  latex  cast  and  lateral  view,  left  valve,  UM  1920-834.  5,  right  valve  UM  K4267.  6,  left  valve, 

UM  K4226.  8,  right  valve,  IGS  GSM  12412.  9,  left  valve,  IGS  GSM  12418.  10,  left  valve,  UM  K4263. 

1 1 , left  and  right  valves,  probably  of  the  same  individual,  UM  K4264,  lit  from  bottom  right  to  accommodate 
both  valves.  12,  nearly  conjoined  valves,  UM  K4233.  14,  right  valve,  IGS  GSM  12415,  lectotype  of 

Area  obliqua  Portlock  1843.  15,  conjoined  valves,  IGS  GSM  1241 1,  lectotype  of  Area  dissimilis  (Portlock 

1843).  16,  right  valve,  UM  K4255.  17,  left  valve,  UM  K4231.  All  Pomeroy,  exact  locality  uncertain, 

x2i  7,  right  valve,  IGS  GU  2087,  13,  left  valve,  IGS  GU  2061.  Both  locality  3 of  Mitchell  1977  (IGR 
H7297  7268),  x2$. 

Figs.  18,  19.  Nuculoid  gen.  et  sp.  nov.,  IGS  NIL  9240-1.  18,  both  valves,  showing  the  fracturing  of  both 

and  the  greater  displacement  of  parts  of  the  left  valve,  x2.  19,  concave  latex  cast  of  composite  mould, 

right  valve  showing  the  fine  dentition,  concentric  ornament,  and  shallow  sulcus,  x 3.  Tirnaskea  Formation 
(Ashgill,  Hirnantian),  Crocknagargan  Stream  section,  Pomeroy  (IGR  H722737). 


PLATE  10 


TUNNICLIFF,  Ordovician  bivalves 


66 


PALAEONTOLOGY,  VOLUME  25 


of  one  form.  There  is  no  clear  relationship  between  growth  and  height-length  ratio  that  can  be 
related  to  Portlock’s  species,  and  although  many  specimens  may,  by  eye,  be  placed  in  one  or  other 
of  Portlock’s  nominal  species,  there  are  others  which  fall  between  any  two  of  the  Portlock  forms. 
I am  unable  to  quantify  any  differences  between  these  ‘species’  and  choose  therefore  to  place  them 
in  synonymy.  There  may  be  an  argument  for  retaining  Portlock’s  obliqua  and  dissimilis  as  terms 
for  morphs. 

Reed’s  (1952)  Clidophorus  occultus  is  a junior  objective  synonym  of  Portlock’s  A.  obliqua  and 
his  Ctenodonta  perangulata  falls  within  the  range  of  the  A.  dissimilis  form.  Clidophorus  diu  Lamont 
(1946),  from  the  Lower  Drummuck  Group  of  Girvan,  Ayrshire,  appears  from  the  figure  and  from 
Lamont’s  observations  to  fall  within  the  range  of  the  A.  cylindricus  form,  differing  only  in  the 
absence  of  the  posterior  fold  commonly  seen  in  specimens  of  that  form. 

Unfortunately,  while  Bretsky  and  Bretsky  could  relate  the  relative  abundances  of  the  different 
forms  of  N.  planulatus  to  a known  stratigraphic  sequence,  no  such  sequence  is  available  in  the 
Killey  Bridge  Formation,  and  each  collection  and  locality  is,  in  effect,  isolated.  Many  specimens 
have  only  vague  locality  information. 

On  the  basis  of  the  description,  figures,  and  measurements  given  by  Bretsky  and  Bretsky  it  is 
not  clear  that  N.  cylindricus  is  distinct  from  N.  planulatus,  but  Portlock’s  name  is  retained  pending 
some  closer  comparison. 

nuculoid  gen.  and  sp.  nov. 

Plate  10,  figs.  18,  19;  text-fig.  6 

Material,  horizon,  and  locality.  A single  specimen  IGS  NIL  9240-9241,  showing  nearly  conjoined  valves, 
crushed  but  apparently  not  distorted,  preserved  as  a composite  mould.  From  the  Tirnaskea  Formation  (Ashgill, 
Hirnantian)  of  the  Crocknagargan  Stream  section,  Pomeroy,  Co.  Tyrone  (IGR  H722  737). 


Description.  Shell  obliquely  ovate  and  modioliform  in  appearance.  For  right  valve,  height  13-3  mm,  length 
22-9  mm  giving  a height-length  ratio  of  0-58.  Inflation  of  the  single  valve  at  least  1 -4  mm.  The  small,  prosogyrate 
umbo  is  at  the  anterior  three-tenths.  A slight  umbonal  ridge  runs  to  the  postero-ventral  angle  at  an  angle  of 


text-fig.  6.  Nuculoid  gen.  and  sp.  nov.,  based  on  latex  cast  of  composite  mould  of  right  valve,  IGS  NIL 
9240,  x6-5.  Compare  with  PI.  10,  fig.  19. 


TUNNICLIFF:  LATE  ORDOVICIAN  BIVALVES 


67 


55°  to  the  hinge-line,  giving  the  shell  its  oblique  appearance.  A little  posterior  to  the  umbonal  ridge  is  a 
shallow  sulcus,  reflected  in  the  sculpture  and  sinus  in  the  posterior  margin.  This  sulcus  is  at  an  angle  of  35° 
to  the  hinge-line.  The  posterior  margin  is  otherwise  rounded  ventrally  and  obliquely  slanted  towards  the 
hinge-line.  The  anterior  margin  is  also  obliquely  slanted.  The  ventral  margin  is  straight.  The  hinge-line  is 
straight  on  both  sides  of  the  umbo,  curving  slightly  at  the  posterior  end.  The  anterior  hinge-plate  bears  3-4 
thin  blade-like  parallel  teeth,  becoming  shorter  towards  the  anterior.  The  posterior  hinge-plate  bears  9 + 
similarly  thin  blade-like  parallel  teeth  extending  to  the  postero-dorsal  angle.  Musculature,  ligament,  pallial 
line  unknown.  Sculpture  of  fine  regular  concentric  lines  (about  7 per  mm)  more  pronounced  on  the  posterior 
slope  of  the  shell. 

Discussion.  Bivalves  are  uncommon  in  the  Tirnaskea  Formation  and  are  usually  poorly  preserved. 
It  is  unfortunate  that  this  specimen,  one  of  the  best  preserved  from  that  horizon,  should  prove 
enigmatic.  In  outline  it  might  be  taken  for  Modiolopsis  sp.,  but  the  presence  of  numerous  taxodont 
teeth  both  before  and  behind  the  umbo  precludes  assignment  to  the  Modiomorphidae.  Suspicion 
that  the  teeth  are  the  result  of  slight  deformation  can  be  dismissed  since  the  anterior  and  posterior 
teeth  are  aligned  contrariwise.  This  is  endorsed  by  the  way  in  which  the  posterior  teeth  of  both 
valves  can  be  seen  in  the  specimen.  Had  these  structures  been  produced  by  some  post-depositional 
deformation,  one  would  have  expected  them  all  to  be  parallel.  It  is  possible  but  unlikely,  that  the 
structures  are  related  to  the  ligament. 

Although  generally  smaller  in  size  than  the  Pomeroy  specimen,  Silicula  Jeffreys,  as  described 
and  illustrated  by  Allen  and  Sanders  1973,  (pp.  263-309)  and  Allen  1978  (p.  394)  shows  many 
similar  features:  small  umbones,  ovate,  flattened  form,  fine,  elongate  teeth.  Silicula  is  a modern 
deep-sea  protobranch  of  the  family  Siliculidae  grouped  by  Allen  (1978,  p.  392)  with  the  Malletiidae 
and  others  in  the  Nuculanoida.  The  strong  resemblance  in  form  to  Silicula  suggests  that  the 
Pomeroy  specimen  was  a deep-water  form,  as  would  seem  likely  since  the  Tirnaskea  Formation 
passes  into  the  overlying  graptolitic,  presumably  offshore,  Llandovery  strata  (Mitchell  1977,  p.  5). 


Subclass  pteriomorphia  Beurlen,  1944 
Order  arcoida  Stoliczka,  1871 
Superfamily  cyrtodontacea  Ulrich,  1894 
Family  cyrtodontidae  Ulrich,  1894 
Genus  cyrtodonta  Billings,  1858 

Type  species.  Cyrtodonta  rugosa  Billings,  1858,  by  subsequent  designation  of  Williams  and  Breger  1916,  p.  149. 


Cyrtodonta ? expansa  (Portlock  1843) 


Plate  11,  figs.  1,  3,  4,  9 


v*  pars.  1843 
v*  pars.  1 843 
v(?)  1843 
v(?).  1875 
1878 
v.  1952 
v.  1952 


Modiola  expansa  Portlock,  p.  425,  pi.  33,  fig.  6 [pars',  longer  variety], 

Modiola  Brycei  Portlock,  p.  425,  pi.  33,  fig.  7 [pars:  see  Cyrtodontal  securiformis  (Portlock)]. 
Modiola  expansa  Portlock;  McCoy,  p.  18. 

Modiolopsis  expansa  Portlock;  Baily,  p.  35,  pi.  12,  fig.  2. 

Modiolopsis  expansa  Portlock;  Baily,  1878  [includes  Brycei  in  expansa  with  doubt]. 
Modiodesma  expansum  (Portlock);  Reed,  p.  71,  pi.  3,  fig.  16. 

Goniophora  brycei  (Portlock);  Reed,  p.  78  [pars:  see  Cyrtodontal  securiformis  (Portlock)]. 


Type-specimens.  Lectotype  here  selected,  IGS  GSM  1 2445,  the  type  of  Portlock’s  long  variety  of  M.  expansa 
(1843,  p.  425).  For  Portlock’s  small  variety  see  Cyrtodontal  securiformis  (Portlock).  Lectotype  of  Modiola 
brycei  Portlock  here  selected  IGS  GSM  12443,  Portlock’s  figured  specimen:  other  syntypes  (paralectotypes) 
IGS  GSM  12444,  GSM  22038  (not  certainly  a syntype),  UM  K4206-4210,  4249.  All  from  the  Bardahessiagh 
Formation,  Pomeroy,  Co.  Tyrone,  exact  localities  uncertain,  south  of  Craigbardahessiagh. 


Material,  localities , and  horizon.  A few  specimens  in  IGS  (GSM  12443-5)  and  UM  (K4244,  4247,  74248, 
74249)  all  from  the  Portlock  Collection.  Two  specimens  in  the  Griffith  Collection  at  NMI  may  belong  to  this 
species.  All  composite  moulds  from  the  Bardahessiagh  Formation,  locality  as  above. 


68 


PALAEONTOLOGY,  VOLUME  25 


Measurements  of  type-specimen.  IGS  GSM  12445:  H 29-6  mm,  L 50-6  mm,  AL  8-6  mm. 

Description.  Elongate,  obliquely  ovate,  becoming  higher  towards  the  posterior  end,  with  a height-length  ratio 
between  about  0-6  and  0-7.  The  straight  hinge-line  is  about  half  as  long  as  the  total  length  of  the  shell. 
Antero-dorsal  margin  almost  straight  or  gently  curved  and  meeting  the  hinge-line  at  an  angle  of  about  150° 
Ventral  margin  almost  straight.  Umbones  prosogyrate  and  placed  at  about  0T5  of  the  length  from  the  front, 
projecting  a little  above  the  hinge-line.  A rounded  umbonal  ridge  extends  backwards  at  about  35°  from  the 
hinge-line  to  the  ventral  margin  at  the  point  furthest  from  the  umbo.  Preservation  poor,  but  two  posterior 
lateral  teeth  can  be  seen  at  the  posterior  end  of  the  hinge-line  in  the  right  valve  and  one  in  the  left, 
anterior  teeth  are  not  known.  The  anterior  adductor  muscle  scar  is  small,  the  posterior  is  unknown.  Maximum 
inflation  seen  in  a single  undistorted  valve  is  4T  mm  in  a valve  45  mm  long.  Sculpture  of  irregular  concentric 
striae. 

Discussion.  C.?  expansa  and  C.?  securiformis  (Portlock)  show  the  effects  of  distortion  more  than 
other  species  from  the  Bardahessiagh  Formation  and  this  has  influenced  their  nomenclatorial 
history.  Portlock  (1843,  p.  425)  noted  two  varieties  of  M.  expansa ; a longer  variety,  which  he 
figured  (1843,  pi.  33,  fig.  6),  and  a smaller  variety  which  he  described  briefly.  The  smaller  variety 
is  now  placed  in  C.?  securiformis.  Portlock’s  M.  brycei  (1843,  p.  425,  pi.  33,  fig.  7)  was  based  on 
specimens  which  he  described  as  ‘somewhat  distorted  by  pressure’  (1843,  p.  426).  Reed  (1952) 
removed  the  specimens  of  Portlock’s  smaller  variety  from  M.  expansa  and  placed  them  in  two  new 
species,  Orthodesma  tyronense  (1952,  p.  71)  and  Whiteavesia  subexpansa  (1952,  p.  72).  These  are 
now  placed  in  C.?  securiformis.  Examination  of  Portlock’s  type  material,  some  of  which  was  not 
available  to  Reed,  shows  that  those  specimens  which  both  Portlock  and  Reed  placed  in  brycei  are 
either  expansa  or  securiformis  which  have  been,  as  Portlock  observed,  crushed  dorso-ventrally.  In 
every  case,  the  line  of  the  umbonal  ridge  has  proved  the  weakest  and  has  produced,  under  pressure, 
a sharp  keel-like  appearance,  described  by  Reed  as  ‘acutely  carinated’.  It  is  clear  that  the  specimens 
of  M.  brycei  which  Portlock  figured  and  described  belong  to  C.?  expansa,  while  those  which  he 
described  as  ‘young  individuals’  (1843,  p.  426)  are  C.?  securiformis. 

Although  the  detail  of  the  hinge  is  not  well  preserved,  and  the  anterior  teeth  are  not  known, 
the  position  and  appearance  of  the  posterior  lateral  teeth  suggests  that  Cyrtodonta  is  the  appropriate 
genus  for  both  expansa  and  securiformis  but  in  the  absence  of  the  anterior  dentition  the  genus 
remains  doubtful.  The  presence  of  the  posterior  teeth,  which  Reed  (1952,  p.  71)  apparently 
interpreted  as  a ligament  groove,  precludes  the  assignment  of  either  species  to  the  edentulous 
Modiolopsis  [ = Modiodesma,  —Orthodesma  Laroque  and  Newell,  1969,  p.  N397].  However,  C.? 
expansa  bears  a strong  resemblance  in  form  to  Sphenolium  sp.  as  figured  by  Pojeta  (1978,  pi.  8, 
fig.  6)  which,  although  previously  placed  in  synonymy  with  Modiolopsis  (as  in  the  Treatise  etc., 
p.  N397),  Pojeta  (1978,  p.  235)  now  appears  to  regard  as  a cyrtodontid.  The  strongly  prosocline 
form  of  C.?  expansa  distinguishes  it  from  the  species  figured  from  North  America  by  Ulrich, 
Pojeta,  and  others  which,  in  illustration,  tend  to  have  a more  rounded  outline. 


v*  pars.  1843 
v*  1843 
v*  pars.  1843 
v.  1846 
1878 
? non.  1946 
v* 1952 
v* 1952 
v* 1952 
v. 1952 
v.  1952 


Cyrtodonta?  securiformis  (Portlock  1 843) 

Plate  11,  figs.  5-7,  10,  12-14 

Modiola  expansa  Portlock,  p.  425  [pars:  smaller  variety], 

Modiola  securiformis  Portlock,  p.  425,  pi.  33,  fig.  8. 

Modiola  Brycei  Portlock,  p.  425  [pars:  see  Cyrtodonta?  expansa  (Portlock)]. 

Modiola  securiformis  Portlock;  McCoy,  p.  18. 

Modiolopsis  securiformis  Portlock;  Baily,  p.  28. 

Whitella  cf.  brycei  (Portlock);  Lamont,  p.  366,  pi.  1,  figs.  1,  2. 

Orthodesma  tyronense  Reed,  p.  71,  pi.  3,  fig.  17. 

Whiteavesia  subexpansa  Reed,  p.  72,  pi.  4,  fig.  2. 

Modiolopsis  concentrica  Hall  and  Whitfield  var.  simulans  Reed,  p.  69,  pi.  3,  fig.  14. 
Modiolopsis  securiformis  (Portlock);  Reed,  p.  70,  pi.  3,  fig.  15. 

Goniophora  brycei  (Portlock);  Reed,  p.  78  [pars:  see  Cyrtodonta?  expansa  (Portlock)]. 


TUNNICLIFF:  LATE  ORDOVICIAN  BIVALVES 


69 


Type-specimens.  Lectotype  of  Modiola  securiformis  here  selected,  IGS  GSM  12448,  Portlock’s  figured 
specimen:  paralectotype  IGS  GSM  12449.  Type  of  M.  brycei:  see  Cyrtodonta?  expansa  (Portlock).  Holotype 
of  Orthodesma  tyronense  IGS  GSM  12447  by  original  designation  of  Reed  (1952,  p.  72).  Holotype  of  Whiteavesia 
subexpansa  IGS  GSM  12446  by  original  designation  of  Reed  (1952,  p.  73).  Holotype  of  Modiolopsis  concentrica 
simulans  Reed  SM  A 1 6294a, b by  original  designation  of  Reed  (1952,  p.  70).  All  from  the  Bardahessiagh 
Formation,  Pomeroy,  Co.  Tyrone,  exact  localities  uncertain,  south  of  Craigbardahessiagh. 

Material , localities,  and  horizon.  Specimens  in  IGS  (GSM  12446-9,  22038,  24172,  GU1880,  Zfl021)  and  UM 
(1920-845,  K4206-4210,  4245-4246).  All  composite  moulds,  from  the  same  locality  and  horizon  as  given 
above. 

Measurements.  Type-specimen,  GSM  12448,  which  is  compressed  dorso-ventrally;  for  left  valve,  the  least 
distorted:  H 13-6  mm,  L 30-7  mm,  AL  8-2  mm.  GSM  12446,  undistorted,  H 30  mm,  L 351  mm,  AL  9-7  mm. 

Description.  Obliquely  ovate,  becoming  broader  towards  the  posterior,  with  a height-length  ratio  of  about 
0-9,  but  very  variable  in  the  distorted  specimens  available.  Maximum  inflation  seen  in  a single  undistorted 
valve  is  3-9  mm  in  a valve  35-6  mm  long.  The  hinge-line  is  straight  or  very  gently  curved  and  is  about  half 
as  long  as  the  total  length  of  the  shell.  At  the  posterior  end  of  the  hinge-line  are  2+  posterior  lateral  teeth 
in  the  right  valve  and  1 + in  the  left,  but  the  preservation  of  the  material  prevents  accurate  description.  The 
antero-dorsal  margin  is  curved  and  meets  the  hinge-line  at  an  angle  of  about  130°.  The  ventral  margin  is 
almost  straight.  Umbones  prosogyrate,  placed  at  about  the  anterior  quarter  projecting  slightly  above  the  hinge- 
line. A rounded  umbonal  ridge  extends  backwards  at  about  50°  from  the  hinge-line  to  the  ventral  margin  at  its 
furthest  point  from  the  umbo.  The  anterior  teeth  are  unknown.  The  musculature  is  unknown.  Sculpture  of 
irregular  concentric  striae. 

Discussion.  See  discussion  of  C.?  expansa.  Portlock  (1843,  p.  425)  described  his  specimens  of 
Modiola  securiformis  as  being  ‘partly  distorted  by  pressure’.  In  the  lectotype  (GSM  12448),  the 
right  valve  approaches  the  M.  brycei  Portlock  form,  with  the  umbonal  ridge  becoming  sharpened; 
the  left  valve  is  less  distorted  and  has  the  form  of  the  smaller  variety  of  M.  expansa.  Another 
specimen  (UM  K4245)  shows  two  sets  of  distorted  conjoined  valves  lying  on  the  same  plane, 
oriented  at  about  45°  to  each  other.  In  each  the  left  valve  is  brycei- like  in  form,  while  the  right 
valve  retains  its  shape  with  perhaps  a reduction  in  height  and  an  accentuation  of  the  umbonal 
ridge.  Reed’s  (1952)  species  O.  tyronense  and  W.  subexpansa  are  here  interpreted  as  almost 
undistorted  valves  of  C.?  securiformis.  The  types  (GSM  12446-12447)  are  so  alike,  apart  from 
being  opposite  valves,  that  it  is  hard  to  understand  why  Reed  separated  them  specifically,  let  alone 
generically.  Portlock’s  ‘younger  individuals’  of  M.  brycei  are  badly  distorted  C.?  securiformis. 

The  reasons  for  referring  securiformis  to  ‘ Cyrtodonta ? are  discussed  under  C.?  expansa.  While 
it  is  possible  that  C.?  expansa  and  C.?  securiformis  may  be  synonymous,  there  are  clear  differences 
between  them:  securiformis  is  noticeably  more  truncate  posteriorly  in  appearance,  its  postero-dorsal 
margin  meeting  the  hinge-line  at  a steeper  angle  than  in  expansa , and  its  umbones  being  relatively 
more  posterior. 

Reed’s  (1952)  Modiolopsis  concentrica  simulans  seems  to  be  a specimen  of  C.?  securiformis  which, 
as  a result  of  distortion,  is  higher  and  shorter  than  normal.  Hind’s  figure  of  C.  transversa  (1910, 
pi.  4,  figs.  19,  19a,  ?20)  from  the  Drummuck  Group  (Ashgill)  closely  resembles  C.?  securiformis, 
but  his  specimen  (BM  L49860)  is  more  inflated  and  slightly  more  prosocline  than  securiformis. 
The  specimen  from  the  Drummuck  Group  figured  by  Lamont  (1946,  pi.  1,  figs.  1,  2)  as  Whitella 
cf.  brycei  (Portlock)  should  be  compared  with  C.  transversa  rather  than  with  C.?  securiformis. 


Cyrtodonta?  sp. 

Plate  13,  fig.  1 

v.  71843  Avicula  orbicularis  J.  de  C.  Sowerby  in  Murchison;  Portlock,  pp.  425,  755  (Synoptical  Table) 
[pars:  see  Cycloconcha?  speciosa  (McCoy)]. 

Material,  horizon,  and  locality.  A single  right  valve,  UM  K4194  from  the  Bardahessiagh  Formation,  exact 
locality  uncertain  but  south  of  Craigbardahessiagh,  Pomeroy. 


70 


PALAEONTOLOGY,  VOLUME  25 


Measurements.  L 26-2  mm,  H 24-5  mm,  AL  2-4  mm,  inflation  of  the  single  valve  3-0  mm,  obliquity  40°. 

Description.  Subcircular  with  the  umbo  at  about  the  anterior  one-tenth.  Umbo  incomplete.  Obliquity  of  valve 
about  40°.  The  antero-dorsal  margin  meets  the  postero-dorsal  margin  at  an  angle  of  about  135°.  Dentition 
poorly  preserved  but  two  parallel  posterior  lateral  teeth  are  suggested  by  faint  grooves;  anterior  teeth  unknown. 
Anterior  muscle  scar  faint,  lying  about  half-way  between  the  dorsal  and  ventral  margins;  posterior  musculature 
unknown.  Sculpture  of  fine  concentric  striae  visible  close  to  the  anterior  margin,  and  faint  traces  of  a coarser 
concentric  ornament  ventrally. 

Discussion.  Although  no  Portlock  label  remains  associated  with  this  specimen  it  is  almost  certainly 
one  which  he  likened  to  Avicula  orbicularis  Sowerby  in  Murchison  1839;  in  shape  it  compares  quite 
well  with  the  type  of  A.  orbicularis  (IGS  Geol.  Soc.  Coll.  6888)  but  this  is  much  larger  with  no 
visible  posterior  lateral  teeth.  Were  it  not  for  the  more  rounded  shape,  this  specimen  might  have 
been  placed  in  ?Cycloconcha  speciosa,  but  in  that  species  one  would  expect  to  see  three  posterior 
lateral  teeth  in  a right  valve  while  in  K4194  only  two  are  apparent. 

Genus  Vanuxemia  Billings,  1858 

Type  species.  Vanuxemia  inconstans  Billings  by  subsequent  designation  of  Miller  1889,  from  the  Black  River 
and  Trenton  Groups. 


Vanuxemia?  contorta  (Portlock  1843) 

Plate  13,  figs.  4,  8 

v*  1843  Inoceramus  contortus  Portlock,  p.  422,  pi.  33,  fig.  5. 
v.  1952  Vanuxemia  contorta  (Portlock);  Reed,  p.  68,  pi.  3,  fig.  12. 

Type-specimen.  IGS  GSM  12435,  holotype  by  monotypy;  Portlock’s  brief  description  is  of  one  shell. 

Material,  locality,  and  horizon.  The  type  specimen  (a  fragmentary  left  valve)  and  a distorted  right  valve,  IGS 
Zf  1020,  both  showing  external  features  and  both  from  the  Killey  Bridge  Formation,  Pomeroy,  exact  locality 
uncertain. 


EXPLANATION  OF  PLATE  1 1 

Figs.  1,  3,  4,  9.  Cyrtodonta?  expansa  (Portlock  1843).  1,  4,  conjoined  valves,  IGS  GSM  12443,  lectotype  of 

Modiola  brycei  Portlock,  1843;  1,  lateral  view  of  left  valve;  4,  dorsal  view  of  both  valves.  3,  right  valve, 
UM  K4247,  probably  the  specimen  figured  by  Baily,  1875,  pi.  12,  fig.  2.  9,  left  valve,  lectotype,  IGS  GSM 
12445.  All  preserved  as  ?composite  moulds,  Bardahessiagh  Formation  (Caradoc),  south  of  Craigbarda- 
hessiagh,  Pomeroy,  x 1 . 

Fig.  2.  Pterineid?  gen.  and  sp.  indet.  ?external  cast  left  valve  IGS  NIL  8981,  Killey  Bridge  Formation 
(Ashgill,  Cautleyan),  Crocknagargan  Stream  section  (IGR  H721737),  x4. 

Figs.  5-7,  10,  12-14.  Cyrtodonta?  securiformis  (Portlock  1843).  5,  right  valve,  IGS  GSM  12447,  holotype 

of  Orthodesma  tyronense  Reed,  1952,  x 1.  6,  left  valve,  IGS  GSM  12446,  holotype  of  Whiteavesia  sub- 

expansa  Reed,  1952,  x 1.  7,  10,  conjoined  valves  lectotype,  IGS  GSM  12448;  7,  dorsal  view,  x 1 10, 

lateral  view  of  left  valve,  x 1|.  12,  external  cast  of  left  valve,  SM  A 16294a,  holotype  of  Modiolopsis 

concentrica  var.  simulans  Reed,  1952,  x 2.  13,  14,  slab  bearing  two  conjoined  pairs  of  valves,  UM  K4245; 

13,  dorsal  view  to  show  the  distortion  of  both  left  valves  which  has  produced  the  sharp  ‘umbonal  ridge’, 
leading  to  confusion  in  the  past  with  Goniophora,  x 1;  14,  lateral  view  of  both  right  valves,  x 1.  All  except 
12  preserved  as  ?composite  moulds,  Bardahessiagh  Formation  (Caradoc),  south  of  Craigbardahessiagh, 
Pomeroy. 

Figs.  8,  11,  15.  Semicorallidomus?  sp.  8,  latex  cast  of  internal  mould  of  left  valve  showing  single  small 
tooth,  IGS  GU  1739,  x4.  11,  composite  mould  of  left  valve,  IGS  GSM  12450,  x 3.  15,  composite 

mould  of  right  valve,  IGS  GSM  12451,  x3.  All  Bardahessiagh  Formation  (Caradoc),  south  of  Craig- 
bardahessiagh, Pomeroy. 


PLATE  1 1 


TUNNICLIFF,  Ordovician  bivalves 


72 


PALAEONTOLOGY,  VOLUME  25 


Measurements.  Holotype  incomplete.  For  Zf  1020  H 32-9  mm,  L 25  1 mm,  maximum  oblique  dimension 
35-8  mm,  apparent  angle  of  obliquity  45  to  50°. 

Description.  Tall,  the  height  being  about  1-3  the  length.  Obliquity  45  to  50°.  Length  of  hinge-line  about  half 
that  of  valve.  Dentition  and  musculature  unknown.  Umbo  anterior  but  not  terminal,  prosogyrate.  Valve 
inflated  to  8-7  mm  in  Zf  1020  after  reduction  by  compression  which  has  sharpened  the  umbonal  ridge,  especially 
towards  the  umbo.  Sculpture  of  fine  concentric  striae,  3-5  per  mm,  becoming  stronger  towards  the  margin. 

Discussion.  Neither  specimen  is  well  preserved;  the  type  shows  only  the  posterior  region  and  Zf 
1020  is  crushed.  Portlock  labelled  Zf  1020  as  ‘indeterminate’.  Reed  did  not  use  Zf  1020  when  he 
assigned  contort  a to  Vanuxemia  with  some  reservation  but  the  more  complete  Zf  1020,  showing  the 
anterior  position  of  the  umbones,  supports  his  opinion.  However,  the  nature  of  the  material  demands 
that  the  generic  assignment  be  qualified  with  a ?.  It  is  here  proposed  that  the  species  be  restricted  to 
these  specimens  and  that  the  species  be  regarded  as  otherwise  unrecognizable. 


Remarks.  In  describing  the  ambonychiids,  reference  is  made  to  Pojeta  1966,  in  which  a terminology 
for  certain  angles  and  dimensions  is  standardized  (text-fig.  7),  and  where  a comprehensive  review 
of  North  American  ambonychiids  is  given. 


Type-species.  Ambonychia  radiata  Hall,  1847,  p.  163,  by  subsequent  designation  of  Stoliczka  (1871,  p.  387). 


Derivation  of  name.  From  the  Latin  arundinea,  of  reeds,  reedy,  referring  both  to  the  ornament  and  by  allusion 
to  F.  R.  C.  Reed. 

v.  1843  Uncites  gryphus  Schlotheim  var.;  Portlock,  p.  455,  pi.  25a,  fig.  8. 

. 1878  Ambonychia  gryphus-,  Baily,  p.  28. 

v Ipars.  1910  Byssonychia  radiata  Hall;  Hind,  p.  487,  pi.  1,  figs.  20,  21,  non  figs.  19,  22. 
v.  1952  Byssonychia  gryphus  (Portlock);  Reed,  p.  77. 


Order  pterioida  Newell,  1965 
Suborder  pteriina  Newell,  1965 
Superfamily  ambonychiacea  Miller,  1877 
Family  ambonychiidae  Miller,  1877 


Genus:  ambonychia  Hall,  1847 


Ambonychia  arundinea  sp.  nov. 


Plate  13,  figs.  2,  3 


Z 


G 


B 


A 


H 


text-fig.  7.  Linear  and  angular  measurements  used  in  describing  Ambonychiids,  after  Pojeta,  1966.  a— angle 
alpha,  b— angle  beta,  G— angle  gamma,  l— length,  h— height,  d— greatest  dimension,  z— length  of  hinge. 


TUNNICLIFF:  LATE  ORDOVICIAN  BIVALVES 


73 


Type-specimens.  Holotype,  right  valve  IGS  GSM  24302.  Paratypes,  left  valves  IGS  GSM  24303,  BM  LL 
40004,  right  valves  IGS  GU  1452,  1566-1567  (counterparts),  and  UM  1920-835  (larger  specimen).  All  external 
moulds  from  the  Killey  Bridge  Formation  of  Pomeroy,  Co.  Tyrone.  GSM  24302,  24303  and  UM  1920-835 
are  not  precisely  localized  but  probably  come  from  locality  3 of  Mitchell  (1977),  the  Little  River  section  (Irish 
grid  ref.  H7297  7268),  the  locality  for  GU  1566-1567.  GU  1452  and  BM  LL  40004  are  from  locality  2 of 
Mitchell,  Warren  Wood  River  section  (Irish  grid  ref.  H7130  7128). 

Description.  Tall  subrectangular  Ambonychia.  The  greatest  dimension  at  an  angle  of  50  to  55°  to  the  hinge-line. 
Length  about  0-7  of  the  greatest  dimension  and  height  about  1-25  the  length.  The  hinge  length  is  0-4  of  the 
length.  Greatest  inflation  of  a single  valve  seen  is  4 mm  in  a valve  33  mm  high  (UM  1920-835).  Prosocline, 
with  an  angle  gamma  of  80  to  85  . Umbones  not  rounded  but  carinate  in  the  sense  of  Pojeta  1962  (i.e.  the 
anterior  portion  of  the  valve  is  flattened  antero-posteriorly),  and  projecting  between  1 -4  mm  and  3 0 mm  above 
the  hinge-line.  Valve  surface  mainly  gently  convex.  The  nature  of  the  byssal  gape  is  not  discernible  but  the 
byssal  sinus  is  shallow.  Dentition  unknown  although  what  may  be  the  damaged  remains  of  the  posterior 
laterals  (?2  in  right  valve)  are  visible  in  GSM  24302.  The  ligament  area  is  damaged  or  obscured  in  all  the 
present  specimens  and  none  shows  musculature.  About  forty  costae  (range  seen  33  + , 37  + , 40,  41,  43). 

Discussion.  Portlock  briefly  described  his  specimens  (GSM  24302,  24303)  as  brachiopods.  Hind 
examined  Portlock’ s specimens  and  assigned  them,  in  museo  to  Byssonychia,  as  noted  by  Reed 
(1952,  p.  78)  who  gave  the  first  full  description,  but  had  only  these  two  specimens  before  him.  He 
discussed  the  similarity  and  differences  between  these  specimens  and  those  Byssonychia  described 
from  Girvan  by  Hind  (1910,  p.  487,  pi.  19)  and  other  species  from  North  America.  However,  he 
chose  to  retain  the  specific  name  gryphus,  and  wrongly  attributed  its  authorship  to  Portlock. 

The  North  American  Richmondian  species  Byssonychia  richmondensis  Ulrich,  and  B.  robust  a 
(Miller),  as  described  by  Pojeta  1962,  and  subsequently  (Pojeta  1966)  transferred  to  Ambonychia , 
have  similar  numbers  of  costae  and  similar  dimensions  to  the  Pomeroy  specimens,  but  the  Irish 
species  has  a more  elongate  appearance,  the  ratio  of  its  length  to  its  greatest  dimension  is  0-7  while  for 
Ambonychia  richmondensis  and  A.  robusta  it  is  0-5-0-6  and  0-6-0-75  respectively  (based  on  Pojeta’s 
figures).  Both  the  American  species  tend  more  towards  the  acline  form  by  about  10°. 

Of  Hind’s  figures  (1910,  pi.  1,  figs.  19-22)  of  B.  radiata,  which  Pojeta  (1962,  p.  184)  excluded 
from  B.  radiata , figs.  19,  22  are  comparable  to  the  Richmondian  species  B.  suberecta  Ulrich,  but 
figs.  20,  21  (BM  L49766-49767)  are  close  in  all  respects  to  A.  arundinea  but  have  a slightly  higher 
number  of  costae  (44  to  48). 


Genus  cleionychia  Ulrich,  1892 
Type  species.  By  original  designation  Ambonychia  lamellosa  Hall,  1862. 


Cleionychia  transversa  (Portlock  1843) 
Plate  12,  figs.  1-6 


v*  1843 
v pars.  1843 

v.  1952 
v.  1952 


Inoceramus  transversus  Portlock,  p.  423,  pi.  33,  fig.  11. 

Inoceramus  vetustus  (J.  de  C.  Sowerby)  Var.  priscus  Portlock,  p.  423,  pi.  33,  figs.  2,  3,  non 
fig.  1. 

Clionychia  subovalis  Reed,  p.  76,  pi.  4,  fig.  7. 

Clionychia  subquadrata  Reed,  pp.  76-77,  pi.  4,  fig.  8. 


Type  specimens.  Lectotype  of  Cleionychia  transversa  IGS  GSM  12439,  selected  by  Reed  1952,  p.  77;  holotype 
of  C.  subovalis  Reed  IGS  GSM  12436  by  original  designation;  lectotype  of  C.  subquadrata  Reed  here  selected, 
IGS  GSM  12437,  the  specimen  figured  by  Reed  (1952,  pi.  4,  fig.  8).  Paralectotypes  of  transversa  IGS  GSM  12440, 
UM  K4199,  all  from  the  Portlock  Collection;  paralectotype  of  C.  subquadrata  IGS  GSM  24304,  from  the 
Wyatt-Edgell  Collection. 


Other  material.  IGS  GSM  24305,  from  the  Wyatt-Edgell  Collection. 


74  PALAEONTOLOGY,  VOLUME  25 

Horizon  and  locality.  All  from  the  Bardahessiagh  Formation,  exact  locality  uncertain  but  south  of 
Craigbardahessiagh,  Pomeroy. 

Measurements.  For  the  lectotype  IGS  GSM  12439,  L 67-6+  mm,  H 40- 1 mm,  maximum  inflation  of  the  single 
valve  9-8  mm. 

Description.  Transversely  elongate  becoming  relatively  longer  with  increased  size  and  with  height  about  0-6 
of  the  length  in  large  specimens.  Inflation  greatest  below  and  slightly  posterior  to  the  umbo,  gradually 
diminishing  towards  the  back.  Prosogyrate  umbo  anterior  and  almost  terminal  and  rising  above  the  hinge-line. 
Straight  hinge-line  about  two-thirds  of  the  valve  length  with  duplivincular  ligament  area  extending  along 
dorsal  margin.  Anterior  end  truncate,  angle  gamma  about  90°.  Anterior  slope  meets  commissure  at  about 
90°  but  may  form  a distinctly  obtuse  angle  in  distorted  specimens.  Ventral  margin  curving  rapidly  up  to 
anterior  margin  and  more  gently  towards  the  posterior  margin  which  is  unclear  in  all  the  larger  specimens 
available,  but  rounded  and  continuous  with  the  ventral  margin  in  smaller  specimens.  Angle  beta  about  140°. 
Musculature  and  dentition  unknown.  Sculpture  of  coarse  concentric  rugae,  undulating  in  section. 

Discussion.  It  is  likely  that  the  specimens  on  which  Reed  based  his  species  C.  subovalis  and  C. 
subquadrata  (1952,  pp.  76-77)  are  juveniles  of  C.  transversa.  This  is  supported  by  measurement  of 
appropriate  early  growth  stages  of  clear  examples  of  C.  transversa.  It  is  found  that  the  height  to 
length  ratio  decreases  from  about  0-9  at  25  mm  long  growth  stage  to  0-6  at  60  mm. 

Early  growth  stages  of  C.  transversa  closely  resemble  C.  undata  Emmons,  but  the  latter  has  a 
slightly  more  acute  angle  gamma  (apparently  about  85°  in  Pojeta’s  illustrations,  1966,  pi.  34,  figs. 
1 -5)  and  there  are  no  figures  of  late  growth  stages  comparable  to  those  seen  in  C.  transversa. 

Cleionychia  prisca  (Portlock  1843) 

Plate  12,  figs.  8,  12 

v*  pars.  1843  Inoceramus  vetustus  (J.  de  C.  Sowerby)  Var.  priscus  Portlock,  p.  423,  pi.  33,  fig.  1,  non 
figs.  2,  3. 

. 1878  Ambonychia  undata  Hall;  Baily,  p.  28. 
v.  1952  Ambonychinia  prisca  (Portlock);  Reed,  p.  73. 


EXPLANATION  OF  PLATE  12 

Figs.  1-6.  Cleionychia  transversa  (Portlock  1843).  1,  left  valve,  lectotype,  IGS  GSM  12439.  2,  left  valve, 

UM  K4199.  3,  right  valve,  IGS  GSM  12436,  holotype  of  Cleionychia  subovalis  Reed,  1952.  4,  right  valve, 

IGS  GSM  12440.  5,  left  valve,  IGS  GSM  12437,  lectotype  of  Cleionychia  subquadrata  Reed,  1952. 

6,  left  valve,  IGS  GSM  24304.  All  external  casts,  Bardahessiagh  Formation  (Caradoc),  south  of 
Craigbardahessiagh,  Pomeroy,  x 1. 

Figs.  7,  11,  13.  Cleionychia  incognita  sp.  nov.  7,  right  valve,  IGS  GSM  24310.  11,  left  valve,  holotype, 

IGS  GSM  12441;  note  the  bryozoan  of  the  type  referred  by  Portlock  (1843,  p.  360)  to  Entobia  antiqua 
Portlock.  13,  small  left  valve,  IGS  GSM  24311.  All  external  casts,  Bardahessiagh  Formation  (Caradoc), 
south  of  Craigbardahessiagh,  x 1^. 

Figs.  8,  12.  Cleionychia  prisca  (Portlock  1843).  8,  right  valve,  UM  K4200.  12,  right  valve,  lectotype, 

IGS  GSM  12438.  Both  external  casts,  Bardahessiagh  Formation  (Caradoc),  south  of  Craigbardahessiagh, 
Pomeroy,  x 1. 

Fig.  9.  Corallidomus  concentrica  (Hall  and  Whitfield  1875),  ?composite  mould  right  valve,  UM  1920-843, 
Killey  Bridge  Formation  (Ashgill,  Cautleyan),  Pomeroy,  exact  locality  uncertain,  x 2. 

Fig.  10.  Corallidomus?  sp.  ?external  cast  of  right  valve,  NMI  G.4.  1979,  Killey  Bridge  Formation  (Ashgill, 
Cautleyan),  Pomeroy,  exact  locality  uncertain,  x 2. 

Fig.  14.  Ambonychiopsis  suspecta  (Reed  1952),  ?composite  mould  of  left  valve,  holotype,  IGS  GSM  22103, 
Bardahessiagh  Formation  (Caradoc),  south  of  Craigbardahessiagh,  x 1. 

Fig.  15.  Goniophora  sp.,  external  cast  of  right  valve,  IGS  Zs  2751,  Killey  Bridge  Formation  (Ashgill, 
Cautleyan),  locality  3 of  Mitchell  1977  (IGR  H7297  7268),  x4. 


PLATE  12 


TUNNICLIFF,  Ordovician  bivalves 


76 


PALAEONTOLOGY,  VOLUME  25 


Type-specimens,  horizon,  and  localities.  Lectotype  IGS  GSM  12438,  right  valve,  (with  a second  valve  only 
partly  visible)  selected  by  Reed  1952,  p.  74;  paralectotype  (not  known  to  Reed)  UM  K4200.  Both  from  the 
Bardahessiagh  Formation  from  Portlock’s  locality  ‘sheet  37  no.  6’  south  of  Craigbardahessiagh,  Pomeroy. 

Other  material.  A single  specimen  in  the  Griffith  Collection,  NMI;  locality  and  horizon  as  for  the  type 
specimens. 

Measurements.  For  the  lectotype,  L 66-5+  mm,  H 60-2+  mm,  maximum  inflation  of  the  single  valve  12-2  mm, 
greatest  dimension  (oblique)  74T  mm,  angle  alpha  c.  55°,  angle  gamma  c.  85°,  angle  beta  c.  120°,  length  of 
hinge  uncertain,  height  of  umbo  above  hinge-line  8 mm. 

Description.  Obliquely  ovate  form,  narrow  anteriorly  and  expanding  rapidly  towards  the  posterior  end.  Height 
about  0-9  of  length;  precise  length  and  detail  of  hinge-line  unknown.  Greatest  dimension  approximately  1-25 
times  length;  angle  alpha  about  50°,  angle  gamma  about  80  to  85°,  angle  beta  about  120°.  Maximum  inflation 
seen  in  a single  valve,  in  the  lectotype  12.2  mm.  Umbo  terminal.  Antero-ventral  slope  steep,  meeting  the  plane 
of  commissure  at  an  angle  of  90  to  100°.  Posterior  surface  convex,  becoming  flatter  towards  the  posterior 
end.  Ligament,  musculature,  and  dentition  unknown.  Sculpture  of  coarse  concentric  rugae  undulating  in 
section,  about  2 per  10  mm  along  the  plane  of  the  greatest  dimension. 

Discussion.  Although  not  well  preserved,  the  specimens  of  Cleionychia  prisca  (Portlock)  are 
sufficiently  distinct  to  be  separated  from  C.  transversa  and  C.  incognita  sp.  nov.  While  the  sculpture 
in  C.  prisca  is  very  like  that  in  C.  transversa,  angle  gamma  is  more  acute  in  C.  prisca  and  the  shell 
has  a generally  more  oblique  and  inflated  appearance.  Angle  alpha  in  C.  prisca  is  less  acute  than 
in  C.  incognita  which  lacks  the  strong  concentric  rugae  of  C.  prisca. 

Pojeta  (1966,  pp.  177,  180)  pointed  out  that  Reed’s  figures  were  poor  and  that  it  was  not  clear 
(despite  his  comments,  1952,  p.  74)  how  Reed  distinguished  Cleionychia  spp.  from  Ambonychinia 
spp.  None  of  the  specimens  which  Reed  assigned  to  Cleionychia  is  greatly  inflated  but  they  have 
concentric  undulating  sculpture;  those  which  he  placed  in  Ambonychinia  Isberg  are  more  inflated, 
elongate  obliquely,  and  have  prominent  umbones  and  a more  acute  angle  gamma,  but  while  having 
concentric  ornament  they  apparently  lack  the  radial  costellae  diagnostic  of  Ambonychiopsis  Isberg 
although  in  shape  they  resemble  the  latter.  The  specimens  treated  by  Reed  as  Ambonychinia  are 
here  assigned  to  Cleionychia  following  the  suggestion  inherent  in  Pojeta’s  remarks  (1966,  p.  177). 

Reed  (1952,  p.  74)  pointed  out  the  similarity  of  C.  prisca  to  one  of  Isberg’s  (1934,  pi.  4,  fig.  6) 
figures  of  Ambonychinia  corrugata  (Lindstrom)  and  also  noted  the  different  appearance  of  the 
remaining  figures  of  A.  corrugata  (Isberg  1934,  pi.  4,  figs.  1-5,  7).  The  specimen  in  Isberg’s  pi.  4, 
fig.  6 appears  to  have  stronger  rugae  than  seen  in  C.  prisca  and  each  of  its  angles  alpha,  beta,  and 
gamma  is  a little  more  acute  than  seen  in  C.  prisca. 


Cleionychia  incognita  sp.  nov. 

Plate  12,  figs.  7,  11,  13 

Derivation  of  name.  From  the  Latin  incognita  unknown  or  unrecognized,  referring  to  the  nomenclatorial 
history  of  the  type  specimens. 

v.  1843  Inoceramus  trigonus  (Munster);  Portlock,  p.  422,  pi.  33,  figs.  4,  4a. 

. 1878  Ambonychia  trigona  Portlock;  Baily,  p.  28. 
v.  1952  Ambonychinia  cf.  amygdalina  (Hall);  Reed,  p.  75,  pi.  4,  fig.  5. 
v.  1952  Ambonychinia  cf.  intermedia  Isberg;  Reed,  p.  75,  pi.  4,  fig.  6. 
v.  1952  Ambonychinia  cf.  volvens  Isberg;  Reed,  p.  74,  pi.  4,  fig.  4. 

Type  specimens.  Holotype  IGS  GSM  12441,  a left  valve,  figured  by  Portlock  1843,  pi.  33,  fig.  4;  paratypes 
IGS  GSM  24310,  24311  from  the  Wyatt-Edgell  Collection,  a right  and  a small  left  valve  and  IGS  GSM 
24306,  24307,  104237,  three  distorted  specimens  showing  conjoined  valves,  all  from  the  Wyatt-Edgell 
Collection. 

Horizon  and  localities.  Bardahessiagh  Formation,  exact  localities  uncertain  but  south  of  Craigbardahessiagh, 
Pomeroy. 


TUNNICLIFF:  LATE  ORDOVICIAN  BIVALVES 


77 


Measurements.  For  the  holotype,  L 41-4  mm,  H 34-7  mm,  greatest  dimension  (oblique)  49-3+  mm,  angle 
alpha  c.  40°,  angle  gamma  75°,  angle  beta  c.  120°,  maximum  inflation  of  the  single  valve  11-5  mm,  length  of 
hinge  29-6+  mm. 

Description.  Obliquely  ovate  to  subquadrate  in  form,  narrow  anteriorly  and  expanding  rapidly  towards  the 
posterior  end.  Height  about  0-8  to  0-9  of  length.  Straight  hinge-line  about  0-7  to  0-75  length.  Greatest  dimension 
approximately  IT  to  1-2  times  length;  angle  alpha  40  to  50°.  Angle  gamma  70  to  80°,  angle  beta  115  to  135°. 
Maximum  inflation  of  a single  valve  seen  1 1 -5  mm  in  the  holotype.  Umbo  terminal,  rising  above  the  hinge-line. 
Antero- ventral  slope  steep,  meeting  the  plane  of  commissure  at  an  angle  of  about  100°  but  varying  with  size 
and  distortion  of  specimen.  Posterior  surface  convex,  becoming  flatter  towards  the  posterior  end  and  with  a 
plano-concave  dorsal  area  below  the  hinge-line.  Ligament,  musculature,  and  dentition  unknown.  Sculpture 
of  faint  indistinct  concentric  lines  and  faint  coarser  rugae  visible  particularly  on  the  antero-ventral  surface. 

Discussion.  Cleionychia  incognita  is  distinct  from  C.  transversa  and  C.  prisca  in  lacking  the  coarse 
rugae  seen  over  the  whole  surface  in  those  species.  It  is  less  transverse  than  C.  transversa  and  less 
oblique  than  C.  prisca  which  it  closely  resembles  in  general  form  but  which  does  not  show  the 
flattened  dorsal  area  seen  in  C.  incognita. 

Reed  (1952,  p.  75)  compared  IGS  GSM  24310  with  Hall’s  Ambonychia  amygdalina  and  placed  it 
in  Isberg’s  genus  Ambonychinia.  Pojeta  (1966,  p.  163)  has  described  amygdalina  as  belonging  to 
Ambonychiopsis  Isberg.  None  of  the  present  specimens  of  C.  incognita  shows  the  radial  ornament 
or  anterior  lobe  associated  with  Ambonychiopsis.  Ambonychinia  intermedia  Isberg,  with  which  Reed 
(1952,  p.  75)  compared  IGS  GSM  12441  does  bear  a close  resemblance  to  C.  incognita  but  Isberg’s 
figure  (1934,  pi.  9,  fig.  1)  shows  more  obtuse  angles  gamma  (c.  90°)  and  beta  (7140°)  than  seen  in 
C.  incognita.  Similarly,  there  is  a recognizable  similarity  between  IGS  GSM  24311  and  A.  volvens 
Isberg  with  which  Reed  (1952,  p.  74)  compared  it,  but  A.  volvens  appears  to  have  a small  anterior 
lobe  (Isberg  1934,  pi.  2,  fig.  4c)  not  seen  in  C.  incognita.  The  flattened  postero-dorsal  area  is  not 
so  pronounced  in  other  species  of  Cleionychia  except  perhaps  C.  intermedia.  The  use  of  the  generic 
name  Cleionychia  rather  than  Ambonychinia  is  discussed  under  C.  prisca. 

Genus  ambonychiopsis  Isberg,  1934 

Type-species.  By  original  designation  of  Isberg  1934,  p.  82,  Ambonychiopsis  osmundsbergensis. 

Ambonychiopsis  suspecta  (Reed  1952) 

Plate  12,  fig.  14 

v*  1952  Vanuxemia?  suspecta  Reed,  p.  69,  pi.  3,  fig.  13. 

Type  specimen,  horizon,  and  locality.  Holotype  by  monotypy  IGS  GSM  22103,  by  original  designation  of 
Reed  1952,  p.  69;  a ?composite  mould  of  a left  valve,  the  only  known  specimen,  from  the  Bardahessiagh 
Formation,  exact  locality  uncertain,  but  south  of  Craigbardahessiagh,  Pomeroy. 

Measurements.  For  the  monotype  L 26-9  mm,  H 19-8  mm,  greatest  dimension  (oblique)  29-1  mm,  angle  alpha 
c.  35°,  angle  gamma  60°,  angle  beta  150°,  maximum  inflation  of  the  single  valve  c.  3 mm,  length  of  hinge-line 
17-9  mm. 

Description.  The  specimen  is  lacking  in  detail  but  shows  an  obliquely  ovate  form,  very  narrow  anteriorly  and 
expanding  rapidly  towards  the  posterior  end.  Height  is  0-7  of  the  length,  hinge-line  is  0-6  of  the  length  and 
is  apparently  straight.  The  greatest  dimension,  along  the  plane  of  obliquity  35°  to  the  hinge-line,  is  1-1  times 
the  length.  Angles  gamma  and  beta  and  inflation  are  given  above.  Anterior  to  the  otherwise  terminal  umbo 
is  a small  anterior  lobe.  The  posterior  surface  is  gently  convex.  The  steep  antero-ventral  slope  meets  the  plane 
of  commissure  at  an  angle  of  about  60°.  Ligament  and  musculature  unknown.  Dentition  represented  only  by 
two  doubtful  posterior  lateral  grooves  below  and  anterior  to  the  postero-dorsal  angle  of  the  margin.  Sculpture 
is  poorly  preserved  and  shows  only  irregular  concentric  striae. 

Discussion.  Portlock  labelled  this  specimen  ‘ Mytilus?—not  specifically  determined' . Reed  described 
two  to  three  short  cardinal  teeth  anterior  to  the  umbo  but  these  are  not  evident.  The  specimen 
has  been  developed  a little  since  Reed  examined  it  and  part  of  the  small  anterior  lobe  now  revealed 


78 


PALAEONTOLOGY,  VOLUME  25 


may  have  led  Reed  to  his  conclusions  and  suggested  to  him  his  determination  Vanuxemia?  suspecta. 
His  use  of  a query  (?)  and  the  name  suspecta  both  suggest  that  Reed  remained  doubtful  as  to  the 
nature  of  this  specimen.  Some  doubt  is  retained  in  applying  the  generic  name  Ambonychiopsis  to 
suspecta  but  this  assignment  is  justified  by  the  ambonychiid  nature  of  the  form  and  sculpture,  and 
the  presence  of  the  anterior  lobe.  However,  there  is  a lack  of  the  radial  sculpture  characteristic  of 
Ambonychiopsis. 

Ambonychinia  balclatchiensis  Reed  (1944,  p.  213,  pi.  2,  fig.  2)  from  the  Balclatchie  Beds  is 
probably  an  Ambonychiopsis : the  illustration  shows  an  anterior  lobe,  oriented  such  that  it  appears 
to  be  dorsal.  Angles  alpha  (c.  30°),  beta  (c.  150°),  and  gamma  (c.  70°)  in  Reed’s  figure  are  close 
to  those  of  A.  suspecta  and  its  greatest  dimension  is  103  of  its  length  (1-1  in  A.  suspecta ),  but  the 
anterior  lobe  in  A.  balclatchiensis  is  larger. 

The  preservation  of  the  specimen  is  such  that  comparison  with  other  species  presents  great 
difficulty.  It  is  here  proposed  that  the  name  be  restricted  to  the  type  specimen  and  the  species  be 
regarded  as  otherwise  unrecognizable. 

pterineid?  gen.  and  sp.  indet. 

Plate  11,  fig.  2 

A single,  very  small  specimen  (IGS  NIL  8981)  from  the  Killey  Bridge  Formation  of  the 
Crocknagargan  Stream  section  (Irish  grid  ref.  H721737),  is  probably  a pterineid.  Although  poorly 
preserved,  it  can  be  compared  in  outline  and  size  to  Palaeopteria  parvula  Whiteaves  (1897,  p.  181, 
pi.  20,  figs.  1-3),  from  rocks  of  Black  River-Trenton  age  (Caradoc)  in  the  area  of  Lake  Winnipeg. 
Pterinea  reticulata  Hind  (1910,  p.  494,  pi.  1,  figs.  7-8)  from  the  Drummuck  Group  of  Girvan  has 
a distinctive  ornament  and  is  larger  than  the  Irish  specimen. 

bivalve?  gen.  and  sp.  indet. 

Plate  13,  fig.  7 

Material,  horizon,  and  locality.  A single  specimen,  IGS  GSM  24308,  from  the  Bardahessiagh  Formation,  exact 
locality  unknown,  south  of  Craigbardahessiagh,  Pomeroy. 

Description.  A subtriangular  fragment  measuring  40  0 by  30-7  mm  and  showing  19+  coarse  concentric  rugae. 

Discussion.  This  fragmentary  specimen  has  in  the  past  been  referred  plausibly  to  Ambonychia  undata 
Hall,  and  indeed  the  coarse  ornament  on  the  specimen  bears  some  resemblance  to  Hall’s  figure 
(1847,  pi.  36,  fig.  la).  It  is  considered,  on  the  ornament  alone,  to  resemble  Cleionychia,  but  this 
view  gives  rise  to  difficulties  of  interpretation  and  a gastropod  or  cephalopod  affinity  is  more 
probable.  No  comparable  material  is  known  from  Pomeroy. 

Subclass  isofilibranchia  Iredale,  1939 
Order  mytiloida  Ferussac,  1822 
Superfamily  mytilacea  Rafinesque,  1815 
Family  modiolopsidae  Fischer,  1887 
(after  Pojeta  and  Gilbert-Tomlinson,  1978) 

Genus  modiolopsis  Hall,  1847 

Type  species.  Pterinea  modiolaris  Conrad,  1838,  by  original  designation  of  Hall  1847,  p.  157. 

Modiolopsis  sp. 

Plate  13,  fig.  11 

Material,  horizon,  and  locality.  A single,  fragmentary,  small  internal  mould  of  a right  valve,  IGS  GU  1380, 
from  the  Killey  Bridge  Formation  at  the  Crocknagargan  Stream  section  (locality  1 of  Mitchell  1977),  Pomeroy, 
Co.  Tyrone. 


TUNNICLIFF:  LATE  ORDOVICIAN  BIVALVES 


79 


Description.  Elongate,  obliquely  ovate,  modioliform,  with  the  umbo  at  about  the  anterior  one-tenth  to  one-fifth. 
Height  (12-5+  mm),  about  half  of  length  (26-0+  mm).  Inflation  of  the  single  valve  2-5  mm.  The  incomplete 
anterior  end  is  narrow  but  the  shell  broadens  rapidly  towards  the  flat  posterior  which  is  also  incomplete.  The 
ventral  slope  and  margin  show  a slight  sinus  a little  posterior  to  the  umbo.  A rounded  umbonal  ridge  runs 
towards  the  postero-ventral  angle  which  is  lost.  An  adductor  muscle  scar  occupies  much  of  the  preumbonal 
portion  of  the  shell  and  two  umbonal  muscle  scars  are  present.  Edentulous  but  with  a dorsal  longitudinal 
groove.  Sculpture  unknown  except  for  a suggestion  of  concentric  ornament  on  the  antero-ventral  slope. 

Discussion.  Although  fragmentary,  this  specimen  is  similar  in  shape  to  two  species  figured  by  Hind 
(1910):  Modiolopsis  exasperatus  (Phillips)  (Hind  1910,  pi.  2,  fig.  15)  and  M.  scotica  Hind  (1910, 
pi.  2,  figs.  18-20)  both  recorded  from  the  Drummuck  Group  (Ashgill)  of  Girvan.  In  both  the 
Scottish  species  the  valves  become  higher  very  rapidly  towards  the  posterior  as  in  the  Irish  specimen, 
but  although  the  latter  is  incomplete,  its  postero-dorsal  margin  suggests  that  it  is  less  elongate. 

Genus  corallidomus  Whitfield,  1895 

Type  species.  C.  concentrica  (Hall  and  Whitfield  1875),  by  monotypy:  the  only  Ordovician  species  assigned 
to  the  genus  (Pojeta  1971,  p.  6). 

Corallidomus  concentrica  (Hall  and  Whitfield  1875) 

Plate  12,  fig.  9 

* 1875  Modiolopsis  concentrica  Hall  and  Whitfield,  pp.  86-87,  pi.  2,  fig.  18. 

1893  [1895]  Corallidomus  concentricus  Whitfield,  pp.  492-493,  pi.  13,  fig.  2 [on  p.  493],  pi.  13,  [as 
n.  gen.,  n.  sp.]. 

1978  Corallidomus  concentrica  (Hall  and  Whitfield);  Pojeta,  pi.  12,  figs.  12-14. 

Material,  horizon,  and  locality.  A right  valve,  UM  1920-843,  from  the  Killey  Bridge  Formation  of  Pomeroy, 
Co.  Tyrone,  exact  locality  uncertain. 

Description.  A small  modioliform  shell,  nearly  twice  as  long  as  high  (12  mm  high,  22  mm  long)  and  inflated 
to  nearly  2 mm  at  one  point  along  the  umbonal  ridge.  The  shell  broadens  towards  the  posterior.  The  anterior 
margin  is  rounded,  the  posterior  end  is  obliquely  truncate.  The  ventral  margin  shows  a slight  flexure  at  about 
the  mid-point.  The  umbo  lies  at  about  the  anterior  one-quarter.  The  preumbonal  part  of  the  shell  is  occupied 
by  an  adductor  muscle  scar.  Dentition  uncertain  but  there  is  a suggestion  of  a short,  thin,  posterior  tooth 
immediately  behind  the  umbo.  The  ornament  is  distinctive:  the  antero-ventral  portion  of  the  shell  bears  fine 
concentric  lines  which  give  way  at  the  umbonal  ridge  to  strong,  coarse,  fairly  regular  concentric  rugae.  The 
umbonal  ridge  becomes  less  marked  away  from  the  umbo  but  is  defined  by  the  change  in  texture  and  direction 
of  the  ornament. 

Discussion.  The  specimen  closely  matches  the  descriptions  and  illustrations  of  Hall  and  Whitfield 
(1875,  pp.  86-87,  pi.  2,  fig.  18)  and  the  specimens  figured  by  Pojeta  (1978,  pi.  12,  figs.  12-14). 
Pojeta  (1978,  pi.  1 2,  figs.  12-14)  figured  Modiolopsis  concentrica  as  Corallidomus  concentrica,  treating 
it  as  synonymous  with  Whitfield’s  type  species,  and  is  followed  here;  however,  Whitfield’s  (1893) 
figures  and  the  lack  of  his  material  (Pojeta  1971,  p.  30)  do  give  rise  to  some  doubt  as  to  whether 
the  two  species  are  identical  and  therefore  whether  M.  concentrica  should  be  assigned  to 
Corallidomus.  Pojeta  (1971,  pp.  30-33)  observed  that  the  known  occurrence  of  Corallidomus  is 
restricted  to  the  Richmondian,  comparable  to  the  Cautleyan  age  of  the  Killey  Bridge  Formation. 
The  mode  of  life  of  Corallidomus,  recorded  by  Whitfield  as  boring  into  masses  of  the  coral  Labechia 
ohioensis,  is  said  by  Pojeta  (1971)  to  be  unique  among  Ordovician  bivalves  and,  although  the 
present  specimen  is  not  in  life  position,  it  is  worth  noting  that  masses  of  the  tabulate  coral  Catenipora 
tapaensis  (Sokolov)  [identified  by  Dr.  D.  E.  White,  IGS]  were  collected  by  Portlock  (IGS  GSM 
103460, 104183,  104 184)  from  the  Killey  Bridge  Formation,  possibly  from  locality  3 of  Mitchell  1977. 

A left  valve,  NMI  G.4.  1979,  in  the  Griffith  Collection  is  from  the  same  horizon  as  UM  1920 
843  but  also  lacks  exact  locality  details  and  is  here  recorded  as  Corallidomus ? sp.  (PI.  12,  fig.  10). 
It  lacks  the  characteristic  coarser  ornament  on  the  posterior  portion  of  the  shell  and  has  a more 
central  umbo  than  C.  concentrica. 


PALAEONTOLOGY,  VOLUME  25 


Subclass  palaeoheterodonta  Newell,  1965 
Order  modiomorphoidea  Newell,  1969 
Superfamily  modiomorphacea  Miller,  1877 
Family  modiomorphidae  Miller,  1877 
Genus  goniophora  Phillips,  1848 

Type  species.  Goniophora  cymbaeformis  (J.  de  C.  Sowerby)  by  original  designation  of  Phillips  1848,  p.  264. 


Goniophora  sp. 

Plate  12,  fig.  15 

Material,  locality,  and  horizon.  A single  specimen,  IGS  Zs  2751,  external  cast  of  a right  valve  collected  by 
Mr.  R.  P.  Tripp  from  the  Killey  Bridge  Formation,  at  locality  3 of  Mitchell  (1977),  the  Little  River  Section 
(Irish  grid  ref.  H7297  7268). 

Measurements.  H 6 0 mm,  L 13-8  mm,  AL  2-8  mm;  inflation  of  the  single  valve  1-8  mm;  obliquity,  measured 
along  umbonal  ridge,  35-40°. 

Description.  Small  Goniophora',  transversely  elongate  with  a pronounced,  slightly  sigmoidal  umbonal  ridge  or 
carina  running  from  the  umbo  to  the  postero-ventral  angle;  height  to  length  ratio  0-43;  umbo  barely  protruding 
above  the  hinge-line  and  situated  at  about  the  anterior  one-fifth.  Dentition  and  musculature  unknown.  The 
surface  anterior  to  the  carina  shows  concentric  striae  of  irregular  strength  and  spacing  but  the  posterior 
surface  has  a well-developed  concentric  sculpture,  regularly  spaced,  with  3-4  striae  per  mm.  The  posterior 
surface  also  shows  a flexure  close  behind  the  carina. 

Discussion.  This  remains  the  only  known  Goniophora  specimen  from  the  Ordovician  of  Pomeroy: 
Goniophora  brycei  (Portlock)  of  Reed  (1952,  p.  78)  is  a Cyrtodontal  (see  Cyrtodontal  spp.).  In  the 
absence  of  any  detail  of  the  hinge,  there  is  some  doubt  as  to  whether  Goniophora  or  Goniophorina 
Isberg  is  the  more  appropriate  genus  for  this  specimen;  but  it  is  reasonable  to  place  it  in  the  more 
familiar  Goniophora.  The  lack  of  radial  costae  precludes  it  from  being  Cosmogoniophora  McLearn. 
Specimens  of  a comparable  form  from  Caradocian  rocks  in  north  Wales  are  under  study. 


Genus  colpomya  Ulrich,  1894 

Type  species.  Monotype,  Colpomya  constricta  Ulrich  by  original  designation  of  Ulrich  1894,  pp.  522-523. 

Colpomya  simplex  (Portlock  1 843) 

Plate  13,  fig.  14 

v*  1843  Cypricardia?  simplex  Portlock,  p.  426. 

v pars.  1910  Grammysia  undata  J.  de  C.  Sowerby;  Hind,  p.  540,  pi.  5,  fig.  14,  Inon  figs.  13,  15,  16. 
v.  1952  Cuneamya  simplex  (Portlock);  Reed,  p.  79,  pi.  4,  fig.  10. 

Type-specimen.  The  only  known  syntype,  a left  valve,  IGS  GSM  24290,  from  the  Killey  Bridge  Formation, 
is  here  selected  as  lectotype;  locality  given  by  Portlock  as  his  ‘sheet  37  no.  2’,  probably  locality  3 of  Mitchell 
(1977),  the  Little  River  section,  Pomeroy  (Irish  grid  ref.  H7297  7268). 

Measurements.  H 22-9  mm,  L 39-5  mm,  AL  14  00  mm,  inflation  of  the  single  valve  is  7-8  mm;  obliquity, 
measured  along  the  umbonal  ridge,  35°. 

Description.  Transversely  elongate,  rounded  at  posterior  and  anterior  ends,  with  dorsal  and  ventral  margins 
subparallel;  posterior  broader  than  anterior.  The  height  is  about  0-6  of  the  length  and  the  umbo  is  at  about  the 
anterior  one-third.  The  single  valve  is  inflated  to  7-8  mm  in  the  type-specimen.  A slightly  sigmoidal  rounded 
umbonal  ridge  at  an  angle  of  35°  to  the  hinge-line  meets  the  postero-ventral  margin  below  the  posterior  end 
of  the  hinge-line.  The  prosogyrate  umbo  extends  above  the  hinge-line  and  is  flattened,  the  flat  area  corresponding 
to  the  very  faint  sulcus  developed  towards  the  margin  anterior  to  the  umbonal  ridge.  Lunule  present  but 


TUNNICLIFF:  LATE  ORDOVICIAN  BIVALVES 


81 


escutcheon  not  clearly  differentiated.  Dentition  and  musculature  unknown.  Sculpture  of  coarse,  rather  irregular 
concentric  rugae,  pronounced  anterior  to  the  sulcus  and  becoming  fainter  and  even  indistinguishable  on  the 
posterior  surface. 

Discussion.  Portlock  (1843,  p.  426)  and  Hind  (MS,  see  Reed  1952,  p.  79)  compared  this  species  to 
Cypricardia  impressa  J.  de  C.  Sowerby  (in  Murchison).  As  Portlock  observed,  the  sulcus  in  C. 
simplex  is  much  less  pronounced  than  in  C.  impressa.  Colpomya  simplex  differs  from  Cuneamya 
in  that  the  umbo  is  not  terminal  (cf.  Cuneamya  miamiensis  Hall  and  Whitfield  as  figured  in  the 
Treatise,  p.  N821  and  Pojeta  1971,  pi.  15,  figs.  9,  10),  but  it  is  closely  comparable  to  the  type 
species  of  Colpomya,  C.  constricta  Ulrich  (1894,  p.  523,  fig.  41)  from  the  upper  Trenton  of  Kentucky. 
The  specimen  figured  by  Hind  (1910,  pi.  5,  fig.  14)  from  the  Drummuck  Group  (BM  L49889)  may 
be  conspecific  with  C.  simplex. 

Genus  semicorallidomus  Isberg,  1934 

Type  species.  Semicorallidomus  whitfieldi  Isberg  by  original  designation  of  Isberg  1934,  pp.  175,  180  from  the 
Ordovician  of  Sweden. 


Semicorallidomus?  sp. 

Plate  11,  figs.  8,  11,  15 

v.  1843  Modiola  Nerei  (Munster);  Portlock,  p.  424,  pi.  33,  fig.  10. 
v.  1843  Mytilus?  Nerei  (Munster);  Portlock,  p.  424. 

1878  Modiolopsis  Nerei  Munster;  Baily,  p.  28. 

? 1952  Modiolodon  speciosus  (McCoy);  Reed,  p.  72,  pi.  4,  fig.  1. 
v.  1952  Paramodiola?  sp.;  Reed,  p.  73,  pi.  4,  fig.  3. 

Material.  IGS  (Portlock  and  Mitchell  collections)  GSM  12450-12452,  GU  1739,  1741;  UM  (Portlock  and 
Grainger  collections)  4125  and  counterpart,  K4169,  4195,  4197-4198,  4242;  NMI  (Griffith  collection);  TCD 
7871  (Portlock  collection);  ?SM  A 16461. 

Localities  and  horizon.  All  from  the  Bardahessiagh  Formation,  south  of  Craigbardahessiagh,  Pomeroy,  exact 
[ localities  uncertain. 


Measurements 


L 

H 

H/L 

AL/L 

Max. 

20  0 mm 

140  mm 

0-85 

015 

Min. 

100  mm 

7-2  mm 

0-62 

0-07 

Mean 

14-63  mm 

10-4  mm 

0-71 

Oil 

Median 

150  mm 

1 0-6  mm 

0-74 

Oil 

Description.  Subovate,  modiomorphoid,  with  a height-length  ratio  about  0-7.  Umbo  not  prominent  and  at 
about  the  anterior  one-tenth  to  one-fifth.  Some  specimens  show  a slight  flexure  of  the  posterior  surface. 
Obliquity  about  30°.  Apparently  isomyarian  with  the  anterior  adductor  muscle  scar  almost  below  the  umbo 
and  the  posterior  scar  poorly  defined.  Single  valve  inflated  to  about  2-5  mm  in  the  larger  specimens.  Concentric 
ornament  of  varying  strength,  with  an  average  of  five  striae  per  mm  on  the  most  inflated  part  of  the  shell 
along  the  plane  of  obliquity.  Precise  nature  of  the  hinge  unknown,  but  the  left  valve  apparently  bears  a tooth 
below  the  umbo  (IGS  GU  1739,  PI.  11,  fig.  8). 

Discussion.  The  name  Paramodiola,  used  by  Reed  (1952)  for  this  form,  is  rejected  here  in  view  of 
the  apparent  presence  of  a tooth  in  the  left  valve  (GU  1739);  Paramodiola  is  said  to  be  edentulous. 
A diagnostic  characteristic  of  Semicorallidomus  is  the  presence  in  the  left  valve  of  a socket,  the 
reverse  of  what  is  seen  in  the  Pomeroy  form.  In  outline,  the  Irish  specimens  most  closely  resemble 
the  type  species  Semicorallidomus  whitfieldi  Isberg.  The  specimen  described  and  figured  by  Reed 
(1952,  p.  72,  pi.  4,  fig.  1)  as  Modiolodon  speciosus  (McCoy)  (SM  A16461)  was  not  available  for 
study  at  the  time  of  writing,  but  his  figure  resembles  the  known  specimens  of  Semicorallidomus ?: 
McCoy’s  specimen  of  M.  speciosus  is  in  the  NMI  (see  Cycloconcha?  speciosa). 


82 


PALAEONTOLOGY,  VOLUME  25 


Subclass  actinodontia  Douville,  1912  (after  Pojeta  1978) 

Family  cycloconchidae  Ulrich,  1884 
Genus  cycloconcha  Miller,  1874 

Type  species.  By  original  designation  Cycloconcha  mediocardinalis  Miller,  1 874,  p.  231. 

Cycloconcha?  speciosa  (McCoy  1 846) 

Plate  13,  figs.  5,  6,  9,  10,  12,  13 

v.  1843  Avicula  obliqua  J.  de  C.  Sowerby  in  Murchison;  Portlock,  p.  425. 

v.  1843  Avicula  orbicularis  J.  de  C.  Sowerby  in  Murchison;  Portlock,  pp.  425,  755  (Synoptical  Table) 
[pars:  see  Cyrtodonta?  sp.]. 

v*  1846  Pullastra  speciosa  McCoy,  p.  17,  pi.  2,  fig.  2 (reversed). 

Inon  1952  Modiolodon  speciosus  (McCoy);  Reed,  p.  72,  pi.  4,  fig.  1. 

Type  specimen.  The  original  of  McCoy’s  figure  (1846,  pi.  2,  fig.  2)  is  in  the  NMI  Griffith  Collection  and  as 
the  only  known  syntype  is  here  selected  as  Lectotype,  NMI  G.5.  1979. 

Material.  Five  right  valves;  GSM  21919,  21920,  104236,  UM  K4213,  4214.  Four  left  valves;  GSM  104235, 
UM  K4212,  4216  and  the  type  specimen.  All  are  composite  moulds. 

Horizon  and  locality.  All  from  the  Bardahessiagh  Formation,  exact  localities  uncertain  but  south  of 
Craigbardahessiagh,  Pomeroy. 


Measurements 

L 

H 

H/L 

AL/L 

Obliquity 

Max. 

24-2  mm 

22-0  mm 

0-94 

0-38 

55° 

Min. 

16  0 mm 

13-0  mm 

0-72 

0-29 

45° 

Mean 

22-4  mm 

18-6  mm 

0-83 

0-34 

50-6° 

Median 

20- 1 mm 

17-5  mm 

0-83 

0-34 

50° 

explanation  of  plate  13 

Fig.  1.  Cyrtodonta?  sp.  ?composite  mould  of  right  valve,  UM  K4194,  Bardahessiagh  Formation  (Caradoc), 
south  of  Craigbardahessiagh,  Pomeroy,  x IF 

Figs.  2,  3.  Ambonychia  arundinea  sp.  nov.  2,  external  cast  of  right  valve,  IGS  GSM  24302.  3,  external  cast 

of  left  valve  IGS  GSM  24303.  Both  Killey  Bridge  Formation  (Ashgill,  Cautleyan),  Pomeroy,  exact  locality 
uncertain,  x IF 

Figs.  4,  8.  Vanuxemia?  contorta  (Portlock  1843).  4,  external  cast  of  left  valve,  holotype,  IGS  GSM  12435. 
8,  external  cast  of  right  valve,  IGS  Zf  1020.  Both  Killey  Bridge  Formation  (Ashgill,  Cautleyan),  Pomeroy, 
exact  locality  uncertain,  x 1 . 

Figs.  5,  6,  9,  10,  12,  13.  Cycloconcha?  speciosa  (McCoy  1846).  5,  left  valve,  lectotype  NMI  G.5.  1979. 

6,  right  valve,  UM  K42 14.'  9,  left  valve,  UM  K4212.  10,  right  valve,  IGS  GSM  21920.  12,  right  valve, 

IGS  GSM  21919.  13,  left  valve,  UM  K4216.  All  preserved  as  composite  moulds,  Bardahessiagh  Formation 

(Caradoc),  south  of  Craigbardahessiagh,  Pomeroy,  x IF 

Fig.  7.  Bivalve?  gen.  and  sp.  indet.  ?external  cast  IGS  GSM  24308,  Bardahessiagh  Formation  (Caradoc), 
south  of  Craigbardahessiagh,  x 1 . 

Fig.  11.  Modiolopsis  sp.  internal  mould  of  right  valve,  IGS  GU  1380,  Killey  Bridge  Formation  (Ashgill, 
Cautleyan),  Crocknagargan  Stream  section,  Pomeroy  (IGR  H721737),  x 2\. 

Fig.  14.  Colpomya  simplex  (Portlock  1843),  external  cast  of  left  valve,  lectotype,  IGS  GSM  24290,  Killey 
Bridge  Formation  (Ashgill,  Cautleyan),  Pomeroy,  exact  locality  uncertain,  x IF 

Figs.  1 5 19.  Lyrodesma  radiatum  (Portlock  1843).  15,  17,  left  valve,  internal  mould,  IGS  GU  3264;  15, 

oblique  dorsal  view  to  show  dentition,  lit  from  bottom  left,  x6;  17,  lateral  view,  lit  from  bottom  left, 
x4.  16,  external  cast  of  left  valve,  lectotype,  IGS  GSM  22178,  x2.  18,  external  cast  of  right  valve,  IGS 

GSM  22177,  x2.  19,  composite  mould  of  conjoined  valves,  IGS  GSM  104185,  x2.  All  Bardahessiagh 

Formation  (Caradoc),  south  of  Craigbardahessiagh,  Pomeroy. 


PLATE  13 


TUNNICLIFF,  Ordovician  bivalves 


PALAEONTOLOGY,  VOLUME  25 


Description.  Cycloconcha?  of  slightly  variable  subovate  shape,  the  posterior  end  being  somewhat  truncate, 
with  height  between  0-72  and  0-94  of  the  length.  Slight  umbonal  ridge  is  at  about  50°  to  the  hinge-line.  The 
anterior  dorsal  margin  is  at  an  angle  of  145  to  155°  to  the  posterior  dorsal  margin.  The  prosogyrate  umbones 
are  at  about  the  anterior  one-third.  Maximum  inflation  of  a single  valve  seen  is  2 mm  in  a valve  24-2  mm 
long.  Hinge-line  shows  two  or  three  lamellar  posterior  lateral  teeth  in  the  left  valve  and  three  in  the  right 
valve,  subparallel  to  the  margin,  and  at  least  one  anterior  lateral  tooth  in  each.  Anterior  adductor  muscle 
scar  faint,  about  half-way  between  umbo  and  anterior;  posterior  musculature  unknown.  Sculpture  of  regular 
concentric  striae  with  more  prominent  lines  at  intervals  of  about  2-4  mm  along  the  umbonal  ridge  especially 
towards  the  margin. 

Discussion.  This  species  is  placed,  with  some  reservation,  in  Cycloconcha  since  it  possesses  straight 
anterior  and  posterior  teeth,  although  no  cardinal  teeth  are  visible  in  the  available  specimens.  I 
have  not  seen  Reed’s  specimen  of  Modiolodon  speciosus  (SM  A16461,  1952,  p.  72,  pi.  4,  fig.  1), 
and  doubt  whether  it  should  be  placed  in  C.?  speciosa.  Reed  noted  radial  striation,  stating  that 
this  appeared  to  be  present  in  McCoy’s  specimen  but  he  was  working  only  from  McCoy’s  figure 
and  no  radial  striation  is  apparent  in  the  type  specimen.  He  also  mentioned  a deeply  impressed 
anterior  muscle  scar  but  in  the  specimens  of  C.?  speciosa  available  the  anterior  muscle  scar  can  at 
best  be  described  as  faint.  His  comparison  of  his  specimen  with  Modiolopsis?  consimilis  Ulrich, 
Modiolodon  obtusus  Ulrich,  and  M.  truncatus  (Hall)  all  suggest  that  we  should  regard  his  specimen 
as  distinct  from  C.  speciosa. 

In  shape,  C.?  speciosa  resembles  Cyrtodonta  parva  Ulrich  (1894,  p.  541,  pi.  39,  figs.  24,  25)  from 
the  Trentonian  of  Minnesota  but  it  is  larger  and  the  posterior  teeth  are  longer.  In  this  last  respect, 
comparison  may  be  made  with  the  upper  Trenton  shale  form  Cypricardites  tenellus  Ulrich  (1892, 
p.  237)  from  Minnesota.  The  teeth  in  C.  tenellus,  later  referred  to  Cyrtodonta  by  Ulrich  (1894,  p. 
546,  pi.  40,  figs.  15-19),  are  similar  in  form  to  those  of  Cycloconcha  speciosa,  but  only  two  are 
recorded  in  the  right  valve. 

Family  lyrodesmatidae  Ulrich,  1894 
Genus  lyrodesma  Conrad,  1841,  p.  51 

Type-species.  By  monotypy  Lyrodesma  planum  Conrad,  1841,  p.  51. 

Lyrodesma  radiatum  (Portlock  1843) 

Plate  13,  figs.  15-19 

v*  1843  Nucula?  radiata  Portlock,  p.  430,  pi.  36,  fig.  11. 
v.  1846  Nucula  radiata  Portlock;  McCoy,  p.  19. 

1878  Ctenodonta  radiata  Portlock;  Baily,  p.  28. 
v.  1952  Lyrodesma  radiatum  (Portlock);  Reed,  p.  67,  pi.  3,  fig.  11. 

Type-specimen.  Lectotype  selected  here  IGS  GSM  22178,  the  specimen  figured  by  Portlock  (1843,  pi.  36,  fig. 
11)  and  used  by  Reed  (1952,  p.  67)  in  his  redescription. 

Material,  localities,  and  horizon.  Specimens  in  IGS,  TCD,  and  NMI,  including  six  Portlock  syntypes.  Mostly 
external  moulds,  or  composite  moulds.  All  from  the  Bardahessiagh  Formation,  south  of  Craigbardahessiagh, 
Pomeroy,  exact  localities  uncertain  except  IGS  GU  3264,  Mitchell  Collection,  from  Mitchell’s  Bardahessiagh 
collecting  area  (Mitchell  1977,  p.  5). 

Measurements  Angle  between  antero- 

and  postero-dorsal 


L 

H 

H/L 

AL/L 

margins 

Max. 

20-0  mm 

12-6  mm 

0-69 

0-40 

c.  150° 

Min. 

13-0  mm 

8 0 mm 

0-50 

0-26 

c.  125° 

compressed 

dorso-ventrally 

Mean 

16-24  mm 

9-86  mm 

0-60 

0-34 

Median 

16-5  mm 

1 1 -0  mm 

0-60 

0-33 

TUNNICLIFF:  LATE  ORDOVICIAN  BIVALVES 


85 


Description.  Lyrodesma  of  elongate,  ovate  form  with  height-length  ratio  about  0-6.  Maximum  inflation  of  a 
single  valve  is  about  2 mm  in  a valve  20  mm  long.  The  umbo  lies  at  about  the  anterior  one-third.  The  anterior 
margin  is  rounded,  the  posterior  end  is  subtruncate  and  obliquely  carinate.  Postero-dorsal  surface  bearing 
17+  radial  striations  of  variable  strength  and  arranged  somewhat  irregularly.  Anterior  portions  of  the  surface 
show  very  faint  concentric  growth  lines.  Umbones  rather  small.  Six  radiating  crenulate  teeth  of  similar  size 
below  umbo  in  left  valve.  A faint  line  or  ridge  is  seen  on  the  internal  mould  running  from  the  umbo  to  the 
ventral  margin  anterior  to  the  umbonal  ridge.  Posterior  and  anterior  adductor  muscle  scars  situated  dorsally, 
the  posterior  scar  being  a little  larger  than  the  anterior.  Accessory  muscle  scars  poorly  discerned  in  the  present 
specimen. 

Discussion.  As  Reed  noted  (1952,  pp.  67-68)  Portlock’s  description  and  figure  of  Lyrodesma  radiatum 
were  poor,  but  Reed  saw  only  two  of  the  Portlock  syntypes  and  had  no  specimens  showing  the 
dentition.  In  shape  and  the  number  of  teeth,  L.  radiatum  resembles  L.  majus  (Ulrich)  and  Reed 
likened  it  to  L.  cincinnatiense  Hall  (of  Ruedemann  1926)  and  to  L.  poststriatum  elongatum  Stewart, 
1920,  but  until  more  material  is  available  to  provide  further  detail  of  the  internal  structure,  no 
close  comparison  can  be  made  with  other  species. 


Class  rostroconchia  Pojeta,  Runnegar,  Morris  and  Newall,  1972 
Order  conocardioidea  Neumayr,  1891 
Superfamily  conocardiacea  Miller,  1889 
Family  hippocardiidae  Pojeta  and  Runnegar,  1976 
Genus  hippocardia  Brown,  1843 

Type  species.  By  monotypy  Cardium  hibernicum  Sowerby,  1815.  The  classification  used  here  is  that  of  Pojeta 
and  Runnegar  1976. 


Hippocardia  praepristis  (Reed),  1952 
Plate  9,  fig.  8 

v*  1952  Conocardium  praepristis  Reed,  p.  80,  pi.  4,  fig.  11. 

v.  1976  Hippocardia  praepristis  (Reed);  Pojeta  and  Runnegar,  p.  76. 

Type  specimens.  Holotype  by  original  designation,  IGS  GSM  24147,  the  only  specimen  known  to  Reed  and 
to  the  present  author.  The  specimen  is  part  of  the  Portlock  Collection  (Tunnicliff  1980)  and  bears  a label 
reading  ‘Cypricardia— Analogous  to  C.  cymbaeformis  but  differs  from  it  and  also  from  Cardium  carpo- 
morphum  (J.  P.)’  but  Portlock  (1843)  published  no  reference  to  it. 

Locality  and  horizon.  The  horizon  of  this  specimen  is  uncertain,  but  Reed  likened  the  matrix  to  that  of  the 
Tirnaskea  Beds  (Tirnaskea  Formation).  If  this  is  correct,  a likely  locality  would  be  the  Tirnaskea  Stream 
section  or  perhaps  the  Crocknagargan  Stream  section,  both  of  which  have  exposures  of  Tirnaskea  Formation 
(Ashgill,  Hirnantian)  but  there  is  no  record  of  Portlock  having  specimens  from  the  latter  and  there  is 
evidence  that  he  had  none  from  the  Tirnaskea  Stream  section  (Portlock  1843,  p.  230  . .on  the  Tirnaskea 

or  small  river  there  are  thin  calcareous  layers  of  three  or  four  inches  thick  mixed  with  quartzose  bands, 
no  fossils,  however,  occurring  in  them . . .’).  Lithologically,  the  specimen  resembles  known  Killey  Bridge 
Formation  specimens,  and  it  must  be  assumed  that  the  specimen  is  from  that  horizon,  but  the  locality  is 
unknown. 

Description.  Adequate  description  is  provided  by  Reed  (1952),  and  Pojeta  and  Runnegar  (1976) 
placed  praepristis  in  Hippocardia  on  this  basis.  The  specimen  is  figured  here  to  supplement  Pojeta 
and  Runnegar. 


Acknowledgements.  Specimens  were  made  available  through  the  kindness  of  the  following:  Dr.  N.  J.  Morris, 
Mr.  R.  J.  Cleevely,  and  Mr.  R.  P.  Tripp  (Brit.  Mus.  Nat.  Hist.),  Mr.  P.  Doughty  and  Mr.  K.  James  (Ulster 
Museum,  Belfast),  Dr.  C.  O’Riordan  (Nat.  Mus.  Ireland,  Dublin),  Miss  V.  Burns  (Trinity  College,  Dublin), 
and  Mr.  M.  Dorling  (Sedgwick  Museum,  Cambridge).  Dr.  J.  Kriz  helped  with  Barrande  type  specimens 
and  discussion.  Dr.  Morris  and  Drs.  D.  E.  Butler,  A.  W.  A.  Rushton,  and  D.  E.  White  (IGS)  also  discussed 


PALAEONTOLOGY,  VOLUME  25 


aspects  of  the  work  and  gave  advice.  Drs.  Butler,  Rushton,  and  White  read  the  manuscript  critically  at  various 
stages  of  preparation.  Dr.  J.  Pojeta  and  Professor  P.  Bretsky  kindly  provided  copies  of  papers.  Text-fig.  1 is 
reproduced  by  permission  of  the  Trustees,  Ulster  Museum.  This  paper  is  published  with  the  permission  of  the 
Director,  Institute  of  Geological  Sciences. 


REFERENCES 

allen,  j.  a.  1978.  Evolution  of  deep-sea  protobranch  bivalves.  Phil.  Trans.  R.  Soc.  Lond.  B.  284,  387-401. 

— and  Sanders,  h.  l.  1973.  Studies  on  deep-sea  protobranchia  (Bivalvia);  the  families  Siliculidae  and 
Lametilidae.  Bull.  Mus.  comp.  Zool.  Harv.  145  (6),  263-310. 

babin,  c.  1966.  Mollusques  bivalves  et  cephalopodes  du  Paleozoique  armoricain.  472  pp„  18  pis.  Brest. 

— and  melou,  m.  1972.  Mollusques  bivalves  et  brachiopodes  des  ‘Schistes  de  Raguenez’  (Ordovicien 
superieur  du  Finistere);  consequences  stratigraphiques  et  paleobiogeographiques.  Ann.  Soc.  geol.  N.  92  (2), 
79-94,  pis.  7-10. 

— and  robardet,  M.  1973  (February).  Quelques  palaeotaxodontes  (Mollusques  bivalves)  de  l’Ordovicien 
Superieur  de  Saint-Nicolas  de  Pierrepont  (Normandie).  Bull.  Soc.  geol.  miner.  Bretagne,  1972,  (C),  4 (1), 
25-38,  pis.  1-3. 

baily,  w.  H.  1875.  Figures  of  characteristic  British  fossils,  1,  Palaeozoic,  lxxx.  126  pp.,  42  pis.  London. 

— 1878.  In  Nolan,  J.,  Explanatory  memoire  to  accompany  sheet  34  of  the  maps  of  the  Geol.  Surv.  of  Ireland. 
Mem.  geol.  Surv.  G.B.  Dublin  and  London. 

barrande,  j.  1881.  Systeme  silurien  de  la  Boheme,  4,  Mollusques  Acephales.  340  pp.,  361  pis.  Prague  and 
Paris. 

billings,  E.  1858.  New  genera  and  species  of  fossils  from  the  Silurian  and  Devonian  formations  of  Canada. 
Can.  nat.  and  Geol.  3,  419-444,  24  text-figs. 

blake,  j.  f.  1882.  A monograph  of  the  British  fossil  cephalopoda,  Pt.  1.  Introduction  and  Silurian  species, 
248  pp.,  31  pis. 

bradshaw,  M.  a.  1 970.  The  dentition  and  musculature  of  some  Middle  Ordovician  (Llandeilo)  bivalves  from 
Finistere,  France.  Palaeontology,  13,  623-645. 

bretsky,  p.  w.  and  bretsky,  s.  s.  1977.  Morphological  variability  and  change  in  the  Palaeotaxodont  bivalve 
mollusk  Nuculites  planulatus  (Upper  Ordovician  of  Quebec).  J.  Paleont.  51,  256-271,  1 pi. 
carter,  r.  M.  1971.  Revision  of  Arenig  Bivalvia  from  Ramsey  Island,  Pembrokeshire.  Palaeontology,  14, 
250-261,  pis.  38-39. 

conrad,  T.  A.  1838.  Report  on  the  Palaeontological  Department  of  the  Survey.  A.  Rep.  geol.  Surv.  N.Y.,  2, 
107-119. 

— 1841.  Fifth  annual  report  on  the  palaeontology  of  the  strata  of  New  York.  Ibid.  5,  25-57. 
donald,  J.  1902.  On  some  of  the  Proterozoic  gastropoda  which  have  been  referred  to  Murchisonia  and 

Pleurotomaria  Q.  Jl.  geol.  Soc.  Lond.  58,  313-339,  pis.  7-9. 

— ( = longstaff,  j.)  1924.  Descriptions  of  gastropoda  etc.  Ibid.,  80,  408-446,  pis.  31-38. 

fuchs,  A.  1919.  Beitrag  zur  Kenntnis  der  Devonfauna  der  Verse-  und  der  Hobracker  Schichten  des 
sauerlandischen  Faciesgebietes.  Jahrbuch  der  Preufiischen  geol.  Landesanstalt  (1918),  39  (1),  58-95, 
pis.  5-9. 

hall,  j.  1847.  Natural  history  of  New  York:  Palaeontology  Vol  1,  338  pp.,  87  pis. 

— and  whitfield,  r.  p.  1875.  Descriptions  of  Silurian  Fossils.  Geol.  Surv.  Ohio,  2 (2),  65-161,  pis.  1-9. 
havlicek,  v.  and  vanek,  j.  1966.  The  biostratigraphy  of  the  Ordovician  of  Bohemia.  Sbornik  geologickych 

ved,  paleontologie,  rada  P.  svazek  8,  7-69,  pis.  1-16.  With  Czech  summary.  [In  English.] 
hind,  w.  1910.  The  lamellibranchs  of  the  Silurian  rocks  of  Girvan.  Trans.  R.  Soc.  Edinb.  47  (3),  No.  18, 
479-548,  pis.  1-5. 

isberg,  o.  1934.  Studien  iiber  lamellibranchiaten  des  Leptaenakalkes  in  Dalarna.  493  pp.,  32  pis.  Lund. 
kriz,  j.  and  pojeta,  j.  1974.  Barrande’s  colonies  concept  and  a comparison  of  his  stratigraphy  with  the 
modern  stratigraphy  of  the  Middle  Bohemian  Lower  Palaeozoic  rocks  (Barrandian)  of  Czechoslovakia. 
J.  Paleont.  48  (3),  489-494,  2 text-figs. 

lamont,  a.  1946.  Lamellibranchs  from  the  Lower  Drummuck  Group  (Ashgillian),  Girvan,  Scotland.  Cem. 

Lime  Gravel , 20  (10)  (April  1946),  364-366,  1 pi.,  1 text-fig. 

LAROQUE,  a.  and  NEWELL,  n.  D.  1969.  In  MOORE,  r.  c.  1969. 

LONGSTAFF,  J.  1924.  See  DONALD,  j. 

mcalester,  A.  l.  1962.  Mode  of  preservation  in  early  Paleozoic  pelecypods  and  its  morphological  and  ecological 
significance.  J.  Paleont.  36  (1),  69-73,  pi.  16,  1 text-fig. 


TUNNICLIFF:  LATE  ORDOVICIAN  BIVALVES 


87 


mcalester,  a.  l.  1968.  Type  species  of  Paleozoic  Nuculoid  bivalve  genera.  Mem.geol.  Soc.  Am.  105,  i-ix,  143  pp., 
36  pis. 

1969.  In  moore,  r.  c.  1969. 

MCCOY,  F.  1846.  A synopsis  of  the  Silurian  fossils  of  Ireland.  72  pp.  5 pis.  Dublin. 

miller,  s.  A.  1874.  Monograph  of  the  lamellibranchiata  of  the  Cincinnati  group.  Cincinnati  Q.  Jl.  Sci.  1 (3), 
211-231. 

— 1877.  American  Palaeozoic  fossils,  a catalogue  of  the  genera  and  species.  253  pp.  Cincinnati. 

— 1889.  North  American  geology  and  palaeontology  for  the  use  of  amateurs,  students  and  scientists.  718  pp. 
Cincinnati. 

mitchell,  w.  i.  1977.  The  Ordovician  Brachipoda  from  Pomeroy,  Co.  Tyrone.  Palaeontogr.  Soc.  [ Monogr.\ , 
138  pp.,  28  pis. 

moore,  R.  c.  ed.,  1969.  Treatise  on  invertebrate  paleontology,  pt.  N.  Mollusca  6,  Bivalvia.  952  pp.  Boulder. 
murchison,  R.  i.  1839.  The  Silurian  System,  founded  on  geological  researches  in  the  counties  of  Salop,  Hereford, 
Radnor,  Montgomery,  Caermarthen,  Brecon,  Pembroke,  Monmouth,  Gloucester,  Worcester  and  Stafford:  with 
descriptions  of  the  coalfields  and  overlying  formations.  London:  John  Murray,  i-xxxii,  1-168,  pis.  1-37. 
pfab,  L.  1934.  Revision  der  Taxodonta  des  bohmischen  Silurs.  Palaeontographica,  A,  80,  195-253,  3 pis. 
pojeta,  J.  1962.  The  pelecypod  genus  Byssonychia  as  it  occurs  in  the  Cincinnatian  at  Cincinnati,  Ohio. 
Palaeontogr.  am.  4 (30),  169-216,  pis.  22-31. 

— 1966.  North  American  Ambonychiidae  (Pelecypoda).  Ibid.  5 (36),  129-241,  pis.  19-47. 

1971.  Review  of  Ordovician  Pelecypods.  Prof.  Pap.  U.S.  geol.  Surv.  695,  46  pp.,  20  pis. 

— 1978.  The  origin  and  early  taxonomic  diversification  of  pelecypods.  Phil.  Trans.  R.  Soc.  Lond.,  B.  284, 
225-246,  pis.  1-15. 

— and  gilbert-tomlinson,  j.  1977.  Australian  Ordovician  pelecypod  molluscs.  Bull.  Bur.  Miner.  Resour. 
Geol.  Geophys.  Aust.  174  (6),  64  pp.,  29  pis. 

— kriz,  j.  and  berdan,  J.  M.  1976.  Silurian-Devonian  Pelecypods  and  Palaeozoic  Stratigraphy  of  subsurface 
Rocks  in  Florida  and  Georgia  and  Related  Silurian  Pelecypods  from  Bolivia  and  Turkey.  Prof.  Pap.  U.S. 
geol.  Surv.  879,  32  pp.,  5 pis. 

— and  runnegar,  b.  1976.  The  Palaeontology  of  rostroconch  mollusks  and  the  early  history  of  the  phylum 
Mollusca.  Ibid.  968,  pp.  1-88,  pis.  1-54. 

portlock,  j.  e.  1837.  In  Larcom,  T.  Ordnance  survey  of  the  County  of  Londonderry,  vol.  1 (with  a preface 
by  Colby,  T.).  Notices  i (at  rear  of  volume),  pp.  3-6,  pis.  1-3. 

— 1843.  Report  on  the  geology  of  Londonderry  and  parts  of  Tyrone  and  Fermanagh,  i-xxxi,  784  pp.,  pis. 
1-38,  A-I,  map.  Dublin  and  London. 

reed,  F.  R.  c.  1944.  Some  new  Ordovician  lamellibranchs  from  Girvan.  Ann.  Mag.  nat.  Hist.  Ser.  11,11  (76), 
xxiii,  pp.  209-215,  pi.  2. 

— 1952.  Revision  of  certain  Ordovician  fossils  from  County  Tyrone.  Proc.  R.  Ir.  Acad.  55  B (3),  31-136, 
pis.  1-5. 

sharpe,  d.  1853.  Description  of  the  new  species  of  zoophyta  and  mollusca.  [Appendix  B of  ribeiro,  c.  On 
the  Carboniferous  and  Silurian  Formations  of  the  neighbourhood  of  Bussaco  in  Portugal.]  Q.  Jl.  geol.  Soc. 
Lond.  9,  146-158,  pis.  7-9. 

soot-ryen,  H.  1964  (1965).  Nuculoid  pelecypods  from  the  Silurian  of  Gotland,  Arkiv  for  miner,  geol.  3 (28), 
489-519,  pis.  1-5. 

sowerby,  j.  de  c.  1839.  In  murchison,  1839. 

tunnicliff,  s.  p.  1980.  A catalogue  of  the  Lower  Palaeozoic  fossils  in  the  collection  of  Major  General  J.  E. 

Portlock,  R.E.,  LL.D.,  F.R.S.,  F.G.S.  etc.  112  pp.,  5 figs.  Ulster  Museum  Publication. 
ulrich,  e.  o.  1892.  New  Lower  Silurian  lamellibranchiata  chiefly  from  Minnesota  rocks.  Minnesota  geol.  and 
nat.  Hist.  Surv.  19,  211-248. 

— 1893  [1895].  New  and  little  known  Lamellibranchiata  from  the  Lower  Silurian  rocks  of  Ohio  and  adjacent 
states.  Ohio  Div.  geol.  Surv.  Rept.  Inv.  7 (2),  627-693,  pis.  45-56. 

— 1894.  The  Lower  Silurian  Lamellibranchiata  of  Minnesota:  from  vol.  3 of  the  Final  Rept.,  Minnesota 
geol.  and  nat.  Hist.  Surv.  475-628,  pis.  35-42.  (Published  under  separate  cover  prior  to  the  entire  v.  3:  fide 
Pojeta,  1971,  p.  41.) 

Watkins,  R.  1978.  Bivalve  ecology  in  a Silurian  shelf  environment.  Lethaia , 11,  41-56. 
whiteaves,  J.  F.  1897.  The  fossils  of  the  Galena-Trenton  and  Black  River  formations  of  Lake  Winnipeg  and 
its  vicinity.  Geol.  Surv.  Can.  Palaeozoic  Fossils,  3 (3),  129-242,  pis.  16-22. 
whitfield,  r.  p.  1893  [1895],  Contributions  to  the  palaeontology  of  Ohio.  Ohio  Div.  geol.  Surv.  7 (2),  407-454, 
pi.  13. 


PALAEONTOLOGY,  VOLUME  25 


WILLIAMS,  A.,  STRACHAN,  I.,  BASSETT,  D.  A.,  DEAN,  W.  T.,  INGHAM,  J.  K.,  WRIGHT,  A.  D.,  and  WHITTINGTON,  H.  B. 

1972.  A correlation  of  Ordovician  rocks  in  the  British  Isles.  Geol.  Soc.  Lond.,  Spec.  Rep.  3,  74  pp. 
williams,  h.  s.  and  breger,  c.  l.  1916.  The  fauna  of  the  Chapman  Sandstone  of  Maine,  including  descriptions 
of  some  related  species  from  the  Moose  River  sandstone.  Prof.  Pap.  U.S.  geol.  Surv.  89,  347  pp.,  27  pis. 

S.  P.  TUNNICLIFF 

Palaeontology  Unit 
Institute  of  Geological  Sciences 
Exhibition  Road 
London  SW7  2DE 


Typescript  received  26  February  1980 
Revised  typescript  received  27  November  1980 


JUVENILE  SPECIMENS  OF  THE 
ORNITHISCHIAN  DINOSAUR  PSITTACOSAURUS 

by  WALTER  P.  COOMBS,  JR. 


Abstract.  Hitherto  undescribed  specimens  of  Psittacosaurus  mongoliensis  from  the  Oshih  Fm.,  Mongolian 
Peoples’  Republic,  include  two  almost  complete  skulls  and  numerous  postcranial  elements.  A rostral  bone, 
present  in  these  and  other  specimens  of  Psittacosaurus,  is  a cranial  element  otherwise  known  only  in 
ceratopsians,  and  its  presence  indicates  a sister  group  relationship  for  the  Psittacosauridae  and  Ceratopsia. 
Each  of  the  two  new  specimens  of  Psittacosaurus  is  a juvenile,  and  are  among  the  smallest  dinosaur  specimens 
yet  described.  Parental  attendance  of  nests  was  common,  possibly  universal  among  dinosaurs,  but  post-hatching 
parental  care  is  uncertain.  Juveniles  of  Psittacosaurus,  as  well  as  those  of  other  dinosaurs,  may  have  formed 
sibling  groups. 

Two  skeletons  of  small  bipedal  ornithischians  collected  by  the  Third  Asiatic  Expedition  (1922)  of 
the  American  Museum  of  Natural  History  were  described  by  Osborn  (1923,  1924)  as  the  types  of 
Psittacosaurus  mongoliensis  (Oshih  Fm.)  and  Protiguanodon  mongoliense  (Andai  Sair  Fm.),  thought 
to  be  a primitive  ankylosaur  and  a primitive  iguanodontid,  respectively.  To  the  family  Psitta- 
cosauridae, erected  for  reception  of  these  two  species  (Osborn  1923),  have  been  added  several  new 
species  of  Psittacosaurus  and  additional  fragmentary  remains  from  various  localities  in  Mongolia 
and  northern  China  (Young  1931,  1958;  Bohlin  1953;  Maleev  1954;  Rozhdestvensky  1955;  Chao 
1962).  Described  below  are  two  additional  specimens  of  Psittacosaurus  from  the  same  locality  as 
the  type  of  P.  mongoliensis. 


MATERIAL 

AMNH  6535  (=  American  Museum  of  Natural  History,  New  York),  partial  skull  and  jaws. 

AMNH  6536,  almost  complete  skull  and  jaws  with  numerous  distarticulated  postcranial  elements  including: 
cervical,  dorsal,  and  caudal  vertebrae;  ribs;  scapulae;  coracoid;  partial  ilium  and  ischium;  humeri;  femora; 
tibiae;  fibulae;  and  an  almost  complete  left  pes.  The  postcranial  material  belongs  to  several  individuals  (there 
are  fourteen  distal  ends  of  tibiae)  of  at  least  two  different  sizes  (some  femoral  fragments  are  tiny  and  more 
compatible  in  size  with  the  skull  AMNH  6535).  Rather  than  allot  separate  numbers  to  every  element,  all  of 
the  postcranial  material  is  arbitrarily  assigned  to  AMNH  6536. 

Locality.  Both  specimens  come  from  the  Oshih  Fm.,  Artsa  Bogdo  basin,  western  Mongolian  Peoples’ 
Republic,  and  were  apparently  collected  at  the  same  time  and  in  the  same  place  as  the  type  of  Psittacosaurus 
mongoliensis  (Osborn  1923,  1924). 


DESCRIPTION 

Skull,  general.  In  lateral  view  the  skull  is  arched  dorsally,  truncated  posteriorly,  and  has  a short  snout  (text-figs. 
1,  3,  4).  AMNH  6536  is  laterally  crushed,  but  AMNH  6535  retains  the  characteristic  triangular  shape  of 
psittacosaur  skulls,  with  the  greatest  breadth  across  the  flaring  jugals  and  a sharply  pointed  apex  at  the  beak 
(PI.  14).  The  apparent  relatively  large  size  of  the  brain  in  AMNH  6535  is  a juvenile  character  (PI.  14, 
figs.  1 and  2).  The  diameter  of  the  orbit  equals  33  % of  cranial  length  in  AMNH  6536;  38  % in  AMNH  6535, 
the  relatively  large  orbit  also  being  a juvenile  characteristic.  The  nostril  is  clearly  delineated  only  on  the  right 
side  of  AMNH  6536,  on  which  it  is  small,  circular,  and  approximately  at  the  level  of  the  upper  half  of  the  orbit. 

Cranial  fenestrae.  Both  skulls  are  damaged  posteriorly  and  borders  of  all  temporal  fenestrae  are  incomplete. 
The  left  supratemporal  fenestra  of  AMNH  6536  is  oval  with  its  long  axis  paralleling  the  midline.  The  braincase 
is  broad  and  devoid  of  a sagittal  crest  between  the  upper  fenestrae  (PI.  14).  Lateral  temporal  fenestrae  are 


IPalaeontology,  Vol.  25,  Part  1,  1982,  pp.  89-107,  pi.  14.) 


90 


PALAEONTOLOGY,  VOLUME  25 


evidently  tall  and  narrow,  and  considerably  smaller  than  the  orbit,  unlike  the  condition  in  adult  Psittacosaurus 
in  which  the  fenestrae  are  larger  than  the  orbit  (e.g.  Osborn  1923,  fig.  2).  AMNH  6536  has  a shallow  depression, 
its  depth  somewhat  exaggerated  by  crushing,  between  nostril  and  orbit  in  the  region  where  an  antorbitai 
fenestra  might  be  located,  but  there  is  no  opening  into  the  interior  of  the  skull  (text-fig.  1).  AMNH  6535  is 
badly  crushed  in  the  immediate  pre-orbital  region,  but  it  appears  that  a true  antorbitai  fenestra  was  not  present. 


text-fig.  1.  AMNH  6536,  skull,  lower  jaws,  and  anterior  cervical  vertebrae;  right 
side  with  key  drawing.  Abbreviations:  at,  atlas;  ax,  neural  arch  of  axis;  cax, 
centrum  of  axis;  d,  dentary;  en,  external  naris;  f,  frontal;  j,  jugal;  l,  lacrimal; 
mx,  maxilla;  n,  nasal;  nax,  neural  spine  of  axis;  pa,  parietal;  pd,  predentary; 
pf,  prefrontal;  pm,  premaxilla;  po,  postorbital;  pp,  palpebral;  q,  quadrate;  r,  rostral; 
ra,  retroarticular  process;  sq,  squamosal;  and  v,  cervical  vertebrae  (post-atlas). 

Length  of  reference  line  = 1 0 cm  (approximately  twice  natural  size). 


COOMBS:  JUVENILE  DINOSAURS 


91 


Rostral.  Following  Maryariska  and  Osmolska  (1975,  p.  172),  the  most  anterior  element  of  the  snout  is  identified 
as  a rostral,  a bone  diagnostic  of  ceratopsians,  rather  than  as  a premaxilla  as  suggested  by  Osborn  (1923;  see 
also  Young  1958;  Chao  1962).  The  rostral,  best  preserved  on  the  smaller  skull  (text-fig.  4),  is  roughly  triangular 
in  lateral  view  with  an  ascending  ramus  that  curves  posteriorly  to  terminate  near  the  ventral  margin  of  the 
nostril.  Most  of  the  cutting  margin  of  the  beak  is  composed  of  the  rostral,  but  it  is  unclear  to  what  extent 
the  rostral  contributes  to  the  anterior  palatal  shelf. 

Premaxilla.  Exact  limits  of  the  premaxillae  are  unclear,  but  the  bone  probably  intervenes  between  the  rostral 
and  the  external  nares  and  forms  all  the  ventral  border  of  the  latter  opening  (text-fig.  3;  Maryanska  and 
Osmolska  1975).  Because  other  authors  have  identified  the  rostral  as  the  premaxilla,  the  true  premaxilla  has 
been  identified  as  part  of  the  maxilla.  In  consequence,  Psittacosaurus  has  been  reconstructed  with  a very  large 
maxilla  that  separates  the  premaxilla  from  the  lacrimals  as  well  as  from  the  margin  of  the  nostril  (Osborn 
1923,  fig.  2a;  Young  1958,  fig.  51;  Chao  1962,  fig.  1)  a pattern  atypical  for  ornithischians.  The  alternative 
interpretation  presented  here  (text-fig.  3)  indicates  extensive  contact  of  premaxilla  and  lacrimal  with  exclusion 
of  the  maxilla  from  the  border  of  the  nostril,  a pattern  common  among  ceratopsians  and  ornithischians  in 
general  (Lull  1933;  Romer  1956).  Ornithischians  typically  have  a bony  roof  to  the  anterior  part  of  the  buccal 
cavity  composed  of  premaxillae,  maxillae,  and  sometimes  the  vomers.  In  the  Ceratopsia  the  rostral  bone 
forms  much  of  the  cutting  edge  of  the  beak,  but  forms  only  a small  segment  of  the  anterior  palatal  shelf 
(Lull  1933,  figs.  5 and  30;  Maryanska  and  Osmolska  1975,  fig.  9).  Psittacosaurus  may  have  a similar  arrange- 
ment, but  sutures  are  unclear  on  the  present  specimens  (PI.  14,  figs.  3 and  4). 

Maxilla.  Anteriorly  the  maxilla  may  contact  the  posterior  limit  of  the  rostral  along  the  edge  of  the  mouth 
in  the  present  specimens,  but  contact  is  lost  in  adult  Psittacosaurus.  Maxillary  teeth,  numbering  at  least  five  in 
AMNH  6535  and  at  least  six  in  AMNH  6536,  are  withdrawn  medially,  so  that  the  maxilla  forms  a lateral  shelf 
thus  delineating  the  dorsal  border  of  a ‘cheek-pouch’,  a common  feature  among  ornithischians  (Galton  1973). 

Jugal.  All  the  ventral  margin  of  the  orbit  is  formed  by  the  jugal,  the  suborbital  bar  being  relatively  slender 
in  AMNH  6535,  considerably  wider  and  giving  an  impression  of  massiveness  in  AMNH  6536,  but  in  neither 
skull  is  the  jugal  as  wide  as  in  adult  Psittacosaurus.  A ventro-laterally  projecting  jugal  spine,  located  below 
the  lateral  temporal  fenestra,  is  present  on  the  left  jugal  of  AMNH  6536  (text-fig.  2),  but  is  considerably 
smaller  than  the  prominent  flange  of  adult  Psittacosaurus  skulls  (Osborn  1923,  1924;  Young  1958;  Chao 
1962).  A slender  ascending  ramus  of  the  jugal  forms  part  of  the  postorbital  bar.  A quadratojugal  presumably 
intervened  between  the  jugal  and  the  ventral  end  of  the  quadrate,  but  the  element  cannot  be  distinguished 
on  either  of  the  present  specimens. 

Quadrate.  The  elongate  quadrate  curves  anteriorly  from  its  dorsal,  squamosal  articulation  and  is  two  to  three 
times  wider  than  the  postorbital  bar  (text-figs.  2 and  3).  The  mandibular  cotylus  is  compressed  antero- 
posteriorly  and  has  a bulbous  lateral  region  adjacent  to  a narrower  medial  area  (PI.  14,  figs.  3 and  4).  The  cotylus 
projects  below  the  level  of  the  tooth  row,  as  in  most  ornithischians. 

Skull  roof.  The  nasals  presumably  form  the  narrow  bar  between  the  nostrils  as  well  as  the  wider  region 
immediately  posterior  to  them,  but  sutures  of  the  nasals  with  rostral,  maxillae  and  premaxillae  are  not  visible 
on  either  specimen.  Chao  (1962)  illustrates  P.  youngi  with  the  nasals  excluded  from  the  narial  border  (text-fig. 
7c),  but  this  is  not  true  of  the  type  of  P.  mongoliensis  (text-fig.  7d).  A prefrontal  nestles  into  the  antero-dorsal 
margin  of  the  orbit  excluding  the  nasal  from  the  orbital  rim  (right  side  of  AMNH  6536,  text-figs.  1 and  3). 
Ventrally  the  prefrontal  contacts  the  lacrimal  and  premaxilla,  but  sutures  are  indistinct.  A displaced  palpebral 
(‘supraorbital’  of  Osborn  1923;  ‘prefrontal’  of  Chao  1962;  see  Coombs  1972)  originally  articulated  with  the 
prefrontal  at  the  anterior  rim  of  the  orbit  (text-figs.  1 and  3).  The  palpebral  has  an  expanded  base  and  a 
tapered  posterior  extension  that  is  incompletely  preserved.  Posterior  to  the  indistinct  contact  of  nasals  and 
frontals  above  the  orbits  the  skull  roof  widens  and  the  postorbital  contacts  the  lateral  edge  of  the  frontal.  A 
slender  descending  rod  of  the  postorbital  bone  forms  most  of  the  postorbital  bar.  The  relative  contributions 
of  the  postorbitals,  frontals,  and  parietals  to  the  margins  of  the  supratemporal  fenestrae  cannot  be  determined. 
Parietals  form  most  of  the  arched  posterior  skull  roof,  without  the  sagittal  crest  present  in  adult  skulls  (see 
also  skull  of  Psittacosaurus  osborni  Young,  1931,  fig.  2).  The  parietals  bend  sharply  downward  at  the  posterior 
margin  of  the  skull  roof  to  form  the  most  dorsal  surface  of  the  occipital  region. 

Occipital  area.  Very  little  of  this  region  is  exposed.  The  rugose  distal  tip  of  the  left  opisthotic  of  AMNH 
6536  projects  from  matrix  well  below  the  level  of  the  skull  roof  and  the  dorsal  tip  of  the  quadrate  (text- 
fig.  2).  The  element  is  roughly  triangular  in  section,  not  as  flattened  as  is  typical  for  Ornithischia. 


92 


PALAEONTOLOGY,  VOLUME  25 


Palate.  Some  palatal  details  are  visible  on  the  smaller  skull  (PI.  14,  figs.  3 and  4).  An  anterior  palatal  shelf 
is  formed  partly  of  the  rostral  and  partly  of  the  premaxillae,  and  is  roughly  semicircular  in  shape,  being  less 
pointed  and  less  triangular  than  in  adult  Psittacosaurus.  The  vomers  form  a median  vertical  plate  that 
dominates  the  interior  palatal  region.  The  ventral  keel  of  this  plate  extends  below  the  level  of  the  tooth  rows, 
a condition  found  in  several  ornithischians  (e.g.  the  ankylosaur  Panoplosaurus;  Russell  1940).  A narrow  shelf 
from  the  maxillae  appears  to  contact  the  vomer  keel  anteriorly,  thus  shifting  the  internal  nares  posteriorly. 
The  internal  nares  are  bounded  laterally  and  anteriorly  by  maxillae,  medially  by  the  vomer  keel,  and  are 
bounded  posteriorly  by  a broad  plate  of  bone  that  slants  obliquely  anteriorly  and  upward  toward  the  skull 


text-fig.  2.  AMNH  6536,  skull,  lower  jaws  and  anterior  cervical  vertebrae;  left 
side  with  key  drawing.  Abbreviations:  cat,  centrum  of  atlas;  cax,  centrum  of  axis; 
d,  dentary;  j,  jugal;  mrj,  medial  side  of  right  lower  jaw;  mx,  maxilla;  o,  opisthotic; 
pd,  predentary;  pm,  premaxilla;  po,  postorbital;  q,  quadrate;  sp,  spine  of  jugal; 
v,  cervical  vertebra  (post-atlas).  Length  of  reference  line  = 10  cm  (approximately 
twice  natural  size). 


COOMBS:  JUVENILE  DINOSAURS 


93 


text-fig.  3.  Restoration  of  a juvenile  Psittacosaurus  skull  based  primarily  on 
AMNH  6536.  Abbreviations:  d,  depression  in  premaxilla;  f,  frontal;  j,  jugal; 
l,  lacrimal;  mx,  maxilla;  n,  nasal;  o,  opisthotic;  pd,  predentary;  pf,  prefrontal; 
pm,  premaxilla;  pp,  palpebral;  q,  quadrate;  r,  rostral;  and  s,  spine  of  jugal. 


roof.  The  latter  plate  is  composed  of  pterygoid,  ectopterygoid,  and  palatine  bones,  but  sutures  are  unclear. 
The  palatal  structure  of  Psittacosaurus  is  similar  to  that  of  many  Ornithischia,  especially  the  quadrupedal 
forms,  and  may  be  compared  particularly  to  the  palate  of  Bagaceratops  (Maryanska  and  Osmolska  1975, 
fig.  9). 

Basicranium.  The  partially  exposed  basicranium  of  AMNH  6535  (PI.  14,  figs.  3 and  4)  is  broad  relative  to 
other  skull  dimensions  (a  juvenile  feature),  and  has  a median  longitudinal  depression  that  is  flanked  by 
anteriorly  converging  ridges.  A small,  subspherical,  posteriorly  directed  occipital  condyle  projects  only  slightly 
below  the  level  of  the  basicranial  floor. 

Lower  jaw.  The  deep,  massive  lower  jaw  has  a straight  ventral  margin  and  medially  displaced  dentary  teeth 
(text-fig.  1).  A curved  shelf  marks  the  ventral  border  of  the  cheek  pouch.  The  predentary  and  dentary  bones 
are  fused  together.  Posteriorly  there  is  a pointed  coronoid  process  projecting  upward  medial  to  the  most 
anterior  ventral  corner  of  the  lateral  temporal  fenestra.  A short  retroarticular  process  projects  backward  from 
the  articular  (text-fig.  1). 

Teeth.  Anterior  maxillary  teeth  have  three  ridges  on  the  lateral  surface  that  upon  wear  produce  the  ‘trilobate’ 
cutting  margin  described  by  Osborn  (1924;  text-fig.  2).  Posterior  maxillary  teeth  may  have  four  ridges  and 
are  generally  larger  but  less  worn  than  anterior  teeth.  All  teeth  that  are  adequately  exposed  on  the  two  skulls 
are  worn  to  some  degree.  There  are  no  premaxillary  teeth. 

Vertebrae.  Cervical  vertebrae  attached  to  the  skull  of  AMNH  6536  include  the  axis,  which  has  a low,  laterally 
compressed,  elongate  neural  spine  as  described  in  the  type  of  Protiguanodon  mongoliense  (AMNH  6253; 
Osborn  1924).  A small  block  of  bones  (part  of  AMNH  6536)  contains  several  vertebrae,  mostly  dorsals 
(text-fig.  5).  The  centra  are  laterally  compressed,  with  expanded,  roughly  heart-shaped  articular  ends.  The 
neural  arches  are  not  fused  to  the  centra,  a juvenile  characteristic.  A particularly  well-preserved  neural  arch 
of  a dorsal(?)  vertebra  has  about  the  same  length  as  its  centrum;  the  incomplete  zygapophyses  arch  upward 
and  there  is  an  almost  circular  diapophysis  high  on  the  side  of  the  neural  arch.  A sacral  vertebra  in  the  same 
block  has  a centrum  similar  in  shape  to  dorsal  centra,  and  has  a low,  broad  neural  arch  with  an  elongate, 
laterally  compressed  neural  spine  most  of  which  is  broken  off.  The  sacral  rib  articulated  on  anteroposteriorly 
compressed  diapophyses  positioned  at  the  extreme  anterior  end  of  this  sacral,  which  judged  from  the  sacral 


94 


PALAEONTOLOGY,  VOLUME  25 


text-fig.  4.  AMNH  6535,  skull,  partial  lower  jaw,  and  fragmentary  anterior 
cervicals,  right  side.  Length  of  reference  line  = 10  cm  (approximately  twice 
natural  size). 

series  of  Protoceratops  (Brown  and  Schlaikjer  1940),  is  from  the  middle  of  the  sacral  series.  Another  sacral 
vertebra  in  the  block  also  has  an  elongate,  compressed  diapophysis  but  it  is  oriented  almost  horizontally, 
suggesting  that  this  vertebra  belongs  to  the  posterior  end  of  the  sacral  series.  A group  of  three  caudal  centra 
(text-fig.  6n),  all  lacking  the  neural  arch,  are  smaller  than  either  dorsals  or  sacrals  but  have  similar  shape  and 
proportions.  Transverse  processes  are  not  present  on  any  of  the  caudals.  The  largest  of  these  three  caudal 
centra  is  about  4-5  mm  in  length. 

Ribs.  Rib  fragments  embedded  in  the  block  with  other  postcranial  elements  have  cross  sections  that  are  either 
circular  (proximal  ends)  or  compressed  (distal  ends,  text-fig.  5).  One  proximal  end  fragment  has  a large, 
almost  circular  head  that  matches  in  size  and  shape  the  articular  surface  on  the  dorsal  vertebra  described 
above,  and  a very  small  tubercle  positioned  a short  distance  from  the  head. 

Pectoral  girdle.  Scapulae  are  represented  by  two  almost  complete  elements  (text-fig.  5),  plus  a partial  scapular 
blade.  The  scapulae  are  elongate  and  narrow,  with  a much  wider,  slightly  concave  area  adjacent  to  the  glenoid 
and  coracoid  articulation.  The  upper  end  of  the  blade  is  broad  but  flattened.  A low  scapular  spine  arises 
from  the  extreme  anterior  edge  of  the  scapula  directly  opposite  the  glenoid.  The  glenoid  articulation  is  short, 
broad,  slightly  concave,  and  has  a rather  massive  dorsal  lip.  The  coracoid  articulation  is  elongate  and 


EXPLANATION  OF  PLATE  14 

Two  views  of  AMNH  6535,  a skull,  partial  lower  jaw  and  fragments  of  anterior  cervical  vertebrae. 

Fig.  1.  Dorsal  view,  stereo  pair. 

Fig.  2.  Key  to  structures  visible  in  the  dorsal  view.  Stippling  indicates  the  approximate  extent  of  a natural 
endocranial  cast  that  has  been  partially  exposed  through  loss  of  some  dorsal  cranial  bones.  The  anterior 
limit  of  the  endocranial  space  was  determined  in  part  by  examination  of  AMNH  6536  which  has  the 
anterior  tip  of  a natural  endocranial  cast  exposed  through  loss  of  some  skull  bones. 

Fig.  3.  Ventral  view,  stereo  pair. 

Fig.  4.  Key  to  structures  visible  in  ventral  view.  Abbreviations:  at,  atlas;  ax,  axis;  bo,  basioccipital;  c,  left 
coronoid  process  (protruding  through  matrix);  f,  frontal;  j,  jugal;  l,  lacrimal;  lj,  lower  jaw;  m,  maxilla; 
n,  nasal;  o,  opisthotic;  pm,  premaxilla;  po,  postorbital;  pt,  pterygoid;  q,  quadrate;  r,  rostral;  t,  tooth.  Length 
of  reference  line  = 1-0  cm  (approximately  twice  natural  size). 


PLATE  14 


COOMBS,  Psittacosaurus 


96 


PALAEONTOLOGY,  VOLUME  25 


text-fig.  5.  AMNH  6536,  three  views  of  a block  containing  various  postcranial  elements,  with  key 
drawings.  Abbreviations:  c,  centrum;  f,  femur;  il,  ilium;  is,  ischium,  n,  neural  arch,  r,  rib;  sc,  scapula; 
sr,  sacral  rib;  and  a question  mark  denoting  an  unidentified  fragment.  Length  of  reference  line  =10 
cm  (approximately  twice  natural  size). 


text-fig.  6.  AMNH  6536,  various  postcranial  elements,  a,  partial  left  pes  with  distal  ends  of  tibia  and  fibula; 
b,  posterior  view  of  a partial  left  tibia  with  astragalus  and  metatarsals  II,  III,  and  IV;  c,  same  as  b in  anterior 
view;  d,  same  as  b in  distal  view;  E,  left  humerus,  flexor  surface;  f,  right  coracoid,  lateral  view;  G,  H,  i,  and 
J,  distal  ends  of  four  femora,  posterior  views;  K,  L,  and  m,  proximal  ends  of  three  tibiae,  anterior  views; 
N,  block  containing  three  caudal  centra;  o and  p,  proximal  ends  of  two  small  femora,  anterior  views;  Q,  R,  s, 
and  t,  distal  ends  of  four  tibiae,  anterior  views,  with  distal  ends  of  fibulae  on  r and  T.  Scale  divisions  in  mm 
(approximately  twice  natural  size). 


PALAEONTOLOGY,  VOLUME  25 


triangular,  exactly  matching  an  associated  left  coracoid  (text-fig.  6f).  The  coracoid  has  a broad,  slightly 
concave  glenoid  and  a foramen  that  is  roughly  equidistant  from  the  glenoid  and  scapular  articulations.  There 
is  a prominent  biceps  tuberosity  on  the  anterior  edge  of  the  coracoid,  from  which  a ridge  extends  posteriorly 
toward  the  glenoid.  Below  the  biceps  tuberosity  the  coracoid  has  a shallowly  concave,  medially  inclined 
surface  that  is  the  probable  region  of  coracobrachialis  origin.  Clavicles  have  been  reported  for  Psittacosaurus 
mongoliensis  (Osborn  1924)  and  P.  sinensis  (Young  1958),  but  there  are  no  clavicles  associated  with  the 
scapulae  of  the  present  specimens  (text-fig.  5).  In  view  of  the  jumbled  condition  of  the  bones,  the  clavicles 
could  well  be  displaced  and  easily  mistaken  for  rib  fragments. 

Pelvic  girdle.  Of  the  pelvic  girdle  there  is  one  fragmentary  piece  each  of  ilium  and  ischium  (text-fig.  5).  The 
ilial  fragment  is  flat  with  a blunt,  ventrally  projecting  ischial  peduncle  and  a laterally  compressed,  rather  thin 
postacetabular  segment.  Only  the  base  of  the  pubic  peduncle  is  preserved.  The  ischial  fragment  includes  an 
elongate,  laterally  compressed  pubic  articulation,  and  an  ilial  articulation  that  appears  to  be  more  compact 
and  rounded,  although  most  of  the  latter  is  hidden  below  the  ilial  fragment.  The  ischial  blade  slants  posteriorly 
and  is  narrow  and  laterally  compressed.  Most  of  the  distal  part  of  the  ischium  is  missing. 

Forelimb.  In  addition  to  a perfect  left  humerus  (text-fig.  6e),  there  are  two  proximal  ends  of  right  humeri, 
one  about  the  same  size  as  the  left  humerus  and  one  considerably  smaller.  Proximally,  the  humerus  is  broad, 
but  it  tapers  quickly  to  a roughly  circular  shaft,  then  widens  at  the  distal  articulations  which  are  rotated 
relative  to  the  head  such  that  the  extensor  surface  is  twisted  medially.  A deltopectoral  crest  projects  ventrally 
and  is  rather  close  to  the  proximal  end  of  the  humerus.  No  other  forelimb  elements  have  been  identified. 

Hindlimb.  There  are  numerous  fragments,  but  unfortunately  there  is  no  association  of  a particular  femur  with 
a particular  tibia,  nor  with  the  pes  described  below,  so  that  hindlimb  proportions  cannot  be  established.  The 
proximal  end  of  a relatively  large  femur  has  a compressed,  crest-like  greater  trochanter  with  an  adjacent, 
finger-like  lesser  trochanter  separated  from  the  former  by  a deep  furrow,  and  an  irregularly  circular  shaft 
(text-fig.  5).  Two  much-smaller  femora  have  a virtually  identical  set  of  trochanters  and  a medially  displaced 
head  that  is  as  much  cylindrical  as  spherical  (text-fig.  6o,  p).  The  femoral  shaft  is  curved  with  the  convex 
surface  lateral,  a common  shape  for  femora  of  bipedal  Ornithischia.  None  of  the  eight  distal  femoral  pieces 
included  in  AMNH  6536  can  be  fitted  on  to  any  of  the  proximal  ends.  Distal  articulations  follow  the  usual 
bipedal  ornithischian  pattern  of  a larger  lateral  and  smaller  medial  articular  surface  (text-fig.  6g,  h,  i,  j). 

There  are  seven  proximal  and  fourteen  distal  ends  of  tibiae  (text-fig.  6k,  l,  m,  q,  r,  s,  t).  Proximally,  the 
tibia  has  an  antero-posteriorly  elongated  articular  surface  that  has  a smoothly  curved  medial  surface  and  a 
notched  lateral  surface  that  gives  the  proximal  end  a trilobate  shape  viewed  end-on.  The  tibial  shaft  tapers 
to  a slender  rod  of  almost  circular  cross-section.  Distally,  the  tibia  is  strongly  compressed  antero-posteriorly, 
creating  a wide  articular  surface  that  is  a little  thicker  medially  and  has  a slight  depression  near  the  centre 
for  reception  of  the  astragalus.  There  is  a slight  ventral  projection  at  the  lateral  corner  of  the  distal  end, 
marking  the  fibular  articulation.  Fibular  fragments  still  attached  to  two  of  the  tibiae  (text-fig.  6r,  t)  are  oval 
in  section,  only  a fraction  of  the  diameter  of  the  tibial  shaft,  and  fit  into  a shallow  depression  on  the 
antero-lateral  edge  of  the  distal  end  of  the  tibia. 

Pes.  The  astragalus  is  compressed  against  and  covers  about  two-thirds  of  the  distal  end  of  the  tibia  (text-fig. 
6a,  b,  c,  d)  leaving  a small  space  laterally  for  the  calcaneum.  The  lateral  edge  of  the  astragalus  is  notched 
for  reception  of  the  calcaneum,  and  anteriorly  the  astragalus  curves  up  on  to  the  flexor  surface  of  the  tibia. 
A roughly  cup-shaped  calcaneum  nestles  against  the  ventrally  projecting  lateral  corner  of  the  tibia  (text-fig. 
6a),  and  as  preserved  does  not  contact  either  the  astragalus  or  fibula.  A distal  row  of  tarsals  is  not  present. 

As  is  typical  of  bipedal  ornithischians,  the  metatarsals  are  long,  slender,  and  fit  tightly  together,  especially 
at  their  proximal  ends  (text-fig.  6a,  c,  d).  As  in  the  types  of  Psittacosaurus  mongoliensis  and  P.  sinensis, 
metatarsals  I through  IV  are  present,  but  there  is  no  trace  of  metatarsal  V nor  is  there  an  articular  surface 
for  it  on  metatarsal  IV.  Metatarsal  I is  strongly  compressed  at  the  proximal  end  and  through  the  upper  two- 
thirds  of  the  shaft.  Metatarsal  II  is  about  double  the  width  of  metatarsal  I proximally,  but  is  still  somewhat 
compressed.  Metatarsal  II  has  a wider  shaft  and  is  longer  than  metatarsal  I.  Metatarsal  III  has  an  irregularly 
square  proximal  articular  surface  and  is  the  longest,  straightest,  and  most  massive  of  the  metatarsals.  Metatarsal 
IV  is  about  the  same  length  as  metatarsal  II,  but  is  more  compressed  at  the  proximal  end,  and  its  shaft  curves 
slightly  away  from  the  axis  of  the  pes  running  through  metatarsal  III.  Phalanges  have  the  typical  form  found 
in  small  bipedal  ornithischians  (text-fig.  6a).  All  are  relatively  long  (length  greater  than  diameter)  and  are 
circular  to  subrectangular  in  cross-section,  not  dorsoventrally  compressed.  One  almost  complete  ungual  of 
digit  III  is  slightly  longer  than  the  adjacent  phalanx  and  has  a narrow,  semi-claw  shape  as  in  Leptoceratops 
(Brown  1914;  Brown  and  Schlaikjer  1942)  rather  than  the  broad,  hoof-like  unguals  of  Protoceratops  (Brown 


COOMBS:  JUVENILE  DINOSAURS 


99 


and  Schlaikjer  1940).  The  pes  is  narrow,  long,  and  gracile  (text-fig.  6a),  suggesting  a relatively  fleet  animal. 
Of  interest  is  the  manner  in  which  the  pes  is  folded  against  the  anterior  surface  of  the  tibia  in  one  specimen 
(text-fig.  6b,  c,  d),  a post-mortem  posture  also  found  in  the  types  of  Psittacosaurus  mongoliensis,  Protiguanodon 
mongoliense,  and  Stenopelix  valdensis  (Osborn  1923,  1924;  Meyer  1857;  Koken  1887).  When  these  animals 
died,  they  settled  on  to  their  belly  with  the  hindlimbs  folded  into  a sharp  ‘Z’,  the  knee  projecting  forward, 
ankle  backward,  femur  atop  the  tibia,  and  pes  pressed  against  the  anterior  surface  of  the  tibia  and  apparently 
set  flat  against  the  ground  (this  posture  is  imperfectly  preserved  also  in  specimens  of  Psittacosaurus  sinensis'. 
Young  1958,  figs.  52,  53,  and  54).  The  imprint  of  the  pes  would  conform  to  the  sitting  dinosaur  trackways 
from  the  Connecticut  Valley  (e.g.  Sauropus  barrattii  and  Anomoepus  scambus  specimens  described  by  Lull 
1953).  The  belly-down  position  is  not  common  among  bipedal  ornithischians.  The  skeletons  of  Iguanodon 
and  most  hadrosaurids  are  typically  found  lying  on  their  side  with  the  knee  and  ankle  both  flexed  rearward, 
or  with  the  pes  almost  aligned  with  the  tibia.  Psittacosaurs  apparently  had  great  mobility  of  the  knee  and 
mesotarsal  joint,  with  hyperextensibility  of  the  ankle  permitting  the  entire  length  of  the  pes  to  be  laid  flat  on 
the  ground.  The  difference  in  joint  mobility  in  psittacosaurs  and  large  bipeds  may  be  related  to  total  body 
size,  the  larger  ornithischians  sacrificing  some  ankle  extensibility  for  the  sake  of  rigidity  necessary  for 
weight-bearing. 


DISCUSSION 

Taxonomy.  Rozhdestvensky  (1955)  regarded  Psittacosaurus  mongoliensis  Osborn  (1923)  as  the 
senior  synonym  of  the  following:  Protiguanodon  mongoliense  Osborn  (1923  = Psittacosaurus 
protiguanodonensis  of  Young  1958);  Psittacosaurus  osborni  Young  (1931);  ? Psittacosaurus  tingi 
Young  (1931);  and  Protiguanodon  cf.  mongoliense  (of  Young  1931).  Trivial  differences  among  these 
species  are  reasonably  regarded  as  stemming  from  ontogenetic  state  and  individual  variation. 
Psittacosaurus  sinensis  Young  (1958)  and  P.  youngi  Chao  (1962)  may  also  be  junior  synonyms  of 
P.  mongoliensis  although,  without  having  all  the  specimens  in  hand  for  detailed  comparison  of 
measurements  and  proportions  not  available  in  extant  descriptions,  formal  suppression  of  all  but 
one  species  is  premature.  Stenopelix  valdensis  Meyer  (1857;  see  also  Koken  1887;  Huene  1908)  from 
Wealden  deposits  (Swinton  1936)  of  northern  Germany  was  originally  classified  in  the  Hypsilopho- 
dontidae  (Huxley  1870),  but  has  also  been  suggested  as  a ceratopsian  (Huene  1909;  Lull  1910)  and 
a psittacosaur  (Romer  1956).  The  reported  absence  of  a postpubic  process  and  the  exclusion  of 
the  pubis  from  the  acetabulum  in  Stenopelix  are  features  unknown  in  psittacosaurs,  but  have  been 
reported  for  the  pachycephalosaur  Homalocephale  (Maryanska  and  Osmolska  1974),  and  Stenopelix 
has  been  classified  in  the  Pachycephalosauridae  (Galton  1976;  Olshevsky  1978).  Contact  of  the 
ischium  with  the  pubic  peduncle  with  resultant  exclusion  of  the  pubis  from  the  acetabulum  may 
also  be  common  among  ankylosaurs,  although  obscured  in  most  specimens  by  fusion  of  the  pelvic 
elements,  and  ankylosaurs  have  a very  short  postpubic  process.  In  any  case,  assignment  of  Stenopelix 
to  the  Psittacosauridae  cannot  be  defended  on  the  basis  of  common  derived  characters,  and  the  genus 
is  here  considered  as  Ornithischia,  incertae  sedis.  The  family  Psittacosauridae  is  therefore  regarded 
as  having  a single  genus,  Psittacosaurus , with  the  type  species  P.  mongoliensis,  and  other  species 
being  of  questionable  validity. 

AMNH  6535  and  6536  come  from  the  same  stratigraphic  and  geographic  locality  as  the  type 
of  Psittacosaurus  mongoliensis.  While  the  two  new  skulls  differ  in  detail  from  the  type  of  P. 
mongoliensis,  the  differences  are  of  a kind  and  magnitude  that  are  reasonably  expected  in  such 
juvenile  specimens,  as  may  be  appreciated  by  studying  the  hypothesized  growth  series  (text-fig.  7). 
AMNH  6535  and  6536  are  therefore  assigned  to  the  species  Psittacosaurus  mongoliensis. 

Phylogenetic  position  of  the  Psittacosauridae.  Bipedality  has  long  been  the  justification  for  inclusion 
of  several  families,  including  the  Psittacosauridae,  within  the  suborder  Ornithopoda,  despite 
recognized  peculiarities  and  the  absence  of  critical  features.  Maryanska  and  Osmolska  (1974)  argued 
that  bipedality  alone  was  insufficient  evidence  for  such  a grouping,  and  removed  pachycephalosaurs 
from  the  Ornithopoda  (see  also  Olshevsky  1978).  If  the  general  dogma  that  bipedality  is  primitive 
(plesiomorphic)  for  the  order  Ornithischia  is  true  (e.g.  Galton  1978),  then  it  is  unsuitable  as  a 


100 


PALAEONTOLOGY,  VOLUME  25 


unifying  character  for  the  suborder  Ornithopoda.  Moreover,  psittacosaurs,  like  pachycephalosaurs, 
have  no  obturator  process  on  the  ischium,  a feature  apparently  synapomorphic  for  the  Ornithopoda 
sensu  stricto.  Psittacosaurs  have  been  recognized  as  ceratopsian-like  by  several  authors  (Rozhdest- 
vensky 1955,  1960;  Romer  1956;  Gregory  1957;  Young  1958;  Colbert  1965;  Maryanska  and 
Osmolska  1975)  but  Steel  (1969)  noted  that  the  absence  of  premaxillary  teeth  removed  Psittacosaurus 
from  the  immediate  ancestry  of  Protoceratops,  and  the  peculiar  structure  of  the  Psittacosaurus 
manus  (Osborn  1924)  is  a further  barrier  to  placing  the  genus  in  a position  ancestral  to  any  known 
ceratopsian. 

Maryanska  and  Osmolska  (1975)  recognized  that  the  substantial  similarity  in  general  cranial 
morphology  shared  by  psittacosaurs  and  ceratopsians  outweighed  any  arguments  on  the  unsuit- 
ability of  the  former  as  ‘ancestors’  for  the  latter.  They  also  realized  that  failure  to  identify  a rostral 
bone  in  Psittacosaurus  was  the  single  major  objection  to  transferring  the  Psittacosauridae  to  the 
suborder  Ceratopsia,  and  they  boldly  suggested  that  the  snout  of  Psittacosaurus  had  been  incorrectly 
interpreted  by  Osborn  (1923,  1924)  and  others  (Young  1958;  Chao  1962).  Maryanska  and  Osmolska 
(1975)  contended  that  a rostral  hone  is  in  fact  present  in  Psittacosaurus , an  opinion  with  which  I 
fully  concur  (see  also  Olshevsky  1978). 

The  rostral  of  ceratopsians  probably  originated  as  an  epidermally  induced  ossification,  with  the 
large,  overlying  epidermal  scale  eventually  becoming  a thick,  heavily  keratinized  upper  beak.  The 
rostral  and  upper  beak  thus  correspond  to  the  predentary  and  associated  beak  of  the  lower  jaw 
in  origin  and  function.  In  Psittacosaurus  the  rostral  is  ill  defined  and  in  large  part  fused  to  the 
premaxillae,  similar  to  the  way  in  which  epidermal  ossifications  overlie  and  are  fused  to  the  skull 
roof  of  ankylosaurs  (e.g.  Maryanska  1971,  1977).  The  Psittacosaurus  rostral  is  not,  therefore,  a 
completely  separate  cranial  element  as  it  is  in  advanced  ceratopsians,  but  it  is  nevertheless  present. 
Revised  identifications  of  psittacosaur  skull  bones  (text-figs.  3 and  7c)  may  be  compared  with 
identifications  in  original  descriptions. 

The  family  Psittacosauridae  is  here  regarded  as  belonging  in  the  suborder  Ceratopsia  rather 
than  in  the  Ornithopoda,  but  further  refining  of  its  position  among  ceratopsians  is  difficult.  Many 
taxa  included  in  the  Protoceratopsidae  share  synapomorphies  with  the  Ceratopsidae,  for  example 
the  development  of  a posterior  parietal-squamosal  frill  and  at  least  incipient  development  of 
horns,  features  not  present  in  psittacosaurs.  On  the  other  hand,  absence  of  premaxillary  teeth  in 
Psittacosaurus  is  possibly  a derived  character  shared  with  the  Ceratopsidae,  but  not  with  the 
Protoceratopsidae  some  of  which  have  premaxillary  teeth.  The  peculiar  manus  of  Psittacosaurus, 
and  the  related  bipedal  stance,  are  unique  developments  among  ceratopsians,  but  whether  these 
are  derived  characters  defining  the  basal  sister  group  split  of  the  suborder,  or  merely  retention  of 
primitive  features  by  Psittacosaurus,  is  unclear.  Potentially,  at  least,  the  Psittacosauridae  could  be 
the  sister  group  of  all  other  Ceratopsia,  such  a position  leaving  uncertain  the  question  as  to  whether 
bipedal  or  quadrupedal  posture  was  the  primitive  condition  for  the  suborder. 

Juvenile  features.  Differences  between  adult  and  juvenile  ceratopsian  skulls  (Brown  and  Schlaikjer 
1940;  Maryanska  and  Osmolska  1975;  Dodson  1976)  provide  a basis  for  recognition  of  juvenile 
features  in  the  two  Psittacosaurus  skulls  under  consideration.  Among  these  features  are  the  following: 
orbit  very  large;  braincase  relatively  large;  skull  roof  more  curved  than  in  adult;  snout  short;  rostral 
approaches  and  may  contact  maxilla,  thus  excluding  premaxilla  from  border  of  mouth;  small, 
narrow  lateral  temporal  fenestrae;  no  sagittal  crest;  suborbital  bar  slender;  and  jugal  flange  small 
or  absent.  General  changes  in  cranial  proportions  can  be  seen  in  the  hypothetical  ontogenetic  series 
(text-fig.  7).  Quantification  of  allometric  relationships  is  not  justified  in  view  of  the  small  sample 
for  each  age  category.  Consideration  of  juvenile  features  in  AMNH  6535  and  6536  serves  less  for 
establishment  of  growth  trends  than  for  confirmation  of  these  specimens  as  young  Psittacosaurus 
mongoliensis  and  not  a new  taxon  of  small  dinosaur. 

Smallest  dinosaur.  AMNH  6535  and  6536  are  each  smaller  than  any  previously  reported  dinosaur 
specimen.  A listing  of  the  smallest  dinosaurs  based  on  basal  skull  length  yields  the  following 


COOMBS:  JUVENILE  DINOSAURS 


101 


text-fig.  7.  A hypothetical  ontogenetic  series  of  Psittacosaurus  skulls  based  upon  a,  AMNH  6535;  b,  AMNH 
6536;  c,  the  type  of  P.  youngi  (after  Chao  1962);  and  d,  AMNH  6254,  the  type  of  P.  mongoliensis  (after 
Osborn  1923).  Readily  observed  are  a decrease  in  relative  size  of  the  orbit,  a lengthening  of  the  snout  (pre-orbital 
region  of  skull),  and  a shift  from  a rounded  to  a flat-topped  cranial  contour.  Revised  identifications  of  some 
cranial  elements  are  given  in  c.  Abbreviations:  J,  jugal;  l,  lacrimal;  mx,  maxilla;  n,  nasal;  pm,  premaxilla;  pp, 

palpebral;  and  r,  rostral. 


102 


PALAEONTOLOGY,  VOLUME  25 


sequence:  Lesothosaurus  diagnosticus,  94  mm  (Galton  1978;  = Fabrosaurus  australis  of  Thulborn 
1970);  Protoceratops  andrewsi,  76  mm  and  62  mm  (respectively  AMNH  6419  and  ZPAL  MgD  II/7; 
Brown  and  Schlaikjer  1940;  Dodson  1976;  Maryanska  and  Osmolska  1975);  Compsognathus 
longipes,  70  to  75  mm  (Huene  1925;  Ostrom  1978);  Bagaceratops  rozhdestvenskyi,  47  mm  (ZPAL 
MgD  1/123;  Maryanska  and  Osmolska  1975);  juvenile  prosaurauropod(?),  32  mm  (Charig  1979); 
Psittacosaurus  mongoliensis  42  mm  and  28  mm  (respectively  AMNH  6536  and  6535).  An  estimate 
of  total  body  length  for  the  two  Psittacosaurus  juveniles  can  be  obtained  by  scaling  down  the 
dimensions  of  the  two  large,  almost  complete  skeletons  described  by  Osborn  (1923,  1924),  using 
comparisons  of  median  skull  length.  The  calculated  snout  to  tail  tip  length  for  AMNH  6536  is 
about  390  mm,  and  that  for  AMNH  6535  is  about  265  mm.  Possible  allometric  changes  in  skull 
size  relative  to  body  length  were  not  considered  in  these  estimates  but,  since  the  skull  is 
proportionately  larger  in  juveniles,  the  calculated  total  body  lengths  probably  err  on  the  high  side. 
Scaling  the  complete  Psittacosaurus  skeleton  based  on  a comparison  of  humeral  length  (using  the 
humerus  shown  in  fig.  5)  gives  an  estimated  total  length  for  AMNH  6536  of  340  mm.  Allowing 
for  some  allometric  changes  in  skull  size,  I would  estimate  the  total  body  length  of  AMNH  6535 
at  about  230  mm.  For  comparison  with  the  size  of  these  two  juvenile  psittacosaurs,  the  famous 
skeleton  of  Compsognathus  longipes , also  probably  a juvenile,  has  a length  of  between  750  and 
810  mm  (based  on  restorations  by  von  Huene  1925,  1926;  Ostrom  1978).  The  smaller  Psittacosaurus 
specimen  described  herein  (AMNH  6535)  belonged  to  a dinosaur  slightly  smaller  than  a common 
pigeon  ( Columba  livia). 

Parental  care.  Recently  it  has  become  fashionable  to  ascribe  mammalian  or  avian  behaviour  patterns 
to  dinosaurs  despite  the  reptilian  aspect  of  the  brains  of  all  dinosaurs  except  small  theropods 
(Hopson  1977)  and  the  acknowledged  difficulty  of  deducing  habits  from  osteology  even  for  living 
animals.  Among  behavioural  patterns  considered  for  dinosaurs  is  parental  care  of  young  (e.g. 
Horner  and  Makela  1979).  In  attempting  an  objective  analysis  of  parental  care  in  dinosaurs,  three 
points  must  be  kept  in  mind.  First,  there  is  no  such  thing  as  an  ‘accepted  interpretation’.  The 
suggestion  that  dinosaurs  abandoned  their  eggs  as  do  modern  chelonians  is  a hypothesis  that 
requires  proof  as  rigorous  as  that  required  to  prove  that  dinosaurs  cared  for  their  young  in  the 
manner  of  modern  ducks  or  ungulates.  Casting  aspersions  on  arguments  favouring  parental  care 
does  not  constitute  proof  of  the  alternate  hypothesis.  Second,  the  morphologic  diversity  and  long 
evolutionary  history  of  dinosaurs  makes  it  exceedingly  unlikely  that  every  species  practiced  a similar 
amount  of  parental  care.  Indeed,  the  egg-laying  pattern  of  the  hadrosaur  Rhabdodon  has  been 
interpreted  as  precluding  parental  care  (Ginsburg  1980)  while  nests  of  the  hadrosaur  Maiasaurus 
have  been  interpreted  as  requiring  parental  care  (Horner  and  Makela  1979).  Dinosaur  parental 
behaviour  should  be  approached  on  a case-by-case  basis,  and  generalities  applicable  to  all  dinosaurs 
will  necessarily  be  few.  Finally,  the  selection  of  modern  analogues  to  interpret  dinosaur  behaviour 
is  a subtle  trap  that  encourages  interpretations  far  beyond  what  can  reasonably  be  concluded  from 
actual  data.  This  problem  has  been  discussed  elsewhere  (Coombs  1975),  but  the  warning  merits 
repetition.  The  mental  image  of  the  relationship  between  adult  and  juvenile  dinosaur  is  greatly 
altered  if  described  as  being:  (1)  like  elephants;  (2)  like  opossums;  (3)  like  ducks;  (4)  like  crocodiles; 
or  (5)  like  mouth-breeding  fishes.  Use  of  modern  analogues  must  be  approached  with  great 
caution. 

In  the  absence  of  contrary  evidence  it  is  assumed  that  most  dinosaurs  laid  eggs  (Hopson  1977). 
Coelophysis  specimens  from  Ghost  Ranch,  New  Mexico  (Colbert  1961),  are  sometimes  cited  as 
demonstrating  dinosaurian  viviparity,  but  are  more  likely  a case  of  cannibalism  in  so  far  as  jumbled, 
disarticulated  juvenile  skeletons  are  enclosed  in  almost  perfectly  articulated  adult  skeletons  (making 
parental  care  unlikely  for  Coelophysis ).  The  supposed  juvenile  within  the  type  of  Compsognathus 
longipes  is  in  fact  a skeleton  of  the  lacertilian  Bavarisaurus  (Ostrom  1978).  Dinosaur  egg  morphology 
indicates  nests  buried  in  leaf  litter  or  sand  (Seymour  1979;  Seymour  and  Ackerman  1980),  a 
construction  pattern  similar  to  that  of  both  crocodilians  and  megapode  birds  (Frith  1956,  1962; 
Seymour  and  Ackerman  1980).  Megapode  nesting  habits  are  regarded  as  primitive  among  birds 


COOMBS:  JUVENILE  DINOSAURS 


103 


(Welty  1963)  and  are  thus  an  appropriate  analogue  for  at  least  the  primitive  pattern  of  dinosaur 
nest-building  behaviour.  Parental  attendance  of  buried  nests,  involving  protection  (crocodilians), 
temperature  regulation  (megapodes),  and  assistance  to  young  at  hatching  (both  crocodilians  and 
megapodes)  may  be  a primitive  archosaurian  behaviour  pattern  that  is  reasonably  inferred  for 
dinosaurs.  Hatchlings  of  crocodilians,  megapodes,  and  ground-nesting  birds  in  general  (excluding 
those  relatively  free  of  terrestrial  predator  threat,  e.g.  most  insular,  Arctic,  and  Antarctic  marine 
birds)  are  generally  precocial  (mobile,  self-feeding,  and  fast  maturing).  Megapode  hatchlings  are 
notoriously  precocial,  at  least  one  species  being  capable  of  flight  within  24  hours  of  hatching  (Frith 
1962).  The  Psittacosaurus  juveniles  had  been  feeding  on  abrasive  material,  probably  vegetation, 
for  some  time  prior  to  death,  as  evidenced  by  wear  on  teeth  of  both  skulls  (similar  conclusion  for 
juvenile  Maiasaurus : Horner  and  Makela  1979).  It  is  therefore  probable  that  the  juveniles  of 
Psittacosaurus  were  precocial;  parental  offering  of  transported,  premasticated,  or  regurgitated  food 
is  unlikely,  implying  that  parental  care  was  unnecessary  at  least  for  feeding. 

A great  difference  in  physical  size  between  juvenile  and  adult  increases  liability  of  injury  to 
young  in  species  having  large  adults  and  parental  care  after  hatching.  Habits  and  posture  also 
play  a part.  Some  crocodilians  practice  extended  parental  care  after  the  young  hatch,  but  crocodilians 
are  semi-aquatic,  with  young  more  aquatic  than  adults,  and  the  posture  is  sprawling  to  semi-erect 
quadrupedal  with  long,  slender,  flexible  toes.  The  danger  of  injury  to  a juvenile  crocodilian  is  thus 
far  less  than  to  juveniles  of  terrestrial,  bipedal  dinosaurs  that  have  a compact  foot  and  digitigrade 
stance.  Among  both  mammals  and  birds  that  are  ground  dwellers  and  that  have  precocial  young, 
the  minimum  relative  newborn  size  ( = newborn  body  mass  expressed  as  a percent  of  adult  body 
mass,  and  hereafter  abbreviated  RNS)  is  about  0-9%  (in  Struthio\  Welty  1963).  RNS  values  for 
mammals  with  precocial  young  range  from  about  1 to  12%  (calculated  from  data  in  Walker 
1968)  with  typical  values  of  about  3-5  % (e.g.  Odocoileus,  Giraffa , Rhinoceros , Antilocapra , Rangifer, 
Taurotragus,  and  Loxodonta : Walker  1968;  Case  1978).  Adult  body  size  has  little  influence  on  RNS 
for  mammals  with  precocial  young.  Relatively  immobile,  altricial  young  whose  position  and 
movements  are  either  limited  or  controlled  by  the  parent  have  much  lower  RNS  values  (e.g.  0-33  % 
in  Euarctos : Walker  1968;  Case  1978).  The  recent  death  of  a zoo-born  Ailuropoda,  apparently  by 
smothering  under  a careless  (inexperienced)  adult  female,  points  up  the  potential  dangers  to  young 
of  very  low  RNS  and  terrestrial  habits.  RNS  values  of  modern  reptiles  follow  a different  pattern: 
values  decline  as  adult  body  mass  increases  (Table  1).  The  RNS  of  Psittacosaurus  is  0-7  to  0-8% 
(calculated  by  cubing  equivalent  linear  dimensions  of  adult  and  juvenile)  which  is  low  compared 
to  precocial  young  of  mammals  or  birds,  but  is  high  compared  to  that  of  modern  reptiles  of  similar 
adult  body  mass  (10  to  100  kg  category.  Table  1).  RNS  values  for  other  dinosaurs  include  0-24% 
for  Protoceratops  and  0 06%  for  Hypselosaurus  (Case  1978).  Thus  dinosaurs  appear  to  follow  the 
reptilian  pattern  of  declining  RNS  at  high  adult  body  mass.  The  combination  of  low  RNS  and 
mobile,  self-feeding  habits  of  Psittacosaurus  contrasts  with  the  pattern  for  mammals  and  birds  that 
practice  parental  care  after  birth  or  hatching  and  that  have  active,  mobile  young.  On  the  basis  of 
RNS  values,  parental  care  is  marginally  possible,  but  not  probable  for  Psittacosaurus,  and  is  very 
unlikely  for  dinosaurs  of  high  adult  body  mass. 


table  1.  Data  from  Case  (1978)  showing  the  decline  in  relative  newborn  size  (RNS)  at  higher  adult  body 
masses  among  modern  reptiles. 


Adult  mass 

Number  of  species 
averaged 

Average  RNS 

Less  than  1 kg 

6 

3-8% 

1-10  kg 

7 

1-9% 

10-100  kg 

5 

0-42% 

Over  100  kg 

7 

0-04% 

104 


PALAEONTOLOGY,  VOLUME  25 


Indirect  support  for  the  hypothesis  that  some  dinosaurs  did  not  care  for  their  young  after 
hatching  comes  from  Triassic- Jurassic  trackways  known  as  Selenichnus  (text-fig.  8).  The  footprint 
maker  was  a tridactyl  biped  with  a tail  sufficiently  long  to  occasionally  imprint,  and  the  animal 
was  probably  a juvenile  of  one  of  the  many  large  bipedal  dinosaurs  that  are  so  ubiquitous  in  the 
Connecticut  Valley  (e.g.  Eubrontes , Gigandipus,  Anomoepus,  Anchisauripus,  or  Sauropus;  Lull  1953). 
The  Selenichnus  trackmaker  was  intermediate  in  size  between  the  two  juvenile  Psittacosaurus 
described  herein,  although  in  view  of  the  disparity  in  geologic  age,  Selenichnus  trackways  could 
not  belong  to  Psittacosaurus.  The  five  slabs  at  the  Pratt  Museum  (Amherst  College,  Amherst, 
Massachusetts,  U.S.A.)  that  contain  Selenichnus  tracks  have  the  following  characteristics  in 
common:  (1)  there  are  no  instances  of ‘adult’  footprints  in  company  with  Selenichnus ; and  (2)  there 
are  no  instances  of  two  or  more  Selenichnus  tracks  on  a single  bedding  plane,  headed  in  roughly 
the  same  direction.  Insufficient  numbers  of  Selenichnus  tracks  are  known  to  draw  unequivocal 
conclusions,  but  the  evidence  indicates  that  tiny  dinosaurs,  the  size  of  small  juvenile  Psittacosaurus, 
travelled  alone,  unaccompanied  by  either  adults  or  fellow  hatchlings,  at  least  sometimes.  This 
evidence  suggests  that  at  least  some  dinosaurs  did  not  practice  post-hatching  parental  care. 


text-fig.  8.  Selenichnus  breviusculus,  trackway  made  by  a small  tridactyl,  bipedal  dinosaur  that  at  least 
sometimes  left  a sinuous  tail  trace  (pseudo-tail  traces  are  sometimes  made  by  dragging  the  tip  of  one  toe 
through  soft  sediment,  but  that  does  not  appear  to  be  the  case  with  this  specimen).  This  particular 
trackway  was  made  on  a raindrop  splattered  surface.  According  to  Lull  (1953)  the  average  print  length 
for  this  species  is  46  mm  and  the  average  step  length  is  58  mm.  I estimate  the  size  of  the  trackmaker 
to  be  a little  smaller  than  AMNH  6536,  the  larger  Psittacosaurus  specimen  described  herein.  Lull  (1953) 
regarded  these  tracks  as  pertaining  to  a theropod.  Specimen  in  the  collection  of  the  Pratt  Museum, 
Amherst  College  (Amherst,  Massachusetts,  U.S.A.). 


Association  of  adult  and  juvenile  skeletons  might  be  taken  as  an  indication  of  parental  care. 
The  remarkable  assemblage  of  Protoceratops  andrewsi  at  Djadochta  includes  numerous  individuals 
of  every  age  category,  unhatched  eggs  to  old  adults  (Brown  and  Schlaikjer  1940;  Dodson  1976) 
and  association  of  adults  and  juveniles  in  life  is  thus  indisputable.  Horner  and  Makela  (1979) 
interpreted  a group  of  juvenile  hadrosaurs  ( Maiasaurus ) as  requiring  parental  guidance  to  remain 
a coherent  group,  and  an  adult  specimen  was  unearthed  reasonably  nearby.  As  noted  above,  the 
juvenile  Psittacosaurus  described  herein  were  excavated  in  the  vicinity  of  the  type  of  P.  mongoliensis, 
although  exactly  how  close  is  difficult  to  ascertain  from  available  records.  Association  of  adults 
and  juveniles  in  fossil  assemblages  has  little  bearing  on  the  question  of  parental  care  in  so  far  as 
such  an  association  would  sometimes  occur  even  if  the  adults  abandoned  their  eggs  immediately 
after  laying.  Only  if  nests  were  constructed  far  from  the  normal  range  of  the  adult  (note  arguments 
of  Horner  and  Makela  1979),  and  if  the  juveniles  lived  entirely  outside  of  adult  habitats  would 
the  incidence  of  adult-juvenile  association  in  fossil  assemblages  be  reduced  to  nil.  While  such  a 
set  of  conditions  may  have  existed  for  some  dinosaurs,  it  is  unlikely  it  was  a universal  situation. 


COOMBS:  JUVENILE  DINOSAURS 


105 


Thus  the  association  of  adult  and  juvenile  skeletons  even  in  a single  quarry  is  not  sufficient  proof 
of  extended  post-hatching  care. 

Parental  care,  summary.  Available  evidence  for  parental  care  by  dinosaurs  is  scant  and  equivocal. 
Moreover,  there  is  no  clear  indication  of  what  constitutes  necessary  and  sufficient  proof  of  either 
competing  theory  (i.e.  that  dinosaurs  did,  or  that  dinosaurs  did  not,  care  for  their  young).  On  the 
basis  of  the  data  and  arguments  developed  above,  the  following  conclusion  might  be  drawn: 
(1)  guarding  of  nests  may  be  a general  archosaurian  pattern  inherited  as  the  primitive  behaviour 
of  dinosaurs;  (2)  nests  of  some  dinosaurs  were  buried  and  may  have  required  parental  attendance 
for  temperature  regulation  and  perhaps  for  assisting  the  young  at  hatching;  (3)  egg  distribution 
of  other  dinosaurs  seems  to  preclude  parental  care  after  laying;  (4)  adult-juvenile  size  disparity 
makes  parental  care  unlikely  for  very  large  dinosaurs,  but  for  smaller  genera  the  size  disparity 
may  not  have  been  a problem;  (5)  footprints  indicate  that  at  least  some  tiny  dinosaurs,  probably 
juveniles,  were  independent  of  adults;  and  (6)  for  Psittacosaurus,  available  evidence  is  inconclusive. 

Sibling  groups.  One  of  the  two  juvenile  Psittacosaurus  specimens  is  a composite  of  several  skeletons, 
with  fourteen  distal  ends  of  tibiae  indicating  a minimum  of  seven  individuals  represented.  The 
jumbled,  broken,  incomplete  condition  of  the  skeletons  suggests  a post-mortem  assemblage,  yet 
the  majority  of  bones  are  from  juveniles  of  similar,  indeed  almost  identical,  size  (text-fig.  6).  Such 
an  assemblage  could  result  from  mechanical  sorting  processes,  or  from  the  simultaneous  deaths 
of  several  individuals  of  a sibling  group.  Horner  and  Makela  (1979)  described  an  assemblage  of 
juvenile  hadrosaurs  probably  derived  from  a single  clutch,  and  footprints  possibly  belonging  to 
juvenile  dinosaurs  travelling  in  unison  have  been  reported  (Currie  and  Sargeant  1979;  contra 
Selenichnus  evidence  described  above).  Therefore  juveniles,  possibly  clutch  mates  of  some  dinosaurs, 
may  have  formed  cohesive  aggregates,  for  which  the  term  ‘sibling  group’  is  proposed  in  preference 
to  ‘herd’  or  ‘flock’  or  ‘school’  all  of  which  imply  behavioural  interactions  not  yet  proven  for 
dinosaurs.  A behavioural  complexity  similar  to  that  of  modern  gregarious  reptiles  (e.g.  Ambly- 
rhynchus)  is  sufficient  to  explain  sibling  groups  among  dinosaurs.  Aggregates  of  adult  dinosaurs 
might  have  resulted  from  persistence  of  bonds  formed  in  juvenile  sibling  groups. 


REFERENCES 

bohlin,  b.  1953.  Fossil  reptiles  from  Mongolia  and  Kansu.  Palaeont.  sin.  Publ.  37,  1-113. 
brown,  b.  1914.  Leptoceratops,  a new  genus  of  Ceratopsia  from  the  Edmonton  Cretaceous  of  Alberta.  Bull. 
Am.  Mus.  nat.  Hist.  33,  567-580. 

— and  schlaikjer,  e.  m.  1940.  The  structure  and  relationships  of  Protoceratops.  Ann.  N.Y.  Acad.  Sci.  40, 
133-266. 

— 1942.  The  skeleton  of  Leptoceratops  with  the  description  of  a new  species.  Am.  Mus.  Novit.  No.  1 169, 
1-15. 

case,  t.  j.  1978.  Speculations  on  the  growth  rate  and  reproduction  of  some  dinosaurs.  Paleobiology,  4, 
320-328. 

chao,  s.  1962.  A new  species  of  psittacosaurs  from  Laiyang,  Shantung.  Vertebr.  Palasiat.  6,  349-360.  [In 
Chinese.] 

charig,  a.  1979.  A new  look  at  the  dinosaurs.  Mayflower  Books,  Inc.,  New  York.  160  pp. 
colbert,  e.  h.  1961.  Dinosaurs,  their  discovery  and  their  world.  E.  P.  Dutton  and  Co.,  New  York,  xiv  + 300  pp., 
51  figs.,  100  pis. 

— 1965.  The  age  of  reptiles.  W.  W.  Norton  & Co.,  Inc.,  New  York.  228  pp.,  67  figs.,  20  pis. 
coombs,  w.  p.,  jr.  1972.  The  bony  eyelid  of  Euoplocephalus  (Reptilia,  Ornithischia).  J.  Paleont.  46,  637-650. 

— 1975.  Sauropod  habits  and  habitats.  Palaeogeogr.,  Palaeoclimatol,  Palaeoecol , 17,  1-33. 

currie,  p.  j.  and  sargeant,  w.  a.  s.  1979.  Lower  Cretaceous  dinosaur  footprints  from  the  Peace  River  Canyon, 
British  Columbia,  Canada.  Ibid.  28,  103-115. 

dodson,  p.  1976.  Quantitative  aspects  of  relative  growth  and  sexual  dimorphism  in  Protoceratops.  J.  Paleont. 
50,  929-940. 


106 


PALAEONTOLOGY,  VOLUME  25 


frith,  h.  j.  1956.  Breeding  habits  in  the  family  Megapodiidae.  Ibis,  98,  620-640. 

1962.  The  mallee  fowl.  Angus  and  Robertson,  Sidney. 

galton,  p.  m.  1973.  On  the  cheeks  of  ornithopod  dinosaurs.  Lethaia,  6,  67-89. 

— 1976.  The  dinosaur  Vectisaurus  valdensis  (Ornithischia:  Iguanodontidae)  from  the  Lower  Cretaceous  of 
England.  J.  Paleont.  50,  976-984. 

— 1978.  Fabrosauridae,  the  basal  family  of  ornithischian  dinosaurs  (Reptilia:  Ornithopoda).  Paldont.  Z.  52, 
138-159. 

ginsburg,  l.  1980.  Les  gisements  a oeufs  de  Dinosaures  du  Cretace  terminal  du  Midi  de  la  France  et  la 
physiologie  des  Ornithopodes.  Mem.  Soc.  geol.  France,  N.S.,  139,  109-110. 

Gregory,  w.  k.  1951.  Evolution  emerging.  Macmillan  Co.,  New  York.  Vol.  1,  xxvi  + 736  pp.  (text),  vol.  2, 
viii  + 1013  pp.  (figures  and  plates). 

hopson,  j.  a.  1977.  Relative  brain  size  and  behaviour  in  archosaurian  reptiles.  Annu.  Rev.  Ecol.  Syst.  8, 429-448. 
horner,  j.  r.  and  makela,  r.  1979.  Nest  of  juveniles  provides  evidence  of  family  structure  among  dinosaurs. 
Nature,  Lond.  282,  296-298. 

huene,  F.  von  1908.  Die  Dinosaurier  der  Europaischen  Triasformation  mit  Beriicksichtigung  der  aussereuro- 
paischen  Vorkomnnisse.  Geol.  paldont.  Abh.  Suppl.  1,  xii  + 419  pp.,  351  figs.,  pis.  1-111. 

— 1909.  Skizze  zu  einer  Systematik  und  Stammesgeschicte  der  Dinosaurier.  Zentbl.  Miner.  Geol.  Paldont. 
1909,  12-22. 

— 1925.  Eine  neue  Rekonstruktion  von  Compsognathus.  Ibid.  B 5,  157-160. 

— 1926.  The  carnivorous  Saurischia  in  the  Jura  and  Cretaceous  formations,  principally  in  Europe.  Rev.  Mus. 
La  Plata,  29,  35-167. 

huxley,  t.  h.  1870.  On  the  classification  of  the  Dinosauria  with  observations  on  the  Dinosauria  of  the  Trias. 
Q.  Jl  geol.  Soc.  Lond.  26,  32-50. 

koken,  e.  1887.  Die  Dinosaurier,  Crocodiliden,  und  Sauropterygier  des  norddeutschen  Wealden.  Palaeont.  Abh. 
3,  309-419. 

lull,  r.  s.  1910.  Dinosaurian  distribution.  Am.  J.  Sci.  29,  1-39. 

— 1933.  A revision  of  the  Ceratopsia  or  horned  dinosaurs.  Mem.  Peabody  Mus.  nat.  Hist.  3,  1-135. 

— 1953.  Triassic  life  of  the  Connecticut  Valley  (revised  edition).  Conn.  Geol.  Nat.  hist.  Surv.,  Bull.  81, 
1-331. 

maleev,  e.  a.  1954.  (Armored  dinosaurs  from  the  Upper  Cretaceous  of  Mongolia).  Dokl.  Ark.  Nauk  S.S.S.R. 
87,  131-134.  [In  Russian.] 

maryanska,  t.  1971.  New  data  on  the  skull  of  Pinacosaurus  grangeri  (Ankylosauria).  Palaeont.  pol.  25,  45-53. 
1977.  Ankylosauridae  (Dinosauria)  from  Mongolia.  Ibid.  37,  85-151. 

— and  osmolska,  h.  1974.  Pachycephalosauria,  a new  suborder  of  ornithischian  dinosaurs.  Ibid.  30,  45-102. 
1975.  Protoceratopsidae  (Dinosauria)  of  Asia.  Ibid.  33,  133-182. 

meyer,  h.  von  1857.  Stenopelix  valdensis,  ein  reptil  aus  der  Wealden-Formation  Deutschlands.  Palaeonto- 
graphica,  7,  25-34. 

Olshevsky,  G.  1978.  The  archosaurian  taxa  (excluding  the  Crocodylia).  Mesozoic  Meanderings,  1,  1-50. 
osborn,  h.  f.  1923.  Two  lower  Cretaceous  dinosaurs  of  Mongolia.  Am.  Mus.  Novit.  No.  95,  1-10. 

— 1924.  Psittacosaurus  and  Protiguanodon:  two  lower  Cretaceous  iguanodonts  from  Mongolia.  Ibid.  127, 
1-16. 

ostrom,  j.  h.  1978.  The  osteology  of  Compsognathus  longipes  Wagner.  Zitteliana,  4,  73-118. 
romer,  a.  s.  1956.  Osteology  of  the  reptiles.  Univ.  Chicago  Press,  Chicago,  xxi  + 772  pp.,  248  figs. 
Rozhdestvensky,  a.  k.  1955.  (New  data  concerning  psittacosaurs— Cretaceous  ornithopods).  Vop.  Geol.  Asia, 
2,  783-788.  [In  Russian.] 

— 1960.  (Locality  of  Lower  Cretaceous  dinosaurs  in  Khuzbass).  Palaeont.  Jour.  2,  1965.  [In  Russian.] 
russell,  L.  s.  1940.  Edmontonia  rugosidens  (Gilmore),  an  armored  dinosaur  from  the  Belly  River  series  of 
Alberta.  Univ.  Toronto  Stud.  geol.  Ser.  43,  3-28. 

Seymour,  r.  s.  1979.  Dinosaur  eggs:  gas  conductance  through  the  shell,  water  loss  during  incubation  and 
clutch  size.  Paleobiology,  5,  1-11. 

— and  ackerman,  R.  A.  1980.  Adaptations  to  underground  nesting  in  birds  and  reptiles.  Am.  Zool.  20, 
437-447. 

steel,  R.  1969.  Ornithischia.  In  Handbuch  der  Palaoherpetologie  (O.  Kuhn,  ed.),  part  15,  84  pp.,  24  figs. 
swinton,  w.  e.  1936.  Notes  on  the  osteology  of  Hypsilophodon  and  on  the  family  Hypsilophodontidae.  Proc. 
zool.  Soc.  Lond.  1936,  555-578. 

thulborn,  r.  a.  1970.  The  skull  of  Fabrosaurus  australis,  a Triassic  ornithischian  dinosaur.  Palaeontology, 
13,  414-432. 


COOMBS:  JUVENILE  DINOSAURS 


107 


walker,  e.  p.  1968.  Mammals  of  the  world.  Johns  Hopkins  Press,  Baltimore,  xlviii  + 1500  pp.  (in  two  volumes). 
welty,  j.  c.  1963.  The  life  of  birds.  Alfred  A.  Knopf,  New  York,  xiii  + 546  pp. 

young,  c.  c.  1931.  On  some  new  dinosaurs  from  western  Suiyan,  Inner  Mongolia.  Bull.  geol.  Surv.  China , 11, 
159-266. 

— 1958.  The  dinosaurian  remains  of  Laiyang,  Shantung.  Paleont.  Sin.  (n.s.),  16,  1-139.  [In  Chinese  and 


English.] 


Typescript  received  3 March  1980 
Revised  typescript  received  5 December  1980 


WALTER  P.  COOMBS,  JR. 
Western  New  England  College 
Springfield,  Mass.  01119,  U.S.A. 


A DESCRIPTION  OF  THE  GENERATING  CURVE 
OF  BIVALVES  WITH  STRAIGHT  HINGES 

by  MARGARET  JENNIFER  ROGERS 


Abstract.  A method  of  describing  the  whole  of  the  generating  curve  of  a lamellibranch  is  sought.  When 
lengths  and  angles  are  used  to  describe  an  outline,  much  of  the  outline  remains  undefined.  A curve  can  be 
fitted  to  an  outline  and  the  coefficients  in  the  particular  approximation  employed  then  define  the  outline. 
Reasons  are  given  for  fitting  a Tchebychev  polynominal,  rather  than  a spline,  or  Fourier  series  containing 
both  sine  and  cosine  terms.  Polar  coordinates  r and  6 are  calculated  for  each  point  on  a digitized  outline. 
Cos  6,  rather  than  6 is  chosen  as  the  independent  variable  when  the  Tchebychev  coefficients  are  calculated. 
It  is  found  that  about  100  irregular  spaced  data  points  are  required  to  produce  stable  coefficients,  and  that 
adequate  numerical  accuracy  is  obtained  when  the  outline  is  described  by  the  first  six  coefficients.  These  six 
coefficients  can  also  be  used  as  shape  discriminators.  As  the  value  of  c0  is  a measure  of  size,  it  can  be  used 
to  standardize  the  other  coefficients.  The  standardized  coefficients  can  be  used  to  compare  the  shape  of  the 
generating  curve  of  shells  of  different  sizes. 

Descriptions  of  the  morphology  of  organisms  should,  ideally,  be  objective,  reproducible,  and 
enable  different  forms  to  be  distinguished.  In  this  article  the  description  of  the  lateral  outline,  or 
generating  curve  (Raup  1966)  of  the  Carboniferous  non-marine  bivalve  genera  Naiadites  and 
Curvirimula  is  discussed.  The  general  form  of  the  generating  curve  is  elliptical.  A point  (the  origin 
of  growth,  or  tip  of  the  umbo)  and  a line  (the  hinge-line,  which  is  straight,  or  only  slightly  curved 
in  the  genera  studied)  define  the  orientation  of  these  shells  to  which  further  descriptions  can  be 
related.  A coarse  separation  of  shapes  can  be  made  qualitatively  by  simple  terms  such  as  outline 
approximately  triangular  (text-fig.  la),  rectangular  (text-fig.  lb),  or  semicircular  (text-fig.  lc,  d).  The 
distinction  between  text-fig.  lc  and  text-fig.  Id  is  more  difficult  to  describe  qualitatively.  Measure- 
ments of  various  lengths  and  angles  have  proved  useful  in  describing  and  distinguishing  shell  forms 


text-fig.  1 . Outlines  of  four  shells  which  can  be  described  qualitatively  as 
approximately  (a)  triangular,  ( b ) rectangular,  (c  and  d ) semicircular. 

(Davies  and  Trueman  1927;  Deleers  and  Pastiels  1947;  Trueman  and  Weir  1955;  Eager  1973;  Hajkr, 
Lukasova,  Ruzicka,  and  Rehor  1974),  but  only  a few  points  on  the  shell  outline  are  defined.  As 
Bookstein  (1978)  notes,  techniques  which  use  ‘landmarks’,  inter-landmark  distances,  and  angles 
ignore  much  of  the  shape  of  the  outline.  Between  the  landmarks  the  outline  is  undefined  and  cannot 
be  reproduced. 

Papin  and  Khoroshev  (1974)  measured  the  radius  of  curvature  (text-fig.  2,  rc;)  of  successive  parts 
of  the  outline,  indicating  that  sections  of  the  curve  should  be  described,  rather  than  interpoint 
distances.  The  value  of  the  radius  of  curvature  at  Ph  when  calculated  from  the  position  of  three 
successive  points  P,  1?  P;,  P/+1  (text-fig.  2)  on  the  curve,  is  highly  sensitive  to  the  actual  position 
of  the  points  Pi_l,  Pi  + i.  Accordingly,  it  is  virtually  impossible  to  obtain  a reliable  estimate  of  this 


[Palaeontology,  Vol.  25,  Part  1,  1982,  pp.  109-117] 


110 


PALAEONTOLOGY,  VOLUME  25 


text-fig.  2.  Measurements  on  a shell  outline.  The  hinge 
OA  is  produced  to  B.  Pi^1,  Ph  Pi+l  are  three  consecutive 
points  on  the  outline.  rc(  is  the  radius  of  curvature,  t;  the 
tangent  angle,  and  r,  the  distance  from  0 of  point  Pt. 


quantity.  For  this  reason  the  radius  of  curvature,  and  a related  measurement,  the  tangent  angle  f, 
(text-fig.  2)  are  not  suitable  as  shape  descriptors.  Error  in  the  measurement  of  the  length  r,  (text-fig. 
2),  from  the  umbo  to  the  point  is  due  to  the  accuracy  of  the  measuring  device  used  and  is  not 
related  to  the  shape  of  the  curve  at  Pt.  It  is  essential  to  choose  a property  of  the  curve  which  can 
be  measured  accurately  as  a basis  of  a reproducible  description.  The  length  rt  satisfies  this  condition. 

With  the  advent  of  more  sophisticated  devices  such  as  a digitizing  table,  the  (x,y)  coordinates 
(text-fig.  6)  of  many  points  on  an  outline  can  be  recorded  with  ease.  A large  number  of  unequally 
spaced  data  points  will  yield  an  accurate  reproduction.  If  the  recorded  points  are  used  to  redraw 
the  outline  (text-fig.  3),  a few  equally  spaced  data  points  may  not  reproduce  the  outline  accurately. 
Although  a large  number  of  data  points  produce  an  accurate  reproduction,  for  the  purposes  of 
comparison  it  is  necessary  to  look  for  a more  economical  way  in  which  to  represent  the  outline. 
A variety  of  curve-fitting  techniques  exist  which  allow  a numerical  approximation  to  the  fossil 
outline  to  be  derived.  A set  of  numbers  which  represent  the  coefficients  in  the  particular 
approximation  employed  can  then  define  the  outline.  For  reasons  which  are  discussed  below,  the 
curve-fitting  method  chosen  here  is  the  Tchebychev  polynomial.  The  purpose  of  this  article  is  to 
determine  the  reliability  of  the  Tchebychev  coefficients  as  shape  descriptors  and  discriminators  in 
the  practical  context  of  bivalves  with  a straight  hinge. 


text-fig.  3.  Reproduction  of  shell  (a)  from  data  points  taken  at  20°  intervals  and  joined  by  ( b ) curved  and  (c) 
straight  lines.  ( d ) is  reproduced  from  many  unequally  spaced  data  points. 


CURVE  FITTING  TECHNIQUES 
All  descriptive,  curve  fitting  techniques  seek  a solution  to  the  equation 
y —f(x)  over  a specified  range  of  x 

where  x is  the  independent  variable  and  y the  dependent  variable.  A length,  rt  (text-figs.  2 and  6), 
called  the  radial  length  below,  can  be  measured  with  reasonable  accuracy  and  is  a suitable  dependent 
variable.  The  angle  (text-fig.  6)  at  the  origin  between  the  radial  length  and  the  reference  line 
A OB  is  an  appropriate  independent  variable  since,  once  this  angle  is  known,  the  value  of  the 
dependent  variable  is  precisely  defined. 

With  these  coordinates,  the  equation  representing  a circle  is  r = a cos  9,  but  the  outline  of  a 
bivalve  shell  is,  in  general,  too  complex  to  be  represented  by  such  a simple  equation.  Accordingly, 
techniques  such  as  splines  and  Fourier  series  have  been  used.  A spline  curve  may  be  fitted  to  a 
series  of  data  points  provided  there  are  no  marked  inflections.  If  such  inflections  occur,  the  data 
points  are  chopped  up  into  segments  at  the  turning-points.  These  turning-points  are  called  knots. 
Suitable  positions  for  knots  (x)  are  shown  on  two  shell  outlines  in  text-fig.  4.  Each  part  of  the  shells 
would  be  accurately  described  by  the  coefficients  of  the  appropriate  spline  function,  but  these 
coefficients  cannot  be  used  to  compare  the  shapes  of  the  shells  (which  may  belong  to  the  same 
species)  because  each  function  relates  to  a different  range  of  the  independent  variable. 


ROGERS:  BIVALVE  SHELL  SHAPE 


text-fig.  4.  Fitting  spline  curves  to  a shell 
outline,  x mark  the  position  of  knots. 


A 


o 


A 


o 


Fourier  series  coefficients  have  been  used  to  define  the  outlines  of  ostracodes  (Kaesler  and  Waters 
1972),  Bryozoa  (Anstey  and  Delmet  1973),  blastoids  (Waters  1977),  and  sand  grains  (Ehrlich  and 
Weinberg  1970).  As  yet  palaeontologists  cannot  readily  understand  what  outline  is  being  described 
by  a particular  set  of  coefficients.  This  led  Scott  (1980),  in  his  study  of  Foraminifera,  to  calculate 
radii  from  the  origin  to  the  outline  at  10°  intervals  from  the  Fourier  series  representation  of  an 
outline.  The  thirty-six  measurements  allow  the  outline  to  be  visualized,  but  the  description  is  not 
economical  and  the  outline  is  undefined  between  the  radii.  A Fourier  series  in  which  both  sine  and 
cosine  terms  are  used  is  periodic,  the  value  of  the  independent  variable  ranging  from  0 to  360°, 
and  is  well  suited  for  describing  the  closed  curves  found  in  the  above  examples,  provided  the  origin 
and  orientation  are  defined.  In  the  case  of  a sand  grain  (text-fig.  5a)  a choice  of  origin  and  orientation 
can  be  made  on  the  basis  of  its  geometry.  The  line  of  orientation  could  be  its  longest  axis  ( AOB ) 
and  the  origin  its  mid-point.  In  radially  symmetric  organisms,  the  centre  is  a point  of  morphologic 
significance  and  the  orientation  can  be  defined  in  morphologic  terms.  As  in  the  case  of  the  sand 
grain,  the  origin  and  orientation  of  a bivalve  can  be  defined  geometrically  (text-fig.  5b),  but  it  is 
not  clear  that  such  definitions  would  be  homologous,  or  have  any  morphologic  relevance.  Biological 
reference  points  within  the  curve,  such  as  muscle  insertions,  are  rarely  preserved  in  fossil  material 
and  are  thus  unsuitable.  In  those  bivalves  having  a straight  hinge,  an  origin  and  orientation  can 
be  chosen  which  are  morphologically  homologous  and  are  usually  visible  in  fossil  material.  Since 
they  lie  on,  rather  than  within,  the  outline,  a periodic  function  may  not  fit  the  outline  most 
economically. 


The  Tchebychev  series  provides  a function  in  which  the  independent  variable  lies  in  the  range 
— 1 to  +1  and  the  NAG  Library  routine  E02ADF  enables  this  function  to  be  fitted  to  an  arbitrary 
set  of  data  points.  For  these  reasons  it  was  decided  to  use  the  coefficients  of  the  polynomials  of  a 
Tchebychev  series  to  describe  the  generating  curve  of  several  bivalves.  Experiments  were  designed 
to  discover  how  many  data  points  are  needed  to  ensure  that  the  calculated  coefficients  are  the 
same  wherever  the  data  points  are  taken,  and  by  whom.  Further,  how  many  coefficients  are  needed 
to  give  a good  representation,  whether  6 or  a transform  of  9 as  the  independent  variable  gives 
better  results;  how  sensitive  the  coefficients  are  to  a change  of  range  in  the  initial  data  points,  and 
how  effective  they  are  as  shape  discriminators.  The  results  of  these  experiments  are  given  below. 

The  basic  objective  of  this  study  is  to  create  attributes  (shape  descriptors)  which  can  be  used  to 
study  variation  within  populations  and  as  aids  to  classification. 

METHODS  OF  TAKING  DATA  POINTS  AND  CALCULATING  THE 
TCHEBYCHEV  COEFFICIENTS 

A digitizing  table  consists  of  a plane  surface  over  which  a pointer  may  be  moved.  In  the  device  used  for  this 
study,  the  pointer  consists  of  the  intersection  of  a pair  of  cross  wires.  When  a button  is  pressed  the  x,y 
coordinates  of  the  pointer  are  recorded  to  an  accuracy  of  0 025  mm.  The  cursor  must  lie  on  the  table  when 
the  coordinates  are  recorded.  Thus  the  outline  to  be  digitized  must  be  drawn  on  a piece  of  paper,  be 
photographed,  or  be  a projected  image.  Material  embedded  in  matrix  cannot  be  digitized  directly. 


text-fig.  5.  Geometric  method  of  defining  the  origin 
and  orientation  of  an  outline.  AOB  = longest  axis, 
O being  the  centre,  (a)  sand  grain,  ( b ) bivalve. 


112 


PALAEONTOLOGY,  VOLUME  25 


Initially,  outlines  of  shells,  scaled  to  be  about  4 cm  square,  were  drawn  using  a camera  lucida.  The  origin 
of  the  polar  coordinates  system,  0,  was  taken  as  the  projection  of  a perpendicular  from  the  tip  of  the  umbo 
on  to  the  hinge  (text-fig.  6b)  and  this  was  the  first  point  digitized.  When  the  specimen  consisted  of  an  external 
representation  of  the  shell,  this  origin  is  concealed  by  the  umbonal  swelling  and  its  position  must  be  inferred 
from  growth  lines.  The  hinge  AOB  (text-fig.  6a)  is  the  line  of  orientation  and  the  posterior  end  of  the  hinge 
(A)  was  the  second  point  digitized.  Subsequent  points,  i.e.  Pt  (text-fig.  6a)  were  digitized  from  A to  B. 


text-fig.  6.  Diagram  showing  the  orientation  of  a 
shell  for  digitizing.  AOB  = hinge  line,  u = tip  of  umbo, 
0 = origin  of  the  polar  coordinates,  O'  = origin  of  x,  y 
coordinates.  xh  y,  are  the  rectangular  coordinates,  and 
rh9t  the  polar  coordinates  of  the  point  P ■ . 


The  (Xj.y,-)  coordinates  can  be  used  to  calculate  the  polar  coordinates  (r;,0;)  of  the  point  Pt.  The  shape  is 
then  defined  by  the  polar  equation: 

r = r(9) 

where  6 is  the  independent  variable  and  r is  the  single  valued  function  of  9.  An  approximation  to  this  function 
may  be  made  by  a Tchebychev  series  of  the  form 

r = jC0T0(v)+  £ cmTJv) 

where  v,  the  independent  variable,  is  9 or  a transform  of  9,  Tmiy)  is  the  Tchebychev  polynomial  of  degree  m, 
and  the  cm  are  the  coefficients  whose  values  define  the  particular  function  r(9)  in  each  case.  A full  account  of 
the  Tchebychev  polynomials  can  be  found,  for  example,  in  Froberg  (1965).  The  general  polynomial  is  defined 
as  Tm(v ) = cos  (m  arccos  v)  and  the  first  few  polynomials  are 
T0(v)  = 1 
7»  = v 
T2(v)  = 2v2  — 1 
T3(v)  = 4v3  — 3v 

When  v = cos  9,  Tmv  = cos  (m9)  and  the  expansion  is  effectively  a Fourier  cosine  series,  the  form  of  which  is 

r = 2^0  + X COS  (W^) 

where  b„  are  the  coefficients  of  the  Fourier  cosine  series.  In  this  case  it  is  a simple  matter  to  calculate  the  area 
of  a specimen,  provided  9 ranges  from  0 to  n radians  (0-180°). 

The  area  = — I-  Y 

4 V 2 

The  coefficients  were  calculated  using  the  NAG  library  routine  E02ADF  which  computes  the  least-squares 
approximation  to  an  arbitrary  set  of  data  points.  The  introductory  remarks  to  this  routine  states  that  more 
points  should  be  recorded  where  the  outline  changes  markedly,  and  also  at  the  ends  of  the  range.  The 
independent  variable  is  scaled  so  that  its  value  ranges  from  — 1 to  + 1 by  the  linear  transform 
V = (2v  — umax  — umin) / (umax  — umin)  where  V = scaled  independent  variable 
vmax  — maximum  value  of  v 
vmin  = minimum  value  of  v 

Thus,  if  9 ranges  from  0 to  180°  and  v = cos  9 then  V = cos  9. 


ROGERS:  BIVALVE  SHELL  SHAPE 


13 


text-fig.  7.  Root  mean  square  residual  (R.M.S.) 
plotted  against  number  of  terms  included. 


The  root  mean  square  residual  (R.M.S.)  is  a measure  of  the  departure  of  the  fitted  curve  from  the  original 
data  points.  Initially  the  residuals  decrease  rapidly  as  successive  terms  are  used  (text-fig.  7);  thereafter  they 
decrease  more  slowly,  and  indeed  may  increase  slightly  as  unwanted  fluctuations  are  produced.  The  number 
of  terms  in  the  series  required  to  give  adequate  accuracy  is  the  number  after  which  the  residuals  decrease 
only  slowly.  In  the  case  shown  in  text-fig.  7,  six  or  seven  terms  give  an  adequate  fit. 


NUMBER  OF  COEFFICIENTS  REQUIRED  AND  CHOICE  OF  THE 
INDEPENDENT  VARIABLE 

The  polar  coordinates  r and  9 of  a digitized  outline  were  used  to  calculate  the  coefficients  of  the  polynomial 
of  the  Tchebychev  series  from  degree  zero  to  degree  eight.  The  plots  of  the  R.M.S.  (text-fig.  7)  for  the  two 
cases  when  v = 6 and  v = cos  9 are  typical  and  show  that  c2  has  little  effect  on  the  accuracy  of  the  fit  of  the 
polynomial.  Thereafter  the  accuracy  of  the  fit  improves  continuously  as  terms  are  added  when  v = cos  9.  The 
R.M.S.  decreases  irregularly  when  v = 9,  and  it  is  difficult  to  identify  where  significant  flattening  of  the  curve 
occurs.  The  values  of  the  coefficients  when  v = cos  9 are  given  in  Table  1.  It  can  be  seen  that  when  a new 
coefficient  (c,)  is  added,  the  value  of  the  preceding  one,  c,  _ j changes  considerably  but  the  value  of  c,  _2  is  only 
slightly  altered.  The  addition  of  a new  coefficient  can  cause  a marked  change  in  the  value  of  all  the  preceding 
coefficients  when  v = 9. 


table  1.  Value  of  the  coefficients  when  polynomials  of  different  degree  are  evaluated 
v = cos  9 is  the  independent  variable 


Degree,  m of 
polynomial 

Terms  included 

c0 

Cl 

c2 

c3 

C4 

c5 

c6 

c7 

c8 

0 

0-798 

1 

0-605 

-0-251 

2 

0-609 

-0-240 

0-032 

3 

0-599 

-0-239 

0-054 

0-050 

4 

0-606 

-0-235 

0052 

0026 

-0063 

5 

0-606 

-0-237 

0-049 

0-030 

-0-048 

0-032 

6 

0-607 

-0-237 

0-048 

0-031 

-0048 

0-026 

-0-016 

7 

0-607 

-0-237 

0-049 

0-030 

-0-049 

0-026 

-0013 

0-007 

8 

0-605 

-0-237 

0-050 

0-030 

-0-048 

0-028 

-0-013 

0-003 

-0-012 

114 


PALAEONTOLOGY,  VOLUME  25 


text-fig.  8.  Original  outline  of  a shell  overlain  by 
the  outline  reproduced  using  5 terms  (a,  b ) and  8 
terms  (c,  d).  The  independent  variable  is  9 (a,  c ) and 
cos  9 ( b , d). 


The  outline  was  recreated  from  the  coefficients  of  the  polynominal,  the  polar  coordinates  of  the  fitted  points 
being  calculated  using  the  NAG  Library  routine  E02AEF.  When  v = 9 the  outline  is  somewhat  more  circular 
than  the  original  (text-fig.  8a,  c)  whereas  the  ends  of  the  outline  are  somewhat  more  inaccurate  when  v = cos  9 
(text-fig.  8 b,  d).  This  is  an  unavoidable  consequence  of  using  cos  9;  the  calculated  curve  must  meet  the  hinge 
at  an  angle  of  90°  when  9 = 0 or  180°.  Cos  0 is  preferred  as  the  independent  variable  because  an  accurate 
portrayal  of  the  ends  of  the  outline  is  considered  of  less  importance  than  the  general  shape  which  is  recreated 
more  accurately  when  v = cos  9.  For  this  reason  and  because  of  the  behaviour  of  the  R.M.S.  and  the  stability 
properties  of  the  coefficients  noted  above,  it  was  decided  to  adopt  cos  9 as  the  independent  variable,  rather 
than  9. 

Inspection  of  the  residuals  suggests  that  the  five  coefficients,  c0-c4,  are  sufficient  to  give  a good  approximation 
(text-fig.  8a,  b).  However,  it  was  found  that  the  outlines  recreated  using  the  eight  coefficients,  c0-c7,  look  much 
more  like  Naiadites  (text-fig.  8c,  d).  The  three  coefficients,  c5-c7,  although  small  in  value  and  accounting  for 
little  of  the  residual  error,  improve  the  detailed  representation  of  the  outline. 


NUMBER  OF  DATA  POINTS 

About  100  data  points  were  recorded  on  the  generating  curve  of  one  shell.  A random  set  of  thirty  of  these 
100  points  were  used  to  calculate  the  coefficients  of  the  Tchebychev  polynomial  to  degree  8.  This  was  done 
ten  times  for  each  of  30,  40,  50,  60,  and  80  data  points.  The  variability  of  c0  is  shown  in  text-fig.  9,  the  value 
of  the  coefficient  being  plotted  against  the  number  of  data  points  used.  It  is  seen  that  when  thirty  data  points 
are  used  the  value  of  the  coefficient  varies  from  0-5  to  1-34  and  is  0-94  when  the  full  data  set  is  used.  About 
100  data  points  yield  a stable  value  of  0-94  for  c0. 

The  same  outline  drawn  and  digitized  by  two  further  people  gave  values  for  c0  ranging  from  0-89  to  0-94 
when  about  100  data  points  were  recorded.  An  error  of  about  5 % was  considered  satisfactory,  particularly 
considering  that  either  the  material  or  the  technique  was  unfamiliar  to  the  participants. 


1-5 


10- 

Co 


0-5- 


0 “I 1 1 1 1 1 1— 

30  A0  50  60  80  126 

No.  of  random  data  points 


text-fig.  9.  Value  of  c0  plotted  against 
number  of  random  data  points  selected 
from  a set  of  about  100  digitized  on  an 
outline.  In  each  case  coefficients  were 
calculated  to  degree  8.  When  forty  or 
more  data  points  are  used,  the  ten  values 
of  c0  cannot  be  recorded  separately  at 
this  scale. 


ROGERS.  BIVALVE  SHELL  SHAPE 


5 


THE  EFFICIENCY  OF  THE  COEFFICIENTS  AS  DISCRIMINATORS  OF  SHAPE 

Two  rather  different  shells  (text-fig.  10)  were  selected,  and  drawn  and  digitized  by  different  operators. 
The  coefficients  were  calculated  to  degree  eight  for  each  digitized  outline.  When  the  value  of  the 
coefficients  is  plotted  (text-fig.  10c),  it  is  seen  that  certain  coefficients  (c0  and  c3)  have  distinctly 
different  values  for  the  two  shells,  whereas  other  coefficients  (cj  are  approximately  the  same  for 
both  shells.  These  two  shells  are  of  similar  size,  and  the  zeroth  coefficient  is  a measure  of  size;  it 
is  the  radius  of  the  semicircle  which  fits  the  outline  giving  the  smallest  root  mean  square  residual. 
Using  this  criterion  as  a measure  of  size,  c0  can  be  used  to  standardize  the  remaining  coefficients. 
Plots  of  the  standardized  coefficients  ( cjc0 , c3/c0)  (text-fig.  10c)  show  that  the  first  coefficient  cannot 
be  used  to  distinguish  the  two  shells,  but  c3  becomes  more  effective  as  a discriminator.  This  result 
is  expected,  because  cy  is  a crude  measure  of  asymmetry  which  is  refined  by  successive  odd-numbered 
coefficients.  Negative  values  of  cy  indicate  that  the  shells  are  asymmetric;  that  is,  the  posterior  lobe 
is  larger  than  the  anterior  lobe.  These  results  indicate  that  shapes  of  Naiadites  shells  of  any  size 
may  be  discriminated  using  at  most  seven  coefficients. 


text-fig.  10.  (a)  and  (b)  two  shells  digitized  and  drawn  by  three  different  people,  (c)  value  of  the 
coefficients  c0-c3  and  the  ratios  cylc0  and  c3/c0  • = shell  a.  + = shell  b. 


EFFECTS  OF  POOR  PRESERVATION 

As  noted  earlier,  the  origin  of  growth  is  often  obscured  in  fossil  material.  However,  provided  the 
umbo  is  not  very  large,  the  tests  made  above  suggest  that  different  operators  will  choose 
approximately  the  same  origin.  A more  serious  problem  is  the  fact  that  the  anterior  lobe  is  often 
distorted,  and  thus  its  outline  cannot  be  digitized  with  confidence.  Although  in  Naiadites  the  anterior 
lobe  is  small,  it  was  found  that  very  different  interpretations  of  the  anterior  outline  gave  rise  to 
markedly  different  coefficients.  Sometimes  the  distortion  brings  the  anterior  lobe  below  the  level 
of  the  hinge,  thus  it  is  not  possible  to  measure  the  radial  length  when  6 = 180°.  It  was  found  that 
provided  v = cos  9 and  9 ranged  from  0 to  about  170°  the  coefficients  obtained  were  similar  to 


16 


PALAEONTOLOGY,  VOLUME  25 


those  obtained  when  6 ranged  from  0 to  180°.  In  reasonably  well-preserved  specimens  of  Naiadites 
it  is  usually  possible  to  use  a range  of  6 of  0-170°. 

In  Curvirimula  it  was  found  that  the  hinge  rarely  extended  beyond  the  umbo,  and  often  the 
maximum  value  of  6 was  120°.  As  noted  earlier,  it  is  impossible  to  compare  the  coefficients  of  two 
curves  if  the  ranges  to  which  they  fit  are  very  different. 

Further,  not  only  should  the  range  of  6 be  the  same,  but  the  morphologic  structures  described 
should  be  the  same  if  the  coefficients  are  to  be  used  as  discriminators.  Thus,  if  the  range  of  6 chosen 
is  0 to  120°,  the  coefficients  calculated  for  one  shell  (text-fig.  11a)  describe  the  whole  of  the  generating 
curve,  whereas  those  for  a second  shell  (text-fig.  lib)  fail  to  describe  the  anterior  lobe,  and  the  two 
sets  of  coefficients  cannot  be  used  to  compare  the  generating  curve  of  the  two  shells.  In  the  genera 
studied  the  umbo  and  the  hinge  are  the  only  structures  on  the  outline  which  are  homologous. 
However,  the  anterior  lobe  contains  the  anterior  adductor  muscles,  part  of  the  foot  and  its  associated 
musculature,  and  the  size  of  the  anterior  lobe  is  a good  indicator  of  the  mode  of  life  of  the  organism. 
A suitable  way  of  comparing  these  shells  may  be  to  describe  the  posterior  lobe  in  terms  of  the 
coefficients  which  define  its  outline,  and  to  describe  the  anterior  lobe  in  terms  of  area,  a parameter 
which  can  easily  be  calculated  from  the  digitized  points.  Such  a description  would  still  be  economical, 
and  a comparison  of  shells  described  in  this  way  would  be  justified  on  theoretical  grounds.  This 
method  has  not  yet  been  tested. 


text-fig.  1 1.  Two  specimens  of  Curvirimula  having  very 
different  maximum  value  of  6. 


SUMMARY  AND  CONCLUSION 

The  shape  of  the  generating  curve  of  those  shells  (including  Naiadites)  which  have  a straight  hinge 
extending  beyond  the  umbo  may  be  described  economically  using  the  coefficients  of  a polynomial 
of  the  Tchebychev  series  of  the  eighth  degree.  Such  descriptors  are  very  stable  provided  that  about 
100  data  points  are  used  and  that  cos0  is  used  as  the  independent  variable.  C0  is  a measure  of 
size,  and  tests  indicate  that  if  it  is  used  to  standardize  the  coefficients  of  higher  degree,  the  first  five 
scaled  coefficients  can  be  used  as  shape  discriminators. 

Describing  shells  in  this  way  has  the  further  advantage  that,  because  the  whole  of  the  generating 
curve  can  be  reproduced,  any  other  features  of  the  shell  such  as  the  longest  radial  length  can  be 
calculated  from  these  nine  coefficients.  If  6 ranges  from  0 to  180°,  other  features  such  as  the  area  can 
be  calculated  directly  from  the  coefficients. 

When  the  hinge  is  not  straight,  or  does  not  extend  beyond  the  umbo,  the  shape  can  still  be 
described,  but  it  is  difficult  to  specify  the  range  of  the  independent  variable  in  such  a way  that 
different  species  can  be  compared.  In  the  case  of  Curvirimula  the  hinge  rarely  extends  beyond  the 
umbo.  For  this  genus  it  is  suggested  that  the  coefficients  of  a very  restricted  portion  of  the  generating 
curve,  together  with  a further  parameter,  the  area  of  the  anterior  lobe,  which  has  functional 
significance,  may  be  used  in  order  to  compare  shells.  For  shells  which  lack  a straight  hinge  it  is 
possible  that  a periodic  function  may  be  more  appropriate  as  a shape  discriminator.  The  umbo 
could  be  retained  for  the  purpose  of  orientation,  but  it  may  be  necessary  to  define  a geometric  origin. 

Acknowledgements.  The  author  would  like  to  thank  Professor  M.  H.  Rogers  for  much  assistance  with  the 
mathematical  techniques  employed,  Dr.  G.  N.  Lance  for  critical  reading  of  the  manuscript,  and  Mrs.  J.  Bees 
for  drafting  the  illustrations.  The  ICL  4-75  computer  at  Bristol  University  was  used  for  all  the  calculations, 
and  the  assistance  of  the  computer-centre  staff  is  gratefully  acknowledged. 


ROGERS:  BIVALVE  SHELL  SHAPE 


117 


REFERENCES 


anstey,  r.  l.  and  delmet,  D.  A.  1973.  Fourier  analysis  of  zooecial  shapes  in  fossil  tubular  bryozoans.  Bull.  geol. 
Soc.  Amer.  84,  1753-1764. 

bookstein,  F.  L.  1978.  The  measurement  of  biological  shapes  and  shape  change.  Lecture  notes  in  biomathematics, 
24,  1-191. 

davies,  J.  h.  and  trueman,  a.  e.  1927.  A revision  of  the  non-marine  lamellibranchs  of  the  Coal  Measures. 
Quart.  Jl  geol.  Soc.  83,  210-259. 

deleers,  c.  and  pastiels,  A.  1947.  Etude  biometrique  des  Anthraconauta  du  Houiller  de  la  Belgique.  Publ. 
Assoc,  strat.  Houilleres,  2,  1-99. 

eagar,  r.  m.  c.,  1973.  Variation  in  shape  of  shell  in  relation  to  palaeoecological  station  in  some  non-marine 
Bivalvia  of  the  Coal  Measures  of  south  east  Kentucky  and  of  Britain.  C.R.  7th  Cong.  Int.  Strat.  Geol.  Carb., 
Krefeld,  1971,  2,  387-416. 

erlich,  r.  and  weinberg,  b.  1970.  An  exact  method  for  characterization  of  grain  shape.  J.  sed.  Pet.  40, 
205-212. 

froberg,  c.  e.  1965.  Introduction  to  numerical  analysis.  Reading,  Mass.,  Addison-Wesley  publ.  Co.,  x + 340  pp. 

HAJK.R,  o.,  lukasova.  A.,  ruzicka,  b.,  and  rehor,  F.  1974.  Die  gattung  Citothyris  (Bivalvia)  aus  dem  Karbon 
und  ihr  statistich-geometrisches  Modell.  Freiberg,  Fortschungsh.  C306,  1-159. 

kaesler,  r.  l.  and  waters,  J.  a.  1972.  Fourier  analysis  of  the  ostracode  margin.  Bull.  geol.  Soc.  Amer.  83, 
1169-1178. 

papin,  yu.  s.  and  khoroshev,  n.  g.  1974.  Numerical  expression  of  the  shape  of  shells  outlined  by  a convex  or 
convexo-concave  smooth  line.  Zh  Paleont.  8,  121-124.  [In  Russian.] 

scott,  G.  h.  1980.  The  value  of  outline  processing  in  the  biometry  and  systematics  of  fossils.  Palaeontology, 
23,  757-768. 

raup,  d.  m.,  1966.  Geometric  analysis  of  shell  coiling:  general  problems.  J.  Paleont.  40,  1178-1190. 

trueman,  a.  e.  and  weir,  j.  1955.  The  Carboniferous  non-marine  Lamellibranchia.  Palaeontogr.  Soc.  Mono- 
graph, 8,  207-242. 

waters,  J.  a.,  1977.  Quantification  of  shape  by  use  of  Fourier  analyses:  the  Mississippian  Blastoid  genus 
Pentremites.  Paleobiology,  3,  288-299. 

nag  library  1978.  Mark  6.  Numerical  Algorithms  Group  Ltd.,  Oxford,  U.K. 


M.  J.  ROGERS 


Typescript  received  7 July  1980 

Revised  typescript  received  30  November  1980 


Department  of  Geology 
Queen’s  Building 
University  Walk 
Bristol  BS8  1TR 


REWORKED  ACRITARCHS  FROM  THE  TYPE 
SECTION  OF  THE  ORDOVICIAN  CARADOC 
SERIES,  SHROPSHIRE 


by  ROBERT  E.  TURNER 


Abstract.  Thirty-seven  acritarch  species  in  the  type  Caradoc  rocks  of  south  Shropshire  are  recognized  as 
reworked  from  strata  of  Tremadoc  and  Arenig/Llanvirn  age,  their  distribution  reflecting  an  inverted 
stratigraphy.  These  microfossils  yield  valuable  palaeoenvironmental  data,  their  excellent  preservation  indicating 
a source  on,  or  adjacent  to,  the  Midland  Platform;  erosion  of  the  relatively  unconsolidated  parent  sediments 
occurred  in  a marine  environment.  Acritarchs  were  eroded  and  redeposited  as  discrete  particles  and  wave  and 
current  action  are  considered  the  most  likely  erosive  agents.  The  rate  of  release  of  microfossils  is  linked  to  a 
shallowing  of  the  water-body;  this  resulted  in  a shift  to  a high  energy  environment  with  storm  surges  influencing 
the  erosion  and  redeposition  of  sedimentary  particles. 

Reworking  of  palynomorphs,  the  erosion  and  redeposition  of  the  microfossils  within  younger 
sediments,  is  a problem  familiar  to  most  palynologists  (Funkhauser  1969;  Richardson  and  Rasul 
1978, 1979)  and  has  been  discussed  by  Wilson  (1964).  Palynomorphs  are  prone  to  reworking  because 
of  their  small  size,  abundance,  and  the  durable  nature  of  their  complex  wall.  This  paper  attempts  to 
show  that  although  reworking  may  present  a problem  in  biostratigraphical  studies,  it  can  provide  a 
valuable  insight  into  the  depositional  environment  of  the  enclosing  sediments.  Murchison  (1839) 
introduced  the  term  Caradoc  to  geology  when  he  applied  the  name  ‘Caradoc  Sandstone’  to  a group 
of  strata  cropping  out  in  the  county  of  Shropshire.  The  clearest  section  was  said  by  Murchison  (p. 
216)  to  occur  in  the  valley  of  the  River  Onny  near  Horderley.  Usage  of  the  term  Caradoc  has  been 
much  restricted  by  subsequent  workers  and  was  given  status  as  a Series  by  Lapworth  (1916)  who 
subdivided  it  into  Groups  based  on  lithology  and  faunal  content.  The  exposures  along  the  valley  of 
the  River  Onny  are  accepted  as  the  type  section  on  historical  grounds  and  these  still  comprise  the  best 
exposed  and  most  complete  succession  of  Caradoc  rocks  in  the  district.  Since  the  recent  works  on  the 
A489  road  adjacent  to  the  river,  exposure  is  continuous  through  much  of  the  sequence  and  access  is 
better  than  it  has  been  for  many  years.  Extensive  researches  have  been  carried  out  on  the  macro- 
faunas and  biostratigraphy  of  these  sediments,  a comprehensive  review  of  these  works  being  provided 
by  Hurst  (1979c).  Palaeontological  studies  have  concentrated  on  the  rich  shelly  faunas,  but  Jenkins 
( 1 967)  published  some  of  the  earliest  microfossil  data  with  his  detailed  examination  of  the  chitinozoa. 
Twenty  samples  were  collected  from  the  Onny  Valley  as  part  of  a larger  study  of  Llandeilo  and 
Caradoc  acritarchs  from  the  British  Isles.  A further  thirteen  samples  were  obtained  for  comparative 
purposes  from  north  Shropshire  in  the  vicinity  of  Chatwall  Farm  (SO  5137  9745,  text-fig.  2).  All 
samples  were  treated  with  standard  laboratory  techniques  for  the  recovery  of  palynomorphs  and 
most  yielded  abundant,  diverse,  and  well-preserved  acritarchs.  The  majority  of  samples  examined 
contain  numerous  acritarchs  of  undoubted  Caradoc  age  (text-fig.  5);  these  form  the  bulk  of  the 
acritarch  assemblages  recovered  and  will  be  described  elsewhere. 

All  figured  specimens  are  registered  and  held  in  the  MPK  series  of  the  palynological  collections  at  the  Institute  of 
Geological  Sciences,  Leeds. 


IPalaeontology,  Vol.  25,  Part  1,  1982,  pp.  119-143,  pis.  15-17.1 


120 


PALAEONTOLOGY,  VOLUME  25 


AGE  OF  THE  REWORKED  MATERIAL 

Samples  from  the  Onny  Valley  contain  reworked  acritarchs  that  evidently  originated  from  different  sources 
since,  on  an  age-basis,  they  fall  into  three  broad  categories. 

Category  1.  Acritarchs  of  Tremadoc  age 

Some  species  recovered  can  be  identified  as  Tremadoc  forms  because  their  known  stratigraphical  ranges  are 
restricted  to  strata  of  this  age.  Their  occurrences  here  cannot  be  interpreted  as  extensions  of  their  ranges  because 
the  species  concerned  remain  unknown  in  strata  of  Arenig  to  Llandeilo  age  (see  text-figs.  3,  4). 

Category  2.  Acritarchs  of  probable  Tremadoc  age 

This  category  consists  of  taxa  that  are  known  both  from  the  Tremadoc  and  the  Arenig/Llanvirn  but  which, 
other  than  this  occurrence,  have  not  been  recorded  from  younger  strata  in  the  British  Isles,  or  elsewhere  (text- 
figs.  3,  4).  That  these  forms  are  reworked  seems  certain,  but  it  is  not  possible  to  assign  to  them  a precise 
stratigraphical  age.  Despite  this,  although  an  Arenig/Llanvirn  age  is  accepted  as  possible  for  some  of  these 
individuals,  it  is  probable  that  the  majority  were  derived  from  rocks  of  Tremadoc  age.  This  is  indicated  by  the 
large  numbers  involved  (see  text-fig.  3)  which  suggest  derivation  from  particularly  rich  pre-existing  acritarch 
assemblages.  It  is  known  that  parts  of  the  Tremadoc  sequence  in  Great  Britain  yield  abnormally  abundant 
acritarch  populations;  a rough  calculation  suggested  a figure  of  100,000  individuals  per  gram  of  rock,  for  a 
sample  from  the  Shumardia  pusilla  Zone  of  the  Shineton  Shales  of  Shropshire  (Downie  1958,  p.  332).  These 
profuse  numbers  are  reflected  in  the  periodic  reworking  of  Tremadoc  acritarchs  into  other  parts  of  the  geological 
column;  for  example,  reworked  Tremadoc  assemblages  have  been  identified  in  Devonian  rocks  in  Oxfordshire 
(Richardson  and  Rasul  1979)  and  are  known  from  Llanvirn  sediments  in  north-west  England  (author’s 
unpublished  data).  Strata  of  Tremadoc  age  are  unquestionably  the  commonest  recognized  source  of  reworked 
acritarchs  in  the  British  Isles.  In  contrast,  the  work  of  Booth  1979  (unpublished  Ph.D.  thesis)  and  the  present 
author’s  own  unpublished  data  suggest  that  in  Britain  the  Arenig/Llanvirn  was  a period  of  substantially  lower 
phytoplankton  productivity.  Thus  most  of  the  acritarchs  in  this  group  were  probably  redeposited  from 
sediments  of  Tremadoc  age. 


text-fig.  1.  Location  and  geological  setting  of  the  Caradoc  type  section. 


TURNER:  REWORKED  ACRITARCHS 


121 


Category  3.  Acritarchs  of  Arenig/Llanvirn  age 

The  species  placed  in  this  category  have  stratigraphical  ranges  restricted  to  strata  of  this  age.  They  are  not  present 
in  the  type  Llandeilo  of  South  Wales  nor  have  they  been  recorded  from  Llandeilo  or  younger  rocks  elsewhere  in 
the  world. 


text-fig.  2.  Sketch-map  of  the  outcrop  of  Ordovician  rocks  around  Church 
Stretton,  Shropshire,  and  the  relative  positions  of  exposures  at  Chatwall  and  the 
River  Onny. 


122 


PALAEONTOLOGY,  VOLUME  25 


text-fig.  3.  The  distribution  and  numerical  abundance  of  reworked  acritarchs  in  samples  from  the  Caradoc  type 
section.  The  figure  for  each  category  of  reworked  acritarchs  identified  in  every  sample  is  given  as  a percentage 
based  on  a count  of  200  specimens,  or  1 00  specimens  where  acritarchs  are  sparse. 


TURNER:  REWORKED  ACRITARCHS 


123 


text-fig.  4.  Previous  records  of  the  reworked  taxa  identified  in  the  type  Caradoc.  Locality  numbers  indicate  the 
following  references.  1.  Vanguestaine  1978:  2.  Timofeev  1959:  3.  Rasul  and  Downie  1974:  4.  Downie  1958:  5. 
Rasul  1974:  6.  Martin  1977:  7.  Rasul  1979:  8.  Combaz  1967: 9.  Martin  1973:  10.  Deunff  1961:  1 1.  Gorka  1967:  12. 
Timofeev  1966:  13.  Martin  1969:  14.  Rauscher  1974:  15.  Rasul  1976:  16.  Vavrdova  1972:  17.  Vavrdova  1973:  18. 
Booth  1979:  19.  Cramer  and  Diez  1977:  20.  Dean  and  Martin  1978:  21.  Vavrdova  1966:  22.  Vavrdova  1977:  23. 
Vavrdova  1976:  24.  Turner  and  Wadge  1979:  25.  Burmann  1970:  26.  unpublished  data:  27.  Cramer  and  Diez 
1976:  28.  Cramer,  Allam,  Kanes,  and  Diez  1974:  29.  Cramer,  Kanes  etal.  1974:  30.  Downie  and  Soper  1972:  31. 
Burmann  1968:  32.  Paris  and  Deunff  1970:  33.  Loeblich  and  Tappan  1978. 


124 


PALAEONTOLOGY,  VOLUME  25 


DESCRIPTIVE  PALAEONTOLOGY 
Group  acritarcha  Evitt  1963 

The  system  of  informal  ‘subgroups’  proposed  by  Downie,  Evitt,  and  Sarjeant  (1963)  has  no  status  under  the 
provisions  of  the  International  Code  of  Botanical  Nomenclature  (I.C.B.N.).  As  indicated  by  Wicander  1974 
(p.  1 1),  the  introduction  of  new  subgroups  which  reflect  generic  names  (Staplin,  Jansonius,  and  Pocock,  1965) 
poses  substantial  nomenclatural  problems.  For  these  reasons  it  is  preferred  here  not  to  organize  the  acritarcha 
into  a suprageneric  classification  but  to  simply  list  them  in  alphabetical  order. 

Genus  acanthodiacrodium  (Timofeev  1958)  Deflandre  and  Deflandre-Rigaud  1961 

The  status  in  the  literature  of  the  ‘diacrodians’  is  confused,  a situation  created  when  many  of  the  early 
species  described  were  assigned  to  invalid  genera  by  Timofeev  (1959).  The  resultant  taxonomic 
confusion  was  exacerbated  by  Deflandre  and  Deflandre-Rigaud  (1961)  who  revised  and  restricted 
most  of  Timofeev’s  original  genera;  unfortunately  many  of  their  emendations  are  either  invalid  or 
illegitimate  under  various  provisions  of  the  I.C.B.N.  and  must  be  rejected.  Leoblich  and  Tappan 
1978,  discussed  this  problem  (pp.  1236-1238)  and  created  a new  genus,  Actinotodissus.  This  appears 
to  be  differentiated  from  Acanthodiacrodium  on  minor  variations  in  morphology  and  has  yet  to  be 
widely  accepted.  Considerable  numbers  of  individuals  attributable  to  Acanthodiacrodium  or  possibly 
to  Actinotodissus  were  recorded  here  but  no  attempt  was  made  to  speciate  them.  Although  the 
stratigraphical  distribution  of  the  ‘diacrodians’  is  not  wholly  understood,  it  is  clear  that  in  the 
Tremadoc  rocks  of  Great  Britain  such  forms  occur  in  abundance,  while  in  Arenig  and  younger  rocks 
they  are  relatively  rare. 

A can  thodiacrodium / Ac tino  todissus  spp . 

Plate  16,  fig.  3 

Description.  Central  vesicle  varying  in  outline  from  ovate  with  rounded  poles  to  elongate-subovate.  Opposite 
poles  bear  similar  processes  which  may  be  hollow  or  solid;  the  central  portion  of  the  vesicle  is  always  without 
processes.  Both  vesicle  and  process  wall  are  usually  smooth,  rarely  granulate;  the  equatorial  zone  may  bear 
longitudinal  striae.  No  excystment  structure  recorded. 

Genus  archaeohystrichosphaeridium  Timofeev  1959  ex  Loeblich  and  Tappan  1976 

The  genus  Archaeohystrichosphaeridium  is  technically  invalid  (see  Loeblich  and  Tappan  1976,  p.  303) 
but  many  forms  originally  described  under  this  name  have  not  yet  been  transferred  to  other  genera. 
The  name  is  retained  here  pending  transfer  of  included  species  to  other  genera. 

Archaeohystrichosphaeridium  zalesskyi  Timofeev  1959 

Description.  Central  vesicle  spherical,  smooth,  bearing  a moderate  number  (15-25)  of  smooth,  simple,  hollow, 
homomorphic  processes  which  have  wide  bases  tapering  to  an  acuminate  distal  termination.  The  process  interior 
communicates  freely  with  the  vesicle  cavity.  No  excystment  structure  recorded.  Vesicle  diameter  24-33  /un; 
process  length  5-10  ^m.  Four  specimens  measured. 

Remarks.  The  spherical  forms  described  here  probably  belong  to  the  genus  Solisphaeridium  Staplin, 
Jansonius  and  Pocock  (1965),  but  transfer  should  await  the  examination  of  in  situ  material. 

Genus  arkonia  Burmann  1970 
Arkonia  tenuata  Burmann  1970 

Description.  Central  vesicle  hollow,  triangular  in  outline,  compressed  with  each  angle  bearing  a long,  hollow, 
smooth,  simple  process  tapering  gradually  to  an  acuminate  distal  termination.  Processes  communicate  freely 
with  the  vesicle  cavity  and  all  lie  in  the  same  plane  as  the  compression  of  the  central  body.  The  vesicle  wall  is 


TURNER:  REWORKED  ACRITARCHS 


25 


ornamented  with  fine,  closely  spaced  striae  which  are  approximately  parallel  to  the  vesicle  sides;  these  striae  do 
not  extend  on  to  the  process  wall  which  is  smooth.  No  excystment  structure  recorded.  Vesicle  height  27-32  /xm; 
process  length  33-36  /u.m.  Three  specimens  measured. 

Remarks.  A.  tenuata  differs  from  A.  virgata  Burmann  1970  by  having  more  numerous,  more  closely 
spaced,  and  finer  striae  ornamenting  the  vesicle  wall. 

Arkonia  virgata  Burmann  1970 

Description.  Similar  to  A.  tenuata  in  over-all  morphology  but  the  vesicle  wall  is  ornamented  with  fewer,  coarse, 
widely  spaced  striae.  No  excystment  structure  recorded.  Vesicle  height  28-3 1 /urn;  process  length  34-39  /xin.  Four 
specimens  measured. 


Genus  coryphidium  Vavrdova  1972 
Coryphidium  australe  Cramer  and  Diez  1976 

Plate  17,  fig.  4 

Description.  Central  vesicle  hollow,  quadrate  in  outline  with  rounded  corners,  strongly  compressed.  The  vesicle 
wall  bears  numerous  (more  than  fifty)  short,  relatively  thick  processes  but  is  otherwise  smooth.  Processes  tend  to 
be  more  closely  spaced  towards  the  corners  of  the  vesicle  and  sparse  on  the  central  portions.  Processes  are 
hollow,  smooth,  and  communicate  freely  with  the  vesicle  cavity;  distal  terminations  may  be  capitate,  bifurcate, 
or  irregularly  bulbous.  No  excystment  structure  recorded.  Vesicle  width  39-42  ^.m;  process  length  6-8  /im.  Eight 
specimens  measured. 


Coryphidium  bohemicum  Vavrdova  1972 

Description.  Central  vesicle  hollow,  quadrate  in  outline  with  rounded  corners,  strongly  compressed,  the  sides  of 
the  central  body  may  be  almost  straight  or  concave.  The  vesicle  wall  bears  numerous  (30-60),  short,  relatively 
thick  processes  which  are  concentrated  towards  the  corners,  with  few  or  none  on  the  central  portions  of  the 
vesicle.  These  processes  are  hollow,  smooth,  and  communicate  freely  with  the  vesicle  cavity;  distal  terminations 
may  be  flat-topped,  bifurcate,  multifurcate,  capitate,  or  irregularly  bulbous.  The  vesicle  wall  also  bears  well- 
developed  striae  which  are  mostly  restricted  to  those  central  areas  having  few  processes;  these  striae  are 
approximately  parallel  to  the  vesicle  sides  around  the  margins  but  towards  the  centre  may  become  strongly 
concave.  No  excystment  structure  recorded.  Vesicle  width  22-27  /xm;  process  length  3-6  fim.  Five  specimens 
measured. 

Remarks.  Probable  occurrences  of  this  species,  recorded  as  ‘Indetermine  forme  A’  in  the  Caradoc  of 
Ombret,  Belgium  (Martin,  Michot,  and  Vanguestaine,  1970)  are  here  interpreted  as  reworked  (see 
also  Martin  1977,  fig.  14). 

Coryphidium  elegans  Cramer,  Allam  et  al.  1974 

Description.  Central  vesicle  hollow,  quadrate  in  outline  with  rounded  corners,  strongly  compressed.  The  vesicle 
wall  bears  numerous  (30-60),  short,  slender  processes  which  tend  to  be  concentrated  towards  the  corners  with 
few  or  none  on  the  central  portions.  Processes  are  smooth,  apparently  solid,  and  the  distal  terminations  may  be 
rounded  or  capitate.  The  vesicle  wall  also  bears  well-developed  striae  which  are  most  prominent  on  the  central 
portions  but  may  extend  into  the  corners;  these  striae  are  approximately  parallel  to  the  vesicle  sides  but  tend  to 
become  concave  towards  the  centre.  No  excystment  structure  recorded.  Vesicle  width  20-21  ^m;  process  length 
3-5  jam.  Two  specimens  measured. 

Genus  cymatiogalea  (Deunff  1961)  Deunff,  Gorka,  and  Rauscher  1974 
Cymatiogalea  cristata  (Downie  1958)  Deunff,  Gorka,  and  Rauscher  1974 

Plate  15,  fig.  2 

Description.  Central  vesicle  spherical  to  sub-spherical;  wall  granular,  divided  into  polygonal  fields  by  low  sutural 
ridges  which  bear  smooth,  apparently  solid  processes  dividing  distally  into  two  to  four  simple  lateral  branches; 
the  polygonal  areas  between  sutural  ridges  lack  processes.  Excystment  is  by  the  development  of  a large  round  to 


126 


PALAEONTOLOGY,  VOLUME  25 


sub-polygonal  polar  opening,  the  periphery  of  which  also  bears  processes.  The  operculum,  which  is  commonly 
preserved  inside  the  vesicle,  is  devoid  of  processes  but  has  a coarse  granular  ornament.  Vesicle  diameter  25-31 
/xm;  excystment  opening  27-30  /xm;  process  length  15-28  fim.  Eight  specimens  measured. 

Cymatiogalea  velifera  (Downie  1958)  Martin  1969 

Plate  15,  fig.  1 

Description.  Central  vesicle  spherical  to  sub-spherical;  wall  ornamented  with  irregular  grana  and  divided  into 
polygonal  fields  by  low  sutural  ridges  that  bear  processes  supporting  thin  membranes.  The  processes  are  smooth, 
slender,  hollow  with  a solid  proximal  plug  separating  the  process  interior  from  the  vesicle  cavity;  distally  the 
processes  divide  into  two  or  four  simple  lateral  branches.  Membranes  are  delicate  and  transparent  although 
sometimes  faint  striations  or  thickenings  may  be  seen.  Polygonal  areas  between  sutural  ridges  lack  processes. 
Excystment  is  by  the  development  of  a large  sub-polygonal  polar  opening  the  periphery  of  which  also  bears 
processes  supporting  membranes.  The  operculum  which  commonly  is  preserved  inside  the  vesicle  lacks  both 
processes  and  membranes  but  has  a coarse  granular  ornament.  Vesicle  diameter  30-40  /xm;  excystment  opening 
26-35  ftm;  process  length  7-12  jam.  Ten  specimens  measured. 

Genus  dasydiacrodium  (Timofeev  1959)  Deflandre  and  Deflandre-Rigaud  1961 
Dasydiacr odium  palmatilobum  Timofeev  1959 

Plate  15,  fig.  4 

Description.  Central  vesicle  ellipsoidal,  smooth  with  rounded  poles,  one  pole  bears  approximately  fifteen  simple, 
smooth,  hollow,  homomorphic  processes  which  have  wide  bases  and  taper  rapidly  to  an  acuminate  distal 
termination.  The  opposite  pole  bears  a larger  number  (about  twenty-five)  of  much  shorter  but  otherwise  similar 
processes.  The  intervening  equatorial  portion  of  the  vesicle  is  without  ornament.  The  interiors  of  all  processes 
communicate  freely  with  the  vesicle  cavity.  No  excystment  structure  recorded.  Length  of  long  axis  31-33  /xm; 
short  axis  26-28  jam;  long  processes  27-30  /xm;  short  processes  11-14  jam.  Three  specimens  measured. 

Genus  dicrodiacrodium  Burmann  1968 
Dicrodiacrodium  normale  Burmann  1968 

Plate  17,  fig.  5 

Description.  Central  vesicle  is  heteropolar,  approximately  oval  in  outline  with  a broadly  rounded  antapical  pole 
and  a more  sharply  rounded  apical  pole.  The  apical  pole  bears  a single,  smooth,  hollow,  cylindrical  process 
which  has  a solid  proximal  plug  separating  the  process  interior  from  the  vesicle  cavity.  Distally  this  process 
terminates  in  five  to  six  short,  simple,  acuminate  branches  giving  a grapnel-like  appearance.  The  antapical  pole 
bears  a dense  anastomosing  network  of  fine  threadlike  processes.  The  vesicle  wall  is  ornamented  with  widely 
spaced  longitudinal  striae.  No  excystment  structure  recorded.  Vesicle  height  58  /xm;  vesicle  width  33  /xm;  length 
of  apical  process  25  /xm.  One  specimen  measured. 


explanation  of  plate  15 
Selected  acritarchs  of  Tremadoc  age 
All  figures  x 1200 

Fig.  1.  Cymatiogalea  velifera  (Downie)  Martin.  OV/A/2b-5,  MPK  2732,  Onny  Valley,  Alternata  Limestone. 
2.  C.  cristata  (Downie)  Deunff,  Gorka  and  Rauscher.  OV/A/2b-l,  MPK  2733,  Onny  Valley,  Alternata 
Limestone.  3.  Stelliferidium  stelligerum  Deunff,  Gorka  and  Rauscher.  OV/A/la-1,  MPK  2734,  Onny 
Valley,  Alternata  Limestone.  4.  Dasydiacrodium  palmatilobum  Timofeev.  OV/UHS/1-2,  MPK  2735,  Onny 
Valley,  Horderley  Sandstone.  5.  Trichosphaeridium  annolovaense  Timofeev.  OV/A/la-1,  MPK  2736,  Onny 
Valley,  Alternata  Limestone.  6.  Saharidia  fragile  (Downie)  Gombaz.  OV/HS/1-1,  MPK  2737,  Onny  Valley, 
Harnage  Shales,  phase-contrast. 


PLATE  15 


TURNER,  Tremadoc  acritarchs 


128 


PALAEONTOLOGY,  VOLUME  25 


Genus  dictyotidium  (Eisenack  1955)  Staplin  1961 
Dictyotidium?  dentatum  (Vavrdova  1976)  Dean  and  Martin  1978 

Description.  Central  vesicle  hollow,  sub-polygonal  in  outline.  The  vesicle  surface  is  divided  into  a small  number 
of  polygonal  fields  (nine  in  the  one  individual  recorded)  delineated  by  prominent,  smooth,  transparent 
membranes  the  outer  edges  of  which  carry  a row  of  short,  capitate,  often  flat-topped  denticles.  No  excystment 
structure  recorded.  Vesicle  diameter  45  /urn.  One  specimen  measured. 

Genus  frankea  Burmann  1970 
Frankea  breviuscula  Burmann  1970 

Description.  Central  vesicle  hollow,  smooth,  triangular  in  outline,  strongly  compressed.  Each  angle  bears  a single 
slender,  smooth  process  of  moderate  length  (up  to  70%  of  vesicle  height)  that  tapers  gradually  towards  the  distal 
end;  here  it  divides  into  five  or  six  short,  simple,  acuminate  lateral  branches  that  arise  in  a single  plane  normal  to 
the  compression  of  the  vesicle.  The  hollow  processes  all  lie  in  the  plane  of  the  central  body  and  communicate 
freely  with  the  vesicle  cavity.  No  excystment  structure  recorded.  Vesicle  height  28  ^m;  process  length  15 /urn.  One 
specimen  measured. 


Frankea  hamata  Burmann  1970 

Description.  Similar  to  F.  breviuscula  in  over-all  morphology  but  differing  in  having  shorter  processes  that 
always  divide  distally  into  two  long,  smooth,  simple,  strongly  recurved  lateral  branches  with  acuminate 
terminations;  these  bifurcations  occupy  the  same  plane  as  the  compression  of  the  vesicle.  No  excystment 
structure  recorded.  Vesicle  height  24-26  /nm;  process  length  13-15  /u,m;  length  of  lateral  branches  11-14  pm. 
Three  specimens  measured. 


Frankea  hamulata  Burmann  1970 

Description.  Similar  to  F.  breviuscula  in  over-all  morphology  but  differing  in  having  short  slender  processes.  No 
excystment  structure  recorded.  Vesicle  height  29  /urn;  process  length  9 /im;  length  of  lateral  branches  1-2  pm. 
One  specimen  measured. 

Frankea  longiuscula  Burmann  1970 

Description.  Similar  to  F.  breviuscula  in  over-all  morphology  but  with  very  long  (up  to  150%  of  vesicle  height) 
slender  processes.  No  excystment  structure  recorded.  Vesicle  height  41  pm;  process  length  63  /urn;  length  of 
lateral  branches  5 /im.  One  specimen  measured. 

Frankea  sartbernardense  (Martin  1966)  Burmann  1970 

Description.  Similar  to  F.  breviuscula  in  over-all  morphology  but  with  very  short  stout  processes.  No  excystment 
structure  recorded.  Vesicle  height  21-24  pm;  process  length  3-4  /um;  length  of  lateral  branches  2-3  /um.  Three 
specimens  measured. 

Remarks.  Records  of  this  species  in  the  Silurian  of  Belgium  (Martin  1969)  are  interpreted  as  reworked 
by  the  present  author. 

Genus  impluviculus  (Loeblich  and  Tappan  1969)  Martin  1977 
Impluviculus  cf.  lenticularis  Martin  1977 

Plate  16,  fig.  2 

Description.  Central  vesicle  hollow,  compressed,  polygonal  in  outline,  apparently  smooth.  Each  angle  bears  a 
slender,  flagelliform  process  which  tapers  gradually  to  a closed  distal  termination;  these  processes  are  hollow 
proximally  and  communicate  freely  with  the  vesicle  cavity  but  may  become  solid  distally.  Sometimes  two 
processes  may  be  closely  located  forming  a pair.  All  processes  arise  around  the  margins  of  the  central  body  and 
lie  in  the  plane  of  compression.  Vesicle  diameter  9 /um;  process  length  24-28  /u.m.  One  specimen  measured. 

Remarks.  Assignment  to  I.  lenticularis  is  not  certain  since  the  processes  are  much  longer  than  those  in 
the  type  material  from  the  Tremadoc  of  Brabant,  described  by  Martin  (1977). 


TURNER:  REWORKED  ACRITARCHS 


29 


Genus  marrocanium  Cramer,  Kanes  et  al.  1974 
Marrocanium  simplex  Cramer,  Kanes  et  al.  1974 

Plate  17,  fig.  1 

Description.  Central  vesicle  hollow,  smooth,  quadrate  in  outline,  strongly  compressed.  Each  angle  bears  a single, 
smooth,  simple  process  which  tapers  to  a slightly  rounded  distal  termination.  The  hollow  processes  lie  in  the 
same  plane  as  the  central  body  and  communicate  freely  with  the  vesicle  cavity.  Thin,  transparent  membranes  are 
suspended  between  the  processes;  those  in  a lateral  position  are  wide,  stretching  out  to  the  process  tips  while  the 
apical  and  antipical  membranes  are  seen  only  immediately  adjacent  to  the  vesicle.  The  membranes  appear  to 
be  smooth  and  are  fragile.  No  excystment  structure  recorded.  Vesicle  length  30  /nm;  process  length  25  ^m.  Three 
specimens  measured. 

Remarks.  The  transparent  membranes  recorded  here  possibly  envelop  the  entire  central  body,  thus 
forming  a delicate  periderm  rather  than  being  simple,  single-layer  structures  as  described  by  Cramer, 
Kanes,  Diez  and  Christopher  (1974).  More  data  are  required  to  determine  this  point. 

Genus  micrhystridium  (Deflandre  1937)  Downie  and  Sarjeant  1963 
Micrhystridium  diornamentum  Rasul  1979 

Description.  Central  vesicle  spherical,  bearing  two  types  of  process.  Some  processes  are  long,  hollow,  relatively 
few  (5-10),  simple,  smooth,  and  taper  gradually  to  a simple  acuminate  distal  termination.  The  remaining 
processes  are  much  more  numerous  (probably  more  than  50),  short,  closely  spaced,  smooth,  apparently  solid 
and  hair-like.  No  excystment  structure  recorded.  Vesicle  diameter  16  ^m;  process  length  13  p.m  and  3 ^m.  One 
specimen  measured. 

Remarks.  This  species,  originally  described  from  the  Tremadoc  of  England  has  subsequently  been 
recorded  from  the  Arenig/Llanvirn  of  North  Wales  (Booth  1979,  p.  127;  as  M.  robustum  in  part). 

Genus  multiplicisphaeridium  (Staplin  1961)  Eisenack,  Cramer,  and  Diez  1976 
Multiplicisphaeridium  maroquense  Cramer,  Allam  et  al.  1974 

Description.  Central  vesicle  hollow,  smooth,  polygonal  in  outline,  formed  from  the  merging  of  process  bases. 
Processes  are  long,  smooth,  broad,  and  widen  rapidly  proximally;  they  divide  distally  in  a characteristic  manner 
with  first-  and  second-order  lateral  branches  the  tips  of  which  recurve  sharply  to  give  a loosely  coiled  appearance. 
Processes  are  hollow  and  communicate  freely  with  the  vesicle  cavity.  No  excystment  structure  recorded.  Vesicle 
diameter  29  ju.m;  process  length  22  ju.ni.  One  specimen  measured. 

Multiplicisphaeridium  multiradiale  (Burmann  1970)  Eisenack,  Cramer,  and  Diez  1976 
Plate  17,  fig.  2 

Description.  Central  vesicle  hollow,  smooth,  polygonal  in  outline,  formed  from  the  merging  of  process  bases.  The 
processes  are  long,  smooth,  broad,  and  widen  rapidly  proximally;  they  divide  distally  by  simple  bifurcation  up  to 
the  third  order,  forming  slender  acuminate  lateral  branches.  The  processes,  varying  in  number  from  five  to  seven, 
are  hollow  and  communicate  freely  with  the  vesicle  cavity.  No  excystment  structure  recorded.  Vesicle  diameter 
22-27  fj.m;  process  length  14-17  /un.  Ten  specimens  measured. 

Remarks.  This  species  resembles  M.  maroquense  but  is  distinguished  by  the  third-order  branching  and 
the  lack  of  sharply  recurved  distal  tips  to  the  process  branches. 

Multiplicisphaeridium  rayii  Cramer,  Allam  et  al.  1974 

Description.  Central  vesicle  hollow,  smooth,  polygonal  in  outline,  formed  from  the  merging  of  process  bases.  The 
processes  are  long,  smooth,  broad  and  widen  rapidly  proximally;  they  divide  distally  into  four  or  five  digitate, 
generally  straight,  dagger-like  branches  with  rare  second-order  branching.  Processes  are  hollow  and 
communicate  freely  with  the  vesicle  cavity.  No  excystment  structure  recorded.  Vesicle  diameter  38  ju.m;  process 
length  44  jum.  One  specimen  measured. 


130 


PALAEONTOLOGY,  VOLUME  25 


Genus  polygonium  Vavrdova  1966 
‘ Polygonium'  spp. 

Plate  16,  figs.  5,  6 

Description.  Central  vesicle  hollow,  polygonal  to  sub-polygonal,  bearing  numerous,  long,  hollow,  simple 
processes  which  communicate  freely  with  the  vesicle  cavity  and  have  acuminate  distal  terminations.  Processes 
may  have  a consistent  concentric  arrangement  or  may  be  apparently  distributed  at  random.  Vesicle  wall  smooth, 
process  wall  smooth  or  rarely  granular.  Processes  always  have  wide  bases  which  thin  rapidly  to  a slender  stem, 
tapering  gradually  to  the  distal  tip.  No  excystment  structure  recorded.  Vesicle  diameter  26-35  ;y,m;  process  length 
15-20  yum.  Fifty  specimens  measured. 

Remarks.  The  acritarchs  included  here  under  the  name  ‘ Polygonium ’ embrace  a wider  variety  of 
forms  than  is  circumscribed  by  the  genus  Polygonium  Vavrdova  1966.  This  taxon  was  considered  by 
its  author  to  be  distinguished  by  always  having  processes  arranged  in  a consistent  concentric  manner 
(Vavrdova  1966,  p.  413).  Some  individuals  showing  this  feature  were  recorded  here  (PI.  16,  fig.  6)  but 
most  specimens,  otherwise  indistinguishable,  exhibit  an  apparently  random  process  arrangement 
(PI.  16,  fig.  5).  These  forms  constitute  a taxonomic  problem  since,  although  they  are  abundant  in 
Tremadoc  to  Llanvirn  strata,  no  valid  generic  name  has  yet  been  proposed  for  them.  The  term 
‘ Polygonium ’ is  used  here  to  denote  forms  with  both  concentrically  and  non-concentrically  arranged 
processes. 


Genus  priscogalea  Deunff  1961 
Priscogalea  distincta  Rasul  1974 

Plate  16,  fig.  4 

Description.  Central  vesicle  spherical  to  sub-spherical,  bearing  about  fifty  processes  which  appear  to  be 
distributed  irregularly  over  the  surface.  Processes  are  smooth,  solid,  and  taper  towards  the  distal  ends  where  they 
are  usually  multifurcate;  a few  bifurcate  or  simple  processes  may  be  present.  The  vesicle  wall  is  ornamented  with 
faint  striae  which  radiate  out  from  each  process  base.  Excystment  is  by  the  development  of  a large  polygonal  to 
sub-polygonal  polar  opening,  the  periphery  of  which  also  bears  processes.  The  operculum  is  smooth  and  without 
processes.  Vesicle  diameter  30-33  ^m;  excystment  opening  15-16  /xm;  process  length  6-8  /xm.  Two  specimens 
measured. 

Remarks.  The  striate  vesicle  wall  justifies  the  transfer  of  this  species  to  the  genus  Stelliferidium,  but 
this  should  await  examination  of  in  situ  material.  This  species,  previously  known  only  from  the 
Tremadoc,  has  been  recorded  from  the  Arenig  and  Llanvirn  of  Britain  (Booth  1979,  pp.  173,  322). 


EXPLANATION  OF  PLATE  16 
Selected  acritarchs  of  probable  Tremadoc  age 
All  figures  x 1200 

Fig.  1.  Vulcanisphaera  cirrita  Rasul.  OV/HS/1-1,  MPK  2738,  Onny  Valley,  Harnage  Shales,  phase-contrast. 

2.  Impluviculus  cf.  lenticularis  Martin.  OV/O/1-7,  MPK  2739,  Onny  Valley,  Onny  Shales,  phase-contrast. 

3.  Acanthodiacrodium/Actinotodissus  sp.  OV/UHS/1-1,  MPK  2740,  Onny  Valley,  Horderley  Sandstone.  4. 
Priscogalea  distincta  Rasul.  OV/A/2b-5,  MPK  2741 , Onny  Valley,  Alternata  Limestone.  5-6.  ‘ Polygonium ’ 
spp.  5.  OV/A/la-1,  MPK  2742.  6.  OV/A/la-1,  MPK  2743,  both  from  the  Alternata  Limestone  of  the 
Onny  Valley. 


PLATE  16 


TURNER,  probable  Tremadoc  acritarchs 


132 


PALAEONTOLOGY,  VOLUME  25 


Genus  saharidia  Combaz  1967 
Saharidia  fragile  (Downie  1958)  Combaz  1967 
Plate  15,  fig.  6 

Description.  Central  vesicle  circular  in  outline,  wall  thin  (<  0-5  /xm),  fragile,  and  ornamented  with  irregularly 
sized  and  spaced  grana.  Concentric  folds  are  developed  in  the  wall  adjacent  to  the  periphery.  Excystment  is  by 
the  development  of  a small  central  pylome  but  often  such  openings  are  not  apparent.  Vesicle  diameter  35-70  jum; 
pylome  diameter  9-15  /u.m.  Ten  specimens  measured. 

Genus  stelliferidium  Deunff,  Gorka,  and  Rauscher  1974 
Stelliferidium  cortinulum  (Deunff  1961)  Deunff,  Gorka,  and  Rauscher  1974 

Description.  Central  vesicle  spherical  to  sub-spherical;  the  wall  is  thick  (1-2  ^m)  and  bears  15-20  smooth 
processes  which  appear  to  be  distributed  irregularly  over  the  vesicle;  these  processes  are  hollow  with  a solid 
proximal  plug  separating  the  process  interior  from  the  vesicle  cavity;  from  the  base  they  taper  gradually  to  a 
bifurcate  or  multifurcate  distal  termination.  The  vesicle  wall  is  ornamented  with  faint  striae  which  radiate  out 
from  the  base  of  each  process.  Excystment  is  by  the  development  of  a large  circular  to  sub-polygonal  polar 
opening,  the  periphery  of  which  always  lacks  processes.  The  operculum  is  smooth  and  is  also  without  processes. 
Vesicle  diameter  29-36  ^m;  excystment  opening  15-23  jixm;  process  length  5-9  ^m.  Fifteen  specimens  measured. 

Stelliferidium  stelligerum  Deunlf,  Gorka,  and  Rauscher  1974 
Plate  15,  fig.  3 

Description.  Central  vesicle  spherical  to  sub-spherical,  thick-walled  (1-2  /xm),  and  bears  about  sixty  processes 
which  appear  to  be  distributed  irregularly  over  the  surface.  Processes  are  hollow  with  a solid  proximal  plug 
separating  the  process  interior  from  the  vesicle  cavity;  they  are  smooth  and  taper  gradually  to  a simple  acuminate 
or  bifurcate  distal  termination.  Processes  tend  to  be  longest  at  the  antapex  becoming  progressively  shorter 
towards  the  polar  opening.  Excystment  is  by  the  development  of  a large,  circular  to  sub-circular  opening,  the 
periphery  of  which  bears  short,  generally  bifurcating  processes.  The  operculum,  which  is  commonly  preserved  in 
situ,  is  granular  and  without  processes.  The  vesicle  wall  is  ornamented  with  thick  prominent  striae  that  radiate 
out  from  each  process  base.  Vesicle  diameter  3 1 -37  ^m;  excystment  opening  1 8-2 1 ^m;  process  length  11-15  /xm. 
Fifteen  specimens  measured. 


Genus  striatotheca  Burmann  1970 
Striatotheca  frequerts  Burmann  1970 

Description.  Central  vesicle  hollow,  quadrate  in  outline,  strongly  compressed.  Each  angle  bears  a single  long, 
broad,  simple  process  which  tapers  gradually  to  a generally  rounded  distal  termination.  The  hollow  processes  lie 
in  the  same  plane  as  the  central  body  and  are  in  free  communication  with  the  vesicle  cavity.  The  vesicle  wall  is 
ornamented  with  fine  striae  that  are  approximately  parallel  to  the  vesicle  margins  around  the  periphery  but 
become  concave  towards  the  centre;  these  striae  pass  on  to  the  processes  but  die  away  distally.  No  excystment 
structure  recorded.  Dimensions  of  central  vesicle  33-37  /xmx  31-36  jam;  process  length  10-16  fxm.  Five 
specimens  measured. 


EXPLANATION  OF  PLATE  17 
Selected  acritarchs  of  Arenig/Llanvirn  age 
All  figures  x 1200 

Fig.  1.  Marrocanium  simplex  Cramer,  Kanes  et  al.  NS/4-4,  MPK  2744,  Chatwall,  Harnage  Shales.  2.  Multi- 
plicisphaeridium  multiradiale  (Burmann)  Eisenack  et  al.  OV/A/2b-5,  MPK  2745,  Onny  Valley,  Alternata 
Limestone.  3.  Striatotheca  quieta  (Martin)  Rauscher.  NS/4-4,  MPK  2746,  Chatwall,  Harnage  Shales.  4. 
Coryphidium  australe  Cramer  and  Diez.  OV/MHS/1-3,  MPK  2747,  Onny  Valley,  Horderley  Sandstone.  5. 
Dicrodiacrodium  normale  Burmann.  NS/4-1,  MPK  2748,  Chatwall,  Harnage  Shales.  6.  Tunisphaeridium 
eligmosum  Vavrdova?  OV/A/lb-1,  MPK  2749,  Onny  Valley,  Alternata  Limestone. 


PLATE  17 


TURNER,  Areniq/Llanvirn  acritarchs 


134 


PALAEONTOLOGY,  VOLUME  25 


Striatotheca  principalis  Burmann  1970 

Description.  Similar  to  S.  frequens  in  over-all  morphology  but  is  larger,  has  long,  slender  processes  with 
acuminate  distal  terminations  and  bears  a vesicle  ornament  of  coarse,  widely  spaced  striae.  No  excystment 
structure  recorded.  Vesicle  dimensions  40-44  /xm  x 36-38  (im;  process  length  20-31  /xm.  Four  specimens 
measured. 

Striatotheca  principalis  var.  parva  Burmann  1970 

Description.  Similar  to  S.  principalis  in  over-all  morphology,  this  variety  is  distinguished  by  its  much  smaller  size. 
It  differs  from  S.  frequens  in  having  slender  processes  with  acuminate  distal  terminations,  and  a vesicle  ornament 
of  coarse,  widely  spaced  striae.  No  excystment  structure  recorded.  Vesicle  dimensions  27-31  /xm  x 20-23  /xm; 
process  length  15-20  pm.  Five  specimens  measured. 

Striatotheca  quieta  (Martin  1969)  Rauscher  1974 
Plate  17,  fig.  3 

Description.  Central  vesicle  hollow,  quadrate  in  outline,  strongly  compressed;  each  angle  bears  an  extremely 
short,  simple  process  which  tapers  to  a rounded  distal  termination.  The  hollow  processes  lie  in  the  same  plane  as 
the  central  body  and  communicate  freely  with  the  vesicle  cavity.  The  vesicle  wall  is  ornamented  with  line,  closely 
spaced  striae  which  are  approximately  parallel  to  the  vesicle  margins  around  the  periphery  but  become  concave 
towards  the  centre;  these  striations  continue  on  to  the  process  wall  almost  out  to  the  distal  tip.  No  excystment 
structure  recorded.  Vesicle  dimensions  29-32  ^m  x 33-35  /xm;  process  length  3-5  /xm.  Twelve  specimens 
measured. 

Remarks.  Records  of  this  species  (as  Veryhachium  quietum)  from  the  Silurian  of  Belgium  (Martin 
1969)  are  interpreted  by  the  present  author  as  reworked. 

Genus  timofeevia  Vanguestaine  1978 
Timofeevia  phosphoritica  Vanguestaine  1978 

Description.  Central  vesicle  spherical  to  sub-spherical,  wall  smooth,  divided  by  raised  ridges  into  about  twenty 
polygonal  fields;  the  junctions  of  the  ridges  bear  smooth  processes  which  taper  gently  to  a bifurcate  or 
multifurcate  distal  termination.  No  excystment  structure  recorded.  Vesicle  diameter  32  /xm;  process  length  9-11 
urn;  diameter  of  polygonal  fields  9-11  /xm.  One  specimen  measured. 

Genus  trichosphaeridium  Timofeev  1966 
Trichosphaeridium  annolovaense  Timofeev  1966 
Plate  15,  fig.  5 

Description.  Central  vesicle  spherical  to  sub-spherical  but  always  compressed,  wall  smooth,  moderately  thick 
(about  1 pm)  with  compression  folds  developed.  The  vesicle  bears  more  than  100  short,  solid,  smooth,  simple 
hairlike  processes  whose  distal  terminations  may  be  evexate  or  acuminate.  No  excystment  structure  recorded. 
Vesicle  diameter  42-49  /xm;  process  length  3-4  /xm.  Six  specimens  measured. 

Genus  tunisphaeridium  Deunff  and  Evitt  1968 
Tunisphaeridium  eligmosum  Vavrdova?  1973 
Plate  17,  fig.  6 

Description.  Central  vesicle  hollow,  sub-polygonal  in  outline,  wall  smooth,  bearing  15-20,  long,  cylindrical, 
smooth  processes  which  widen  proximally  and  divide  distally  by  means  of  simple  bifurcation  up  to  the  fifth 
order;  the  distal  terminations  of  these  branches  are  long,  slender,  curved,  and  sometimes  appear  to  join  those  of 
adjacent  processes  forming  an  anastomosing  network  of  fine  filaments.  Processes  are  hollow  and  communicate 
freely  with  the  vesicle  cavity.  No  excystment  structure  recorded.  Vesicle  diameter  36  /xm;  process  length  28  /xm. 
One  specimen  measured. 

Remarks.  The  specific  assignment  is  not  certain  since  the  branching  pattern  of  T.  eligmosum  is 
described  by  Vavrdova  as  palmate  rather  than  bifurcate  as  recorded  here. 


TURNER:  REWORKED  ACRITARCHS 


35 


Genus  vulcanisphaera  (Deunff  1961)  Rasul  1976 
Vulcanisphaera  africana  Deunff  1961 

Description.  Central  vesicle  spherical  to  sub-spherical,  wall  smooth  or  granular,  bearing  50-100  processes  which 
arise  from  hollow  conical  projections  having  a solid  tip;  normally  three  processes  arise  from  a common  base  in 
this  way,  more  rarely  two  or  four.  Processes  are  slender  and  taper  gradually  to  a bifurcate  distal  termination.  No 
excystment  structure  recorded.  Vesicle  diameter  32-49  ^m,  process  length  10-14  /u.m.  Three  specimens  measured. 

Remarks.  The  distal  bifurcations  are  extremely  fine  and  delicate  and  are  often  broken  off  giving  the 
appearance  of  a simple  acuminate  termination.  Published  records  of  this  species  are  restricted  to 
strata  of  Tremadoc  age;  the  present  author  has  identified  the  taxon  in  assemblages  from  Britain  which 
on  their  palynological  content  are  of  probable  Arenig  age  (unpublished  data).  It  is  thus  possible  that 
this  form  ranges  above  the  top  of  the  Tremadoc. 

Vulcanisphaera  cirrita  Rasul  1976 
Plate  16,  fig.  1 

Description.  Central  vesicle  spherical  to  sub-spherical  with  50-100  processes  arising  from  hollow  projections 
which  have  a solid  tip.  The  number  of  processes  sharing  a common  base  in  this  way  varies  from  two  to  five. 
Processes  are  slender  and  taper  only  slightly  towards  the  distal  tip  where  they  branch  into  numerous  delicate 
thread-like  branches.  The  branches  of  adjacent  process  groups  may  unite  to  form  a complex  anastomosing 
network  of  fine  filaments.  No  excystment  structure  recorded.  Vesicle  diameter  37-48  ^m;  process  length  9-11 
/j.m.  Three  specimens  measured. 

Remarks.  Originally  described  from  the  Tremadoc  of  Shropshire,  this  species  has  been  subsequently 
recorded  from  the  Arenig/Lower  Llanvirn  of  North  Wales  (Booth  1979,  p.  190). 


DISTRIBUTION  OF  THE  REWORKED  MATERIAL 

The  over-all  pattern  of  reworking  within  the  Onny  Valley  section  is  shown  in  text-fig.  5.  Reworked 
acritarchs  first  appear  in  the  Harnage  Shales  which  represent  the  onset  of  fine-grained  sedimentation. 
These  acritarchs  include  species  of  both  Tremadoc  and  Arenig/Llanvirn  age,  indicating  that  strata  of 
both  ages  were  being  eroded  to  supply  sediment  to  the  shelf  during  early  Caradoc  time.  Towards  the 
top  of  the  Harnage  Shales  and  in  the  lower  Horderley  Sandstones,  Arenig/Llanvirn  forms  dominate 
the  reworked  portions  of  the  assemblages;  only  minor  elements  of  probable  Tremadoc  age  are 
present.  A possible  explanation  of  this  is  that  widespread  erosion  of  Arenig/Llanvirn  sediments  was 
occurring  but  that  only  a small  area  of  Tremadoc  rock  was  exposed.  From  the  base  of  the  Harnage 
Shales  up  to  the  middle  Horderley  Sandstone,  reworked  acritarchs  consistently  comprise  10-20%  of 
the  total  assemblages.  Above  this  level  the  proportion  of  reworked  specimens  in  the  sediments 
increases  greatly  to  as  much  as  70%.  This  large  and  sudden  increase  in  reworking  is  associated  with  an 
increase  in  the  percentages  of  both  Tremadoc  and  probable  Tremadoc  forms  present.  The 
simultaneous  increase  in  the  abundance  of  these  two  Categories  tends  to  substantiate  the  Tremadoc 
age  suggested  for  most  individuals  placed  in  Category  2.  A high  percentage  of  reworked  acritarchs  is 
evident  until  the  upper  part  of  the  Cheney  Longville  Flags,  always  with  taxa  of  Tremadoc  and 
probable  Tremadoc  age  predominating.  The  maximum  level  of  reworking  occurs  in  the  Alternata 
Limestone,  where  up  to  94%  of  the  acritarchs  are  derived,  the  contemporaneous  Caradoc  forms  being 
swamped  out.  The  percentage  of  Arenig/Llanvirn  forms  fluctuates  throughout  the  middle  of  the 
Caradoc  sequence  but  is  always  small  (text-fig.  5).  The  large  numbers  of  reworked  acritarchs  of 
Categories  1 and  2 that  are  present  from  the  middle  Horderley  Sandstone  through  to  the  upper 
Cheney  Longville  Flags  suggest  that  an  acritarch-rich  source  rock  of  Tremadoc  age  was  extensively 
breached  and  continued  to  be  eroded  over  a substantial  period  of  time. 

In  the  lower  Acton  Scott  Beds  reworked  acritarchs  constitute  a mere  1 or  2%  of  the  total 
assemblage  and  remain  at  this  much-reduced  level  up  to  the  top  of  the  succession.  This  reduction  in 
reworking  coincides  with  a return  to  a low-energy  mudstone  environment. 


136 


PALAEONTOLOGY,  VOLUME  25 


If  the  majority  of  individuals  placed  in  Category  2 are  accepted  as  having  originated  in  the 
Tremadoc  then  an  interesting  pattern  to  the  reworking  emerges  (text-fig.  5).  The  distribution  of  taxa 
is  essentially  inverted,  reflecting  successive  erosion  of  progressively  older  source  sediments  during 
Caradoc  time.  It  should  be  noted  that  this  pattern  is  modified  by  the  relative  abundance  of  Tremadoc 
forms  at  the  base  of  the  Harnage  Shales.  As  discussed  above,  this  clearly  indicates  early  erosion  of  a 
Tremadoc  source  rock;  however,  the  paucity  of  Tremadoc  acritarchs  in  the  overlying  Harnage  Shales 


text-fig.  5.  Showing  the  strati  graphical  distribution  of  samples  from  the  Caradoc  type  section  together  with  the 
ages  and  percentages  of  the  reworked  acritarchs  recorded  from  each  horizon. 


TURNER:  REWORKED  ACRITARCHS 


137 


and  Horderley  Sandstone  shows  that  reworking  from  this  source  was  subsequently  suppressed 
although  not  eliminated.  Any  explanation  of  this  diminution  in  erosion  of  Tremadoc  sediments 
would  be  purely  conjectural  at  present.  An  inversion  of  this  type,  with  younger  material  redeposited 
in  the  lower  horizons  and  older  forms  appearing  in  the  overlying  strata,  would  be  expected  where 
relatively  undisturbed  sediments  were  being  eroded  and  quickly  laid  down  again.  It  may  be  assumed 
that  the  rocks  discussed  here  would  have  been  little  altered  by  Caradoc  time  since  the  Ordovician  was 
a period  of  tectonic  quiescence  in  this  region  (Earp  and  Hains  1971,  p.  89). 

Samples  from  the  lower  Caradoc  of  the  Chatwall  district  (text-fig.  2)  contain  abundant  reworked 
acritarchs  of  Arenig/Llanvirn  age  with  rare  Tremadoc  and  probable  Tremadoc  forms,  a distribution 
similar  to  that  in  the  type  section.  Since  the  Caradoc  sequence  is  much  less  complete  at  Chatwall  and 
acritarchs  are  rare  above  the  Harnage  Shales,  these  occurrences  are  not  discussed  in  detail  here  but 
they  demonstrate  that  the  reworking  is  not  a local  phenomenon  restricted  to  the  Onny  Valley. 

PROVENANCE  OF  THE  REWORKED  MATERIAL 

The  Tremadoc  and  Arenig/Llanvirn  acritarchs  encountered  in  the  Caradoc  type  section  are 
extremely  well  preserved  suggesting  that  individuals  were  transported  only  short  distances  and 
underwent  rapid  reburial.  Possible  source  rocks  must  therefore  have  been  located  close  to  the  site  of 
redeposition.  It  is  unlikely  that  the  reworked  material  was  derived  from  the  west  since  this  was  itself 
an  area  of  deposition  in  Caradoc  time.  To  the  east  and  south,  the  Midland  Platform  formed  a stable 
block  during  the  Palaeozoic  (text-fig.  6);  Tremadoc  rocks  are  widespread  over  this  platform  although 
Arenig/Llanvirn  strata  are  practically  unknown  (Richardson  and  Rasul  1978,  p.  37).  To  the  south- 
east of  Shropshire,  probably  close  to  the  ancient  margin  of  the  Midland  Platform,  great  thicknesses  of 
Tremadoc  sediments  exist.  Rocks  of  Arenig/Llanvirn  age  were  possibly  deposited  in  such  peripheral 
areas  but  no  traces  have  yet  been  found.  Acritarch  bearing  Tremadoc  and  Arenig/Llanvirn  strata  are 
known  from  North  Wales  (Booth  1979;  author’s  own  unpublished  data)  and  from  the  north  of 
England  (Booth  1979;  Downie  and  Soper  1972),  but  the  distances  involved  here  are  considerable  and 
a closer  source  is  considered  more  likely.  No  sedimentological  evidence  exists  to  indicate  the  direction 
of  transport,  but  it  appears  probable  that  the  reworked  acritarchs  were  derived  from  the  Midland 
Platform  to  the  east  or  south-east.  This  is  consistent  with  the  available  information  on  early  Caradoc 
palaeocurrents  (Williams  1969,  p.  259,  fig.  8). 

MECHANISM  OF  REWORKING 

A widely  accepted  explanation  of  the  mechanism  for  reworking  of  palynomorphs  is  that  they  were 
eroded  and  transported  while  encapsulated  within  particles  of  pre-existing  sediments  and  so  were 
protected  from  damage.  If  such  recycled  rock  particles  are  present  in  a sediment  they  should  be  visible 
under  microscopic  examination  (Richardson  and  Rasul  1978,  p.  37).  In  the  Onny  Valley, 
lithologically  diverse  sediments  such  as  the  Harnage  Shales,  Horderley  Sandstone,  and  Alternata 
Limestone  all  contain  numerous  reworked  acritarchs.  Thin-sections  of  samples  from  these 
formations  were  prepared  and  examined  to  see  if  such  lithoclasts  could  be  recognized  but  none  was 
observed,  the  sediments  presenting  a more  or  less  uniformly  fine-grained  appearance;  the  dimensions 
of  sediment  grains  are  between  5 and  100  /x m with  the  vast  majority  being  between  30  and  60  ju.m. 
Thus  the  grains  are  at  most  only  slightly  larger  than  reworked  acritarchs  recovered  from  the  same 
samples.  In  addition,  although  no  acritarchs  were  recognized  in  thin  section,  the  sediments  are  clearly 
organic-rich  and  the  abundant  organic  matter  visible  is  trapped  in  the  interstices.  These  factors  make 
it  unlikely  that  acritarchs  were  reworked  in  an  encapsulated  state,  the  evidence  suggesting  rather  that 
they  were  eroded  and  redeposited  as  discrete  sedimentary  particles.  This  hypothesis  is  supported  by 
the  distribution  pattern  of  the  reworking  which  is  unaffected  by  changes  in  the  type  of  sediment  being 
deposited  (text-fig.  5).  Similar  reworked  acritarchs  are  found  in  comparable  numbers  in  sandstones, 
limestones,  and  flaggy  micaceous  siltstones.  This  alone  suggests  that  the  acritarchs  were  being  intro- 
duced into  the  sediment/water-body  system  independently  of  the  non-organic  sediment  particles. 


138 


PALAEONTOLOGY,  VOLUME  25 


If  it  is  true  that  the  reworked  acritarchs  were  transported  as  individual  particles  then  further 
conclusions  can  be  drawn.  Considering  the  excellent  state  of  preservation  of  most  reworked 
specimens,  it  is  probable  that  dissolution  of  the  parent  rock  was  both  easy  and  rapid.  Since  an 
indurated  sediment  would  resist  erosion,  the  Tremadoc  and  Arenig/Llanvirn  rocks  being  eroded  were 
probably  at  most  only  partly  lithified.  The  retention  of  the  most  delicate  morphological  features  on 
many  reworked  specimens  suggests  that  erosion  and  transport  were  not  only  rapid  but  did  not  take 
place  in  a sub-aerial  environment.  Structures  such  as  fine  distal  terminations  of  processes  in 


text-fig.  6.  A suggested  palaeogeographic  reconstruction  of  the  British  Isles  during  Caradoc  time  (simplified, 

after  Williams  1969). 


TURNER:  REWORKED  ACRITARCHS 


139 


HARNAGIAN 


LONG  VILLI  AN 

Developing  Regression 


text-fig.  7.  Hypothetical  diagrammatic  cross-section  through  southern  Shropshire  in  Caradoc  time  showing 
the  postulated  sequence  of  events.  The  Caradoc  succession  is  simplified  for  the  sake  of  clarity. 


140 


PALAEONTOLOGY,  VOLUME  25 


Vulcanisphaera  cirrita  (PI.  16,  fig.  1)  and  the  delicate  membranes  of  Marrocanium  simplex  (PI.  17, 
fig.  1)  would  have  been  unlikely  to  survive  long  in  turbulent  conditions  without  the  protection 
provided  by  encapsulation.  Even  if  mechanical  damage  had  been  avoided,  such  features  would  have 
suffered  rapid  oxidation  and  disintegration.  It  is  therefore  postulated  that  erosion  and  redeposition  of 
these  pre-existing  rocks  took  place  in  a shallow  marine  environment;  wave  and  current  action  are 
considered  the  most  likely  agents  for  eroding  and  dispersing  the  unconsolidated  sediments  involved. 
Under  these  circumstances  the  enclosed  acritarchs  would  be  released  directly  into  the  sea,  affording 
them  the  means  both  of  protection  and  rapid  dispersal  and  reburial  (text-fig.  7).  This  agrees  with  the 
limited  sedimentological  evidence  available  which  suggests  that  the  Caradoc  rocks  were  deposited  in 
a shallow  marine  environment,  possibly  with  off-shore  barriers  but  with  no  estuaries  present  to  have 
provided  a potential  source  of  reworked  acritarchs  from  aerially  exposed  sediments  (Hurst  1979a,  p. 
36).  Hurst  (op.  cit.,  19796)  has  shown  that  in  the  Onny  Valley  the  sequence  from  the  upper  part  of  the 
Horderley  Sandstone  to  the  basal  Acton  Scott  Beds  represents  a regressive  phase,  and  that  the 
deposits  of  this  interval  were  greatly  affected  by  storm  surge  activity.  The  shallowing  of  the  water 
body  would  have  led  to  an  increasingly  high-energy  regime,  while  individual  surges  would  have 
resulted  in  mass  sediment  movement  with  rapid  redeposition  on  the  cessation  of  these  geologically 
ephemeral  events  (text-fig.  7).  Text-fig.  5 shows  that  the  high-energy  environment  which  resulted  in 
the  deposition  of  the  upper  Horderley  Sandstone,  Alternata  Limestone,  and  Cheney  Longville  Flags, 
coincided  with  the  period  of  greatest  acritarch  reworking.  Above  the  base  of  the  Actonian  the 
percentage  of  reworked  acritarchs  is  drastically  reduced  and  it  was  at  this  time  that  storm  swells 
ceased  to  exert  any  significant  effect  (Hurst  19796,  p.  196,  Table  1).  Unfortunately  Hurst’s  studies  do 
not  extend  down  into  the  Soudleyan  so  it  is  uncertain  how  close  the  correlation  is  between  the 
increasing  energy  levels  and  the  first  appearance  of  high  levels  of  reworking.  None  the  less  there  is 
clearly  a link  between  the  high-energy  regime  and  abundant  reworking,  supporting  the  view  that 
unlithified  Tremadoc  and  Arenig/Llanvirn  sediments  were  being  eroded  in  a shallow  marine 
environment. 

Much  has  been  written  in  recent  years  on  the  effects  of  storm  surges  on  contemporaneous  marine 
sediments  (Brenner  and  Davis  1973;  Reineck  and  Singh  1972,  1973),  but  there  appear  to  be  few  data 
available  on  the  effects  of  such  events  on  soft  pre-existing  sediments  at  the  water/substrate  interface. 
Although  little  consideration  has  been  given  in  the  literature  to  the  possibility  of  reworking  from  such 
sediments,  the  situation  visualized  here  is  not  unique.  For  example.  Quaternary  clays  in  the  Moray 
Firth,  Scotland,  contain  extensive  assemblages  of  reworked  late  Jurassic  and  early  Cretaceous 
microfossils,  particularly  organic-walled  microplankton  (Owens  and  Marshall,  1978,  pp.  24-26). 
This  unoxidized  material  is  well  preserved  and  is  clearly  derived  from  unconsolidated  Jurassic  and 
Cretaceous  shales  and  clays  upon  which  the  Quaternary  rests  in  this  area  (Dr.  R.  Harland,  pers. 
comm.).  The  physical  state  of  these  Mesozoic  sediments  suggests  that  reworking  within  lithic  clasts 
would  have  been  unlikely  and  redeposition  of  the  microfossils  as  discrete  particles  is  considered 
probable. 


CONCLUSIONS 

The  type  Caradoc  rocks  of  Shropshire  yield  abundant  acritarch  assemblages  which  contain  Caradoc 
species  admixed  with  reworked  Tremadoc  and  Arenig/Llanvirn  taxa. 

The  vertical  distribution  of  reworked  forms  reveals  essentially  an  inverted  stratigraphy  with 
Arenig/Llanvirn  acritarchs  predominating  in  the  lower  horizons  while  the  older  Tremadoc  species 
become  the  most  abundant  forms  in  the  middle  part  of  the  sequence. 

The  lack  of  visible  lithoclasts  in  thin  sections  of  these  rocks  suggests  that  the  microfossils  were 
introduced  into  the  sediment  body  as  individuals  and  were  not  encapsulated  in  redeposited  fragments 
of  pre-existing  rocks;  this  is  substantiated  by  the  fact  that  the  presence  and  abundance  of  reworked 
acritarchs  appears  to  be  entirely  independent  of  lithotype. 

The  excellent  state  of  preservation  of  the  reworked  acritarchs  indicates  that  they  underwent  little 
transportation  before  reburial.  It  also  suggests  that  the  parent  sediments  were  relatively 


TURNER:  REWORKED  ACRITARCHS 


141 


unconsolidated  and  that  their  erosion  was  caused  by  marine  action  in  a shallow-water  environ- 
ment. 

The  high  percentages  of  reworking  in  the  middle  part  of  the  section  are  partly  related  to  the 
increasing  erosion  of  particularly  acritarch-rich  Tremadoc  rocks.  In  addition,  this  sequence 
represents  a regressive  phase  with  storm  surges  having  profound  effects  on  the  shallowing  water 
body.  Such  high-energy  events  would  have  greatly  increased  the  erosion  rate  of  the  unlithified 
sediments  exposed  at  the  sea  bed.  A state  of  continuous  low-level  erosion  and  acritarch  reworking  is 
envisaged,  punctuated  by  periodic  intense  turbidity  associated  with  an  upsurge  in  the  rate  of  release 
of  pre-existing  microfossils. 

Acknowledgements.  I am  most  grateful  to  Professor  C.  Downie  for  his  encouragement  during  this  work  and  for 
his  advice  and  constructive  comments.  I thank  Dr.  G.  A.  Booth  for  access  to  his  unpublished  information  and  for 
much  useful  and  stimulating  discussion.  My  thanks  also  go  to  Dr.  B.  Owens,  Dr.  W.  H.  C.  Ramsbottom,  and 
Dr.  W.  A.  M.  Jenkins  who  read  the  manuscript  and  suggested  improvements. 


REFERENCES 

booth,  G.  a.  1979.  Lower  Ordovician  acritarchs  from  successions  in  England  and  North  Wales.  University  of 
Sheffield,  unpublished  Ph.D.  thesis.  383  pp. 

brenner,  r.  l.  and  davies,  d.  k.  1973.  Storm  generated  coquinoid  sandstones:  genesis  of  high-energy  marine 
sediments  from  the  Upper  Jurassic  of  Wyoming  and  Montana.  Bull.  geol.  Soc.  Am.  84,  1685-1698. 
burmann,  G.  1968.  Diacrodien  aus  dem  unteren  Ordovizium.  Paldont.  Abh.  B,  2,  639-652. 

— 1970.  Weitere  organische  Mikrofossilien  aus  dem  unteren  Ordovizium.  Ibid.  3,  289-347. 

combaz,  a.  1967.  Un  microbios  du  Tremadocien  dans  un  sondage  d’Hassi-Massaoud.  Act.  Soc.  linn.  Bordeaux , 
B,  104,  1-26. 

cramer,  f.  h.,  allam,  b.,  kanes,  w.  h.  and  diez,  m.  d.  c.  R.  1974.  Upper  Arenigian  to  Lower  Llanvirnian 
acritarchs  from  the  subsurface  of  the  Tadla  Basin  of  Morocco.  Palaeontographica,  B,  145,  182-190. 

— and  diez,  m.  d.  c.  R.  1976.  Seven  new  Late  Arenigian  species  of  the  Acritarch  Genus  Coryphidium 
vavrdova,  1972.  Paldont.  Z.  50,  201-208. 

— 1977.  Late  Arenigian  (Ordovician)  acritarchs  from  Cis-Saharan  Morocco.  Micropaleontology , 23, 
339-360. 

— kanes,  w.  H.,  diez,  M.  d.  c.  R.  and  Christopher,  R.  A.  1974.  Early  Ordovician  Acritarchs  from  the 
Tadla  Basin  of  Morocco.  Palaeontographica , B,  146,  57-64. 

dean,  w.  t.  and  martin,  f.  1978.  Lower  Ordovician  acritarchs  and  Trilobites  from  Bell  Island,  Eastern 
Newfoundland.  Bull.  geol.  Surv.  Can.  284,  1-35. 

deflandre,  G.  1937.  Microfossils  des  silex  cretaces,  1 1 . Flagelles  incertae  sedis.  Hystrichospaeridees.  Sarcodines. 
Organismes  divers.  Annls.  Paleont.  26,  51-103. 

— and  deflandre-rigaud,  m.  1961.  Nomenclature  et  systematique  des  Hystrichospheres  (s.l.)  Observations 
et  rectifications.  Lab.  Micropaleont.  E.P.H.E.  Inst.  Paleontol.  Mus.  Paris , Multicop.  14  pp. 

deunff,  j.  1961.  Un  microplancton  a Hystrichospheres  dans  le  Tremadoc  du  Sahara.  Revue  Micropaleont. 
4,  37-52. 

— and  evitt,  w.  r.  1968.  Tunis phaeridium\  A new  acritarch  genus  from  the  Silurian  and  Devonian.  Standford 
Univ.  Pubis.  Geol.  Sci.  12,  1-13. 

— gorka,  h.  and  rauscher,  r.  1 974.  Observations  nouvelles  et  precisions  sur  les  Acritarches  a large 
ouverture  polaire  du  Paleozoique  inferieur.  Geobios,  7,  5-18. 

downie,  c.  1958.  An  assemblage  of  microplankton  from  the  Shineton  Shales  (Tremadocian).  Proc.  Yorks,  geol. 
Soc.  31,  331-349. 

— evitt,  w.  R.  and  sarjeant,  w.  a.  s.  1963.  Dinoflagellates,  Hystrichospheres  and  the  classification  of  the 
acritarchs.  Stanford  Univ.  Pubis.  Geol.  Sci.  7,  2-16. 

— and  sarjeant,  w.  a.  s.  1963.  On  the  interpretation  and  status  of  some  Hystrichosphere  genera. 
Palaeontology,  6,  83-96. 

— and  soper,  N.  J.  1972.  Age  of  the  Eycott  Volcanic  Group  and  its  conformable  relationship  to  the  Skiddaw 
Slates  in  the  English  Lake  District.  Geol.  Mag.  109,  259-268. 

earp,  j.  r.  and  haines,  b.  a.  1971.  British  Regional  Geology:  The  Welsh  Borderland.  Inst.  Geol.  Sci.  1 18  pp.  pub. 
H.M.S.O. 


142 


PALAEONTOLOGY,  VOLUME  25 


eisenack,  a.  1955.  Chitinozoen,  Hystrichospharen  und  andere  mikrofossilien  aus  dem  Beyrichia- Kalk. 
Senckenberg.  leth.  36,  157-188. 

— cramer,  F.  fi.  and  diez,  m.  d.  c.  R.  1976.  Katalog  der  fossilen  Dinoflagellaten,  Hystrichospharen  und 
verwandten  Mikrofossilien , 4,  Acritarcha  part  2,  Stuttgart.  863  pp. 

evitt,  w.  r.  1963.  A discussion  and  proposals  concerning  fossil  dinoflagellates,  hystrichospheres  and  acritarchs. 

Proc.  natn.  Acad.  Sci.  U.S.A.  49  (2),  158-164,  298-302. 
funkhouser,  j.  w.  1969.  Factors  that  affect  sample  reliability.  In  tschudy,  r.  h.  and  scott,  r.  a.  (ed.). 

Aspects  of  Palynology,  John  Wiley  and  Sons.  pp.  97-102. 
gorka,  h.  1967.  Quelques  nouveaux  acritarches  des  silexites  du  Tremadocien  superieur  de  la  region  de  Kielce 
(Montagne  de  Ste  Croix,  Pologne).  Cahiers  de  Micropaleont.  Series  1,  6,  Arch.  orig.  Centre  Documen. 
C.N.R.S.  441,  1-8. 

hurst,  j.  m.  1979a.  The  environment  of  deposition  of  the  Caradoc  Alternata  Limestones  and  contiguous  deposits 
of  Salop.  Geol.  J.  14,  15-40. 

— 19796.  Evolution,  succession  and  replacement  in  the  type  Upper  Caradoc  (Ordovician)  benthic  faunas  of 
England.  Palaeogeogr.  Palaeoclimatol.  Palaeoecol.  27,  189-246. 

— 1979c.  The  stratigraphy  and  brachiopods  of  the  upper  part  of  the  type  Caradoc  of  South  Salop.  Bull.  Br. 
Mus.  nat.  Hist.  (Geol.),  32,  (4),  183-304. 

jenkins,  w.  a.  m.  1967.  Ordovician  chitinozoa  from  Shropshire.  Palaeontology,  10,  436-488. 
lapworth,  c.  1916.  In  Palaeontological  Work,  summ.  progr.  Geol.  Surv.  Gt.  Britain  1915,  36-38. 
loeblich,  A.  R.  jr.  and  tappan,  h.  1969.  Acritarch  excystment  and  surface  ultrastructure  with  descriptions  of 
some  Ordovician  taxa.  Rev.  Esp.  Micropaleontol.  1,  45-57. 

— 1976.  Some  new  and  revised  organic-walled  phytoplankton  microfossil  genera.  J.  Paleont.  50, 
301-308. 

— 1978.  Some  Middle  and  Late  Ordovician  microphytoplankton  from  central  North  America.  Ibid.  52, 
1233-1287. 

martin,  f.  1966.  Les  acritarches  de  Sart-Bernard,  (Ordovicien-belge).  Bull.  Soc.  beige  Geol.  Paleont.  Hydrol.  74 
(for  1965),  423-444. 

— 1969.  Les  acritarches  de  l’Ordovicien  et  du  Silurien  beiges.  Determination  et  valeur  stratigraphique.  Mem. 
Inst.  r.  Sci.  nat.  Belg.  160  (for  1968),  175  pp. 

— 1973.  Les  acritarches  de  l’Ordovicien  inferieur  de  la  Montagne  Noire  (Herault,  France).  Bull.  Inst.  r.  Sci. 
nat.  Belg.  48  (for  1972),  1-61. 

— 1977.  Acritarches  du  Cambro-Ordovicien  du  Massif  du  Brabant,  Belgique.  Ibid.  51  (for  1975),  1-28. 

— michot,  p.  and  vanguestaine,  m.  1970.  Le  Flysch  Caradocien  d’Ombret.  Annls.  Soc.  geol.  Beige.  93, 
337-350. 

murchison,  r.  i.  1839.  The  Silurian  System.  Murray,  London  (2  vols.).  768  pp. 

owens,  b.  and  marshall,  j.  1978.  Micropalaeontological  biostratigraphy  of  samples  from  around  the  coasts  of 
Scotland.  Rep.  Inst.  Geol.  Sci.  78/20.  35  pp. 

Paris,  f.  and  deunff,  j.  1970.  Le  paleoplancton  Llanvirnien  de  la  Roche-au- Merle,  (Commune  de  Vieux-Vy- 
Sur-Couesnon,  Ille-et-Vi-laine).  Bull.  Soc.  geol.  miner.  Bretagne,  C,  2,  25-43. 
rasul,  s.  M.  1974.  The  Lower  Palaeozoic  acritarchs  Priscogalea  and  Cymatiogalea.  Palaeontology,  17,  41-63. 

— 1976.  New  species  of  the  genus  Vulcanisphaera  (Acritarcha)  from  the  Tremadocian  of  England. 
Micropaleontology,  22,  479-484. 

— 1979.  Acritarch  zonation  of  the  Tremadoc  series  of  the  Shineton  Shales,  Wrekin,  Shropshire,  England. 
Palynology,  3,  53-72. 

— and  downie,  c.  1974.  The  stratigraphic  distribution  of  Tremadoc  acritarchs  in  the  Shineton  Shales 
succession,  Shropshire,  England.  Rev.  Palaeobot.  Palynol.  18,  1-19. 

rauscher,  r.  1974.  Recherches  Micropaleontologiques  et  Stratigraphiques  dans  l’Ordovicien  et  le  Silurien  en 
France.  Mem.  Sci.  geol.  Univ.  Louis  Pasteur  Strasbourg,  Inst.  Geol.  38  (for  1973),  209  pp. 
reineck,  h.  e.  and  singh,  i.  b.  1972.  Genesis  of  laminated  sand  and  graded  rhythmites  in  storm-sand  layers  of 
shelf  mud.  Sedimentology,  18,  123-128. 

— 1973.  Depositional  sedimentary  environments-.  Springer- Verlag,  Berlin.  439  pp. 

richardson,  j.  b.  and  rasul,  s.  m.  1978.  Palynomorphs  in  Lower  Devonian  sediments  from  the  Apley  Barn 
borehole,  Southern  England.  Pollen,  Spores,  20,  423-462. 

— 1979.  Palynological  evidence  for  the  age  and  provenance  of  the  Lower  Old  Red  Sandstone  from 
the  Apley  Barn  Borehole,  Witney,  Oxfordshire.  Proc.  Geol.  Ass.  90,  27-42. 

staplin,  f.  l.  1961.  Reef-controlled  distribution  of  Devonian  microplankton  in  Alberta.  Palaeontology,  4, 
392-424. 


TURNER:  REWORKED  ACRITARCHS 


143 


staplin,  F.  L.,  jansonius,  J.  and  pocock,  s.  A.  j.  1965.  Evaluation  of  some  acritarchous  hystrichosphere  genera. 
Neues  Jb.  Geol.  Palaont.  Abh.  123,  167-201. 

timofeev,  b.  v.  1958.  Uber  das  Alter  Sachsischer  Grauwacken  Mikropalaophytologische  Untersuchungen  von 
Proben  aus  der  Weesensteiner  und-Lausitzer  Grauwacke.  Geologie,  7,  826-845. 

— 1959.  Drevneyshaya  flora  Prebaltiki;  i ee  stratigraficheskoe  znachenie.  (The  ancient  flora  of  the  Prebaltic 
and  its  stratigraphic  significance.)  Trudy  vses.  neft.  nauchno-issled.  geol.-razv.  Inst.  129.  319  pp.  [In  Russian.] 

— 1966.  Mikropaleofitologicheskoe  isseldovanie  drevnikh  svit.  (Micropalaeophytological  research  into 
ancient  strata).  Akad.  Nauk.  SSSR  Lab.  Geol.  Dokembriya.  147  pp.  [In  Russian.] 

turner,  r.  e.  and  wadge,  a.  j.  1979.  Acritarch  dating  of  Arenig  volcanism  in  the  Lake  District.  Proc.  Yorks,  geol. 
Soc.  42,  405-414. 

vanguestaine,  m.  1978.  Donnees  palynologiques  nouvelles  dans  l’Ordovicien  inferieur  du  bassin  de  la  Senne, 
Massif  du  Brabant,  Belgique.  Annls.  Soc.  geol.  Belg.  100,  193-198. 
vavrdova,  m.  1966.  Palaeozoic  microplankton  from  Central  Bohemia.  Cas.  Miner.  Geol.  11,  409-414. 

— 1972.  Acritarchs  from  Klabava  Shales  (Arenig).  Vest,  ustred.  Ust.  geol.  47,  79-86. 

— 1973.  New  acritarchs  from  Bohemian  Arenig  (Ordovician).  Ibid.  48,  285-289. 

— 1976.  Excystment  mechanisms  of  early  Palaeozoic  acritarchs.  Cas.  Miner.  Geol.  21,  55-64. 

— 1977.  Acritarchs  from  the  Sarka  formation  (Llanvirnian).  Vest,  ustred.  Ust.  geol.  52,  109-118. 
wicander,  E.  R.  1974.  Upper  Devonian -Lower  Mississippian  acritarchs  and  Prasinophycean  algae  from  Ohio, 

U.S.A.  Palaeontographica,  B,  148,  9-43. 

williams,  A.  1969.  Ordovician  of  British  Isles.  In  marshall  kay  (ed.).  North  Atlantic— Geology  and  Continental 
Drift,  pp.  236-264. 

wilson,  l.  R.  1964.  Recycling,  stratigraphic  leakage  and  faulty  techniques  in  palynology.  Grana  Palynol.  5, 
425-436. 


Typescript  received  20  March  1980 
Revised  typescript  received  12  November  1980 


R.  E.  TURNER 

Amoco  Canada  Petroleum  Company  Limited 
Amoco  Canada  Building 
444  Seventh  Avenue  S.W. 

Calgary,  Alberta  T2P  OY2 
Canada 


LOWER  CARBONIFEROUS  CONODONT  FAUNAS 
FROM  RAVENSTONEDALE,  CUMBRIA 

by  a.  c.  higgins  and  w.  j.  varker 


Abstract.  Conodont  faunas  from  the  Lower  Carboniferous,  Courceyan-Holkerian  Stages,  of  Ravenstonedale 
are  described.  Three  zones,  the  Taphrognathus,  Cloghergnathus,  and  Cavusgnathus  Zones,  are  recognized  in 
the  Chadian-Holkerian  strata  and  one  fauna,  Fauna  A,  in  the  mid-Courceyan.  These  zones  and  the  Fauna 
are  correlated  with  the  standard  British  Stages  and  with  the  American  classic  sections.  The  faunas  are  typical 
of  the  shallow  water  facies  of  the  Lower  Carboniferous  and  are  dominated  by  the  genera  Taphrognathus, 
Clydagnathus,  Cloghergnathus,  and  Cavusgnathus,  which  characteristically  occur  in  intertidal  and  shallow 
subtidal  environments.  These  shallow-water  faunas  are  joined  by  deeper-water  immigrants  only  in  the  Arundian 
but,  even  at  this  level,  they  are  the  dominant  elements,  indicating  only  a slight  water  depth  increase.  The 
absence  of  the  deeper-water  genera  Gnathodus  and  Siphonodella  adds  difficulty  to  the  correlation  of  this 
succession  with  goniatite-bearing  sequences,  but  correlation  with  the  Avon  Gorge  and  with  shallow-water 
facies  in  the  north  of  England  is  possible.  Three  species,  Apatognathus  asymetricus,  A.  scandalenis,  and 
Cloghergnathus  carinatus  are  proposed  as  new  taxa. 

This  account  of  the  conodont  faunas  of  the  Lower  Carboniferous  succession  of  Ravenstonedale 
is  a documentation  of  one  aspect  of  a succession  which  in  recent  years  has  become  increasingly 
important  to  the  stratigraphy  of  the  Carboniferous  period  as  a whole.  The  special  significance  of 
Ravenstonedale  is,  however,  long  standing,  since  it  falls  within  the  area  designated  by  Garwood 
(1913,  p.  451)  as  his  type  area  for  the  Lower  Carboniferous  of  the  north  of  England.  Garwood 
(1907,  1913)  had  produced  a zonation,  based  primarily  upon  corals  and  brachiopods,  which  was 
a complement  to  the  slightly  earlier  and  very  influential  work  of  Vaughan  (1905)  in  the  Bristol 
area.  Vaughan  had  designated  the  Lower  Carboniferous  as  ‘Avonian’  in  view  of ‘the  completeness 
of  the  sequence  in  the  Avon  section’  (1905,  p.  264)  and  had  further  produced  the  well-known  and 
much  used  zonal  scheme  based,  he  supposed,  upon  the  evolutionary  lineages  of  corals  and 
brachiopods.  Garwood  (1913)  recognized  the  great  importance  of  Vaughan’s  paper  and  conse- 
quently included  a correlation  of  his  zones  with  those  of  Vaughan  (Garwood  1913,  p.  452). 

This  faunal  zoning  proved  to  be  outstandingly  successful.  Indeed,  in  the  opinion  of  Ramsbottom 
(1973,  p.  568)  the  advance  of  Carboniferous  stratigraphy  was  for  some  time  ironically  hampered 
by  this  very  fact,  since  it  appears  to  have  induced  a period  when  very  little  attention  was  paid  to 
the  many  other  aspects  of  the  Lower  Carboniferous  sequence,  in  particular  to  details  of  lithology 
and  their  interpretation.  Vaughan’s  zonal  scheme  along  with  Garwood’s  correlation  for  the  north 
of  England,  was  the  foundation  upon  which  Dinantian  stratigraphy  was  based  for  sixty  years,  until 
it  was  superceded  by  a series  of  divisions  based  upon  major  cycles  of  deposition  (Ramsbottom 
1973).  These  cycles  coincide  very  closely  with  the  six  new  stages  proposed  by  George,  Johnston, 
Mitchell,  Ramsbottom,  Sevastopulo,  and  Wilson  (1976)  which  are  designed  to  be  applied  throughout 
Britain. 

One  consequence  of  the  re-examination  of  the  Dinantian  rocks  was  that  Ramsbottom  (1973) 
was  able  to  show  that,  far  from  being  complete,  the  Avon  Gorge  succession  contains  four  major 
non-sequences,  and  as  a result  large  parts  of  the  succession  of  the  north  of  England  are  not  able 
to  be  directly  correlated  with  Vaughan’s  type  area  or  zonal  scheme.  In  addition,  other  biozonations 
based  upon  the  Avon  Gorge  succession  must  necessarily  be  as  incomplete  as  the  coral/brachiopod 
zonation.  One  such  scheme  is  that  of  Rhodes,  Austin,  and  Druce  (1969)  in  which  they  established 
fourteen  conodont  assemblage  zones,  the  three  highest  of  which  were  actually  based  upon  the 

(Palaeontology,  Vol.  25,  Part  1,  1982,  pp.  145-166,  pis.  18-19.1 


146 


PALAEONTOLOGY,  VOLUME  25 


north  crop  of  the  South  Wales  coalfield  rather  than  upon  the  Avon  Gorge.  Substantial  revision 
of  this  zonal  scheme  has,  however,  since  (Austin  1973)  been  necessary,  with  the  result  that  whilst 
the  Brigantian  (part  of  three  conodont  assemblage  zones)  and  Courceyan  (six  assemblage  zones 
plus  a non-sequence)  are  usefully  subdivided,  the  whole  of  the  interval  from  just  below  the  top  of 
the  Asbian  down  to  a horizon  within  the  Chadian  is  occupied  by  the  Cavusgnathus/Apatognathus 
assemblage  zone,  non-sequences,  or  horizons  with  no  conodonts  (see  George  et  al.  1976, 
Table  1). 

The  present  study  in  Ravenstonedale  was  initiated  in  an  attempt  to  provide  a Dinantian  conodont 
sequence  from  an  area  in  which  the  succession  has  been  demonstrated  to  be  much  more  complete 
than  that  of  the  Avon  Gorge  (Ramsbottom  1973,  p.  595). 


STRATIGRAPHY 

At  its  outset  the  Carboniferous  period  presented  in  the  North-west  Province  a landscape  of  probably  some 
considerable  relief,  dominated  by  rocks  of  Ordovician  and  Silurian  age.  During  Dinantian  times  this  landscape 
was  gradually  overwhelmed  by  a major  transgression,  evidence  for  which  can  be  seen  in  the  progressive 
overlap  of  Carboniferous  strata  on  to  Lower  Palaeozoic  and  thereby  giving  rise  to  an  increasing  hiatus  at 
the  base  of  the  Carboniferous  succession.  Complete  submergence  was  not  achieved  until  Asbian  times,  and 
as  a result,  within  the  North-west  Province  the  local  base  of  the  Carboniferous  succession  varies  in  age  from 
Courceyan  to  Asbian. 

Deposition  of  the  relatively  thick  (approx.  1500  m)  Dinantian  succession  of  Ravenstonedale  began  with 
local  subsidence  in  a region  known  as  the  Stainmore  Trough  or  Gulf,  which  opened  to  the  east  and  was 
bounded  to  the  north,  west,  and  south  by  the  positive  massifs  of  the  northern  Pennines  and  the  Lake  District. 
Subsidence  appears  to  have  begun  earlier  in  this  trough  than  elsewhere  and  to  have  continued  throughout 
Dinantian  times.  The  Ravenstonedale  area  thus  provides  the  most  complete  Dinantian  succession  within  the 
North-west  Province,  with  what  appears  to  be  an  almost  continuous  record  of  sedimentation  from  Courceyan 
to  Namurian.  Readers  are  referred  to  Johnson  and  Marshall  (1971)  for  further  details  on  the  regional  setting 
of  the  Ravenstonedale  area. 

The  present  outcrop  of  this  succession  strikes  north-westwards  from  the  Dent  Fault  near  Ravenstonedale, 
towards  Penrith  (text-fig.  1)  on  the  southern  side  of  the  Vale  of  Eden.  The  regional  structure  is  relatively 
simple,  with  gentle  dips  predominantly  in  a north  to  north-easterly  direction.  Natural  stream  exposure  is 
generally  good,  and  when  combined  with  other  natural  exposures,  road-cuttings,  and  the  numerous  small 
quarries,  provides  an  almost  complete  record  of  the  succession,  except  near  its  base,  where  much  of  the  detail 
is  obscured  by  an  extensive  deposit  of  glacial  drift.  The  problems  associated  with  the  lower  part  of  the 
succession  were  amply  outlined  by  Holliday,  Neves,  and  Owens  (1979)  in  their  account  of  a shallow  borehole 
programme  undertaken  near  Ravenstonedale  over  the  period  1975  to  1978. 

Working  in  collaboration  with  Holliday  et  al.  the  present  authors  have  recently  described  the  conodont 
faunas  from  the  exposed  part  of  the  Pinskey  Gill  Beds  (Varker  and  Higgins  1979)  which  form  the  lowest  unit 
of  the  succession.  These  beds,  which  lie  unconformably  upon  steeply  inclined  Silurian  Bannisdale  Slates  and 
beneath  the  felspathic  conglomerate,  were  shown  by  Holliday  et  al.  (1976)  to  be  45-50  m in  total  thickness, 
of  which  only  approximately  15  m near  the  base  are  exposed.  The  conodont  faunas  indicate  an  age  of 
mid-Courceyan,  i.e.  late  K or  Z on  the  coral/brachiopod  scheme,  for  the  base  of  the  Dinantian  succession 
in  Ravenstonedale. 

The  overlying  felspathic  conglomerate,  which  is  also  poorly  exposed,  was  calculated  by  Johnson  and 
Marshall  (1971,  p.  267)  to  be  36-4  m in  thickness  in  Thackthwaite  Gill.  However,  since  neither  of  the  boundaries 
of  the  horizon  are  well  exposed  and  Holliday  et  al.  (1976,  p.  349)  have  shown  them  to  be  gradational  with 
no  major  breaks  in  deposition,  calculations  of  thickness  can  only  be  approximate.  Holliday  et  al.  (1976, 
p.  348)  estimated  the  cumulative  thickness  from  their  borehole  evidence  to  be  around  42  m and  interpreted 
the  felspathic  conglomerate  as  a continental  interlude  between  the  marine  Pinskey  Gill  Beds  and  the  overlying 
marine  Stone  Gill  Beds.  Capewell  (1955)  considered  both  the  Pinskey  Gill  Beds  and  the  felspathic  conglomerate 
to  be  the  lateral  equivalents  of  the  much-thicker  basal  beds  which  occur  to  the  north-west,  where  they  reach 
274  m in  thickness  in  the  Lowther  Valley  area. 

Above  the  felspathic  conglomerate  there  are  the  Stone  Gill  Beds,  which  consist  predominantly  of  limestone 
but  also  include  thin  (usually  less  than  1 m)  siltstone  and  shale/mudstone  horizons,  which  are  usually.calcareous. 
The  lower  part  of  this  sequence  and  its  junction  with  the  conglomerate  is  not  exposed  but  Holliday  et  al. 


HIGGINS  AND  VARKER:  CONODONT  FAUNAS 


147 


(1976,  pp.  349,  350)  estimated  the  thickness  of  the  unexposed  beds  to  be  only  slightly  in  excess  of  42  m from 
borehole  evidence.  Virtually  all  of  the  remaining  part  of  the  Stone  Gill  Beds  is  exposed  in  Stone  Gill  itself, 
where  they  have  been  recorded  and  sampled,  at  intervals  not  exceeding  3 m,  by  the  present  authors  (section 
2,  text-figs.  2,  3).  The  sequence  includes  a great  variety  of  limestone  lithologies,  although  calcite  mudstones 
predominate  and  large  parts  of  the  succession  have  suffered  some  dolomitization.  This  part  of  the  Dinantian 
succession  also  includes  several  of  the  marker  bands  described  by  Garwood  (1913),  in  particular  the  Vaughania 
Band  and  the  Palaeechinus  Bed,  both  of  which  occur  near  the  base  of  the  exposed  succession. 


|Permo-Triassic 
| Westphalian 
| j Namurian 
! Dinantian 

; Pre  - Carboniferous 

| ^ | Shop  Granite 


text-fig.  1 . Location  of  the  Ravenstonedale  Area. 


The  overlying  Coldbeck  Beds  begin  at  the  Spongiostroma  Band  of  Turner  (1950)  and  extend  up  to  a nodular 
algal  band  which  outcrops  beneath  Coldbeck  bridge  at  Ravenstonedale  (NY  7209  0435).  These  beds,  which 
are  also  predominantly  fine  grained  limestones,  may  be  distinguished  from  the  Stone  Gill  Beds  by  their  sparser 
fauna  and  the  occurrence  of  several  nodular  algal  layers.  The  algal  layer  described  by  Turner  (1950)  may  be 
one  of  these,  since  it  appears  to  be  slightly  lower  stratigraphically  than  the  band  taken  by  Ramsbottom  ( 1 974, 
p.  52)  to  mark  the  top  of  the  unit.  Exposure  of  the  Coldbeck  Beds  in  Stone  Gill  is  almost  complete  and  once 
again  conodont  samples  were  taken  at  intervals  not  exceeding  three  metres  (section  2,  figs.  2,  3).  It  is,  however, 
likely  that  these  samples  do  not  fully  represent  the  Coldbeck  sequence  since  Turner  (1950,  pp.  30,  35)  presented 
evidence  which  suggested  that  approximately  30-35  m of  beds  had  been  removed  from  this  stream  section 
by  faulting. 

The  Stone  Gill  Beds  and  the  Coldbeck  Beds  were  considered  by  Ramsbottom  (1973,  p.  574)  to  represent 
the  transgressive  and  regressive  phases  of  his  first  major  cycle  of  the  Dinantian,  and  they  were  both  consequently 
included  in  the  Courceyan  Stage  by  George  et  al.  (1976).  An  inconsistancy  was,  however,  noted  by  the  latter 
authors  (1976,  p.  38)  in  that  foraminifera  from  the  Stone  Gill  Beds  included  forms  typical  of  the  Vla 
of  Belgium,  i.e.  Chadian,  not  Courceyan.  Holliday  et  al.  (1979,  p.  354)  considered  the  then  available 


148 


PALAEONTOLOGY,  VOLUME  25 


text-fig.  2.  Geological  map  of  the  Ravenstonedale  area  showing  the  location  of 
the  sampled  sections  (Based  with  permission  upon  one  inch  Geological  Survey 
Sheet  No.  40) 


palaeontological  evidence  for  the  position  of  the  Chadian/Courceyan  boundary  to  be  inconclusive  but 
concluded  on  sedimentological  grounds  that  the  boundary  should  be  lower  than  indicated  by  George  et  al. 
and  placed  it,  with  reservation,  high  in  the  felspathic  (Shap)  conglomerate.  On  this  basis,  the  whole  of  the 
Stone  Gill/Coldbeck  succession  would  be  of  Chadian  age. 

The  succession  continues  with  the  Scandal  Beck  Limestone,  which  was  considered  by  George  et  al.  (1976, 
p.  38)  to  represent  the  whole  of  the  Chadian  Stage.  These  authors  placed  the  top  boundary  of  the  unit,  and 
stage,  in  an  unexposed  part  of  the  sequence  some  1 5 m below  the  Brownber  Pebble  Bed  and  indicated  a total 
thickness  for  the  unit  of  approximately  100  m.  Johnson  and  Marshall  (1971,  p.  265)  considered  the  same 
interval  to  be  closer  to  125  m in  thickness.  The  Scandal  Beck  Limestone  was  sampled  for  conodonts  only 
from  approximately  10  m of  partly  dolomitized  bioclastic  limestone  and  calcite  mudstone  which  occur  near 
the  base  of  the  unit.  It  is  this  sequence  which  outcrops  at  the  confluence  of  Stone  Gill  and  Scandal  Beck 
(section  3,  text-fig.  2),  which  is  represented  in  text-fig.  4. 

The  Brownber  Pebble  Bed,  which  forms  a useful  marker  horizon  in  the  region  as  an  approximation  for 
the  base  of  the  Arundian  Stage,  is  4 to  6 m in  thickness  and  consists  of  calcareous  sandstone  with  quartz 
pebbles,  particularly  in  its  upper  part.  Above  the  pebble  bed  the  Arundian  succession  of  Ravenstonedale  may 
be  conveniently  divided  into  two  parts,  the  Michelinia  grandis  Beds  and  the  overlying  Ashfell  Sandstone. 
There  are  considerable  variations  in  the  published  thicknesses  attributed  to  these  units.  Turner  (1950,  p.  33) 
described  the  M.  grandis  Zone  as  consisting  of ‘nearly  150  feet  (c.  46  m)  of  well  bedded,  somewhat  iron-stained 


HIGGINS  AND  VARKER:  CONODONT  FAUNAS 


149 


SG13  H I 


gap0-30m 


SG4  - P p p 


O 
LU 

OQ  SG23  H 


gapOIOm^ 


s s s 
s s 
s s s 
s s 


STONE  GILL 

.roadbridge 


> 4a  fl~  A~~fpl 


gapl-OOm 


gap3-00m 


OSG32- 

LjJ 

HI 

o SG31- 


EJ  SG16- 
O 


gapl-OOm 


gapOTOm 


gap250m 
— fault 


krkp^n 


text-fig.  3.  Section  in  Stone  Gill  of  the  Stone  Gill  and  Coldbeck  Beds.  For 
legend  see  text-fig.  4. 


limestones  with  a horizon  of  chert  nodules  30  feet  above  the  base’.  Johnson  and  Marshall  (1971,  p.  265), 
however,  allocated  110  m to  the  M.  grandis  Zone  and  approximately  145  m to  the  whole  interval  between 
the  Brownber  Pebble  Bed  and  the  base  of  the  Ashfell  Sandstone.  Ramsbottom  (1974,  p.  55)  indicated  that 
the  beds  of  this  same  interval,  which  represent  the  transgressive  phase  of  his  third  major  cycle,  reach  about 
240  m in  the  Ravenstonedale  district.  Finally,  George  et  al.  (1976,  p.  38,  fig.  11)  allocated  a thickness  of 
about  130  m.  The  ten  conodont  samples  collected  by  the  present  authors  (section  4,  text-fig.  2)  were  collected 
at  regular  intervals  through  the  succession  as  exposed  in  Scandal  Beck. 

The  Ashfell  Sandstone  is  reported  to  reach  170  m in  thickness  in  the  Ravenstonedale  area  (Ramsbottom 
1974,  p.  57)  but  it  does  thin  rapidly  southwards.  Only  the  top  13-25  m have  been  examined  in  the  present 


150 


PALAEONTOLOGY,  VOLUME  25 


study  at  the  well-known  roadside  exposures  of  the  Ravenstonedale/Kirkby  Stephen  road,  Ashfell  Edge  (sec- 
tion 5,  text-fig.  2).  At  this  locality  the  Ashfell  Sandstone  is  represented  by  massive,  false-bedded  sandstone 
with  slump  structures,  soft  purple  and  green  mudstones,  and  thin,  sometimes  lenticular,  limestones,  several 
of  which  are  highly  fossiliferous,  containing  large,  in  situ,  dendritic  coral  colonies  (text-fig.  5).  Turner  (1950, 
p.  34)  considered  these  latter  beds,  the  Lithostrotion  martini  horizon  of  Garwood  (1913)  to  be  confined  to 
the  immediate  vicinity  of  the  Ashfell  Edge  locality. 

The  Arundian/Holkerian  boundary  is  well  exposed  at  this  locality  and  is  marked  by  the  incoming  of  the 
predominantly  limestone  sequence  known  as  the  Ashfell  Limestone,  which  represents  the  whole  of  the  Holkerian 
Stage.  The  lowest  60  m of  this  unit  was  almost  continuously  exposed  in  the  road  section  on  Ashfell  Edge 
where,  along  with  the  upper  part  of  the  Ashfell  Sandstone,  it  was  measured  and  recorded  by  Dr.  I.  C.  Burgess, 
Mr.  M.  Mitchell,  and  Dr.  W.  H.  C.  Ramsbottom  (Institute  of  Geological  Sciences)  in  September  1972.  The 
sequence  represented  in  text-fig.  5 is  based  upon  their  measured  section  (with  their  kind  permission),  but 
unfortunately  recent  civil  engineering  work  has  obscured  some  of  the  higher  horizons.  A total  of  thirteen 
conodont  samples  were  taken  from  the  horizons  indicated  (text-fig.  5).  The  upper  part  of  the  Ashfell  Limestone 
sequence  is  not  continuously  exposed  although  exposures,  both  natural  and  in  quarries,  are  frequent  enough 
to  allow  conodont  samples  to  be  taken  at  fairly  regular  intervals.  The  total  thickness  of  the  Holkerian 
succession  on  Ravenstonedale  Common  approaches  200  m. 


SCANDAL  BECK 


limestone 

dolomitised  limestone 

shale /mudstone 

siltstone 

sandstone 

Pa/aeechinus  bed 

Vaughania  band 

Spongiostroma  band 

algal  horizon 

sample  number 

I.G.S.  bed  no.  for  Ashfell  Edge 


metres 

text-fig.  4.  Section  of  Scandal  Beck  Limestones  seen  in  Scandal  Beck 
and  Legend  for  text-figs.  3-5. 


SB8-\C\E 


SB4 

SB3 


gap1-90m 


p p p p 
V V V V 
S S S S 
A A A A 
SG19 
19 
■3 


HIGGINS  AND  VARKER:  CONODONT  FAUNAS 


15: 


The  Asbian  Stage  is  represented  in  this  region  by  two  main  divisions,  the  Potts  Beck  Limestone  (new  name 
George  et  al.  1976,  p.  40)  and  the  overlying  Knipe  Scar  Limestone.  Only  the  Potts  Beck  Limestone  has  been 
sampled  for  conodonts  in  the  present  study,  from  a series  of  quarries  on  the  north-west  side  of  the 
Ravenstonedale/Kirkby  Stephen  road  (section  6,  text-fig.  2).  According  to  George  et  al.  (1976)  this  limestone 
is  only  found  in  the  Stainmore  Trough  around  Kirkby  Stephen,  in  the  I.G.S.  Borehole  at  Raydale,  and  in 
the  River  Clough  near  Sedbergh.  Elsewhere  there  is  a considerable  non-sequence  between  the  Holkerian  and 
Asbian.  The  Potts  Beck  Limestone  forms  the  youngest  stratigraphic  horizon  in  this  conodont  study  and  was 
the  only  limestone  which  did  not  yield  faunas. 


ASHFELL  EDGE 


X AL4  - 

in 
< 


gap200m 

36 


co 

§ AS3- 

< 

in 


_l  AS2- 

l±J 

Ll  ASI-t 
X 


text-fig.  5.  Section  of  Ashfell  Sandstone  and  Ashfell  Limestone  seen 
in  road  cutting  at  Ashfell  Edge. 


CONODONT  FAUNAS 


52 


PALAEONTOLOGY,  VOLUME  25 


CONODONT  FAUNAS 

A sequence  of  four  conodont  faunas  is  recognizable  in  the  Ravenstonedale  succession.  However, 
because  the  element  ranges  of  the  lowest  fauna  are  incompletely  known  this  is  not  named  as  a 
formal  zone.  The  upper  three  zones  are  from  an  almost  completely  exposed  and  continuous  sequence 
and  are  named  and  defined  with  a greater  degree  of  certainty  (see  text-fig.  6). 

Fauna  A 

The  elements  of  Fauna  A have  recently  been  described  by  Varker  and  Higgins  (1979).  They  occur  in  the 
outcrop  section  of  Pinskey  Gill  and  a series  of  dolomites  and  dolomitic  limestones  in  the  Pinskey  Gill  Borehole 
at  depths  between  33  and  36  m.  The  fauna  probably  typifies  the  whole  of  the  Pinskey  Gill  Beds,  but  is  so 
impoverished  in  both  species  and  abundance  that  the  full  range  of  the  elements  is  not  known. 

The  fauna  is  dominated  by  Bispathodus  aculeatus  aculeatus  (Branson  and  Mehl)  and  Clydagnathus  unicornis 
Rhodes,  Austin  and  Druce  together  with  hindeodellids  and  ozarkodinids.  The  association  of  these  forms  led 
Varker  and  Higgins  to  conclude  that  the  fauna  was  mid-Courceyan  in  age  belonging  to  late  K or  early 
Z Zones  of  the  Avon  Gorge  and  South  Wales,  and  correlating  with  the  costatus  costatus / Gnathodus  delicatus 
Zone  of  Rhodes,  Austin,  and  Druce  (1969).  Sevastopulo  and  Johnston  (personal  communication)  would  place 
the  fauna  firmly  in  the  Z-Zone  in  terms  of  the  Irish  sequence. 


text-fig.  6.  Stratigraphical  ranges  of  conodont  species  in  the  Lower  Carboniferous  of  the  Ravenstonedale 

area. 


HIGGINS  AND  VARKER:  CONODONT  FAUNAS 


53 


Taphrognathus  Zone 

This  is  a partial  range  zone  occurring  in  the  upper  part  of  the  Stone  Gill  Beds  and  in  the  Coldbeck  Beds. 
The  Shap  Conglomerate  sequence  above  the  Pinskey  Gill  Beds  and  the  lowest  part  of  the  Stone  Gill  Beds 
do  not  contain  conodonts.  Faunas  first  appear  some  40  m above  the  base  of  the  exposed  section  in  Stone 
Gill  and  continue  up  to  the  top  of  the  Coldbeck  Beds  at  Coldbeck  Bridge. 

The  fauna  remains  impoverished  in  species,  but  is  somewhat  richer  in  abundance  than  that  of  the  Pinskey 
Gill  Beds.  It  is  characterized  by  Taphrognathus  varians  Branson  and  Mehl,  Cloghergnathus  carinatus  sp.  nov., 
Spathognathodus  scitulus  (Hinde),  Apatognathus  cuspidatus  Varker,  and  a variety  of  non-platform  elements. 
T.  varians  is  probably  the  best-known  species  for  it  has  a very  wide  distribution.  In  the  Upper  Mississippi 
Valley  and  Missouri  (Collinson,  Rexroad,  and  Thompson  1972)  it  characterizes  the  middle  part  of  the 
Valmeyeran  Series  (Upper  Osage/Lower  Meramec)  ranging  from  the  Keokuk  to  the  lower  part  of  the  Upper 
St.  Louis  Formations.  It  is  also  known  from  Australia  (Jenkins  1974)  in  strata  which  Jenkins  concluded  were 
of  early  Visean  age.  Taphrognathus  has  also  been  recorded  from  Britain  in  the  Main  Algal  Limestone  of 
Roxburghshire  (Rhodes  et  al.  1969),  where  it  is  of  early  Visean  age.  Finally  it  occurs  in  Ireland  (Austin  and 
Mitchell  1975)  again  in  the  early  Visean.  Cloghergnathus  also  occurs  with  Taphrognathus  in  Ireland  as  does 
Spathognathodus  scitulus.  S.  scitulus  is  characteristic  of  the  St.  Louis  Formation,  Upper  Valmeyeran  in  the 
Upper  Mississippi  Valley,  where  it  overlaps  the  range  of  Taphrognathus  although  the  closely  related  species 
S.  coalescens  occurs  earlier  and  coexists  with  the  greater  part  of  the  range  of  Taphrognathus  (fig.  7). 

The  Taphrognathus  Zone  correlates  broadly  with  the  Bactrognathus-Taphrognathus  Zone  which  occurs  in 
the  upper  part  of  the  Burlington  Formation  of  the  Upper  Mississippi  Valley.  In  Europe  this  fauna  is  uncommon 
but  does  occur  in  Ireland  (Austin  and  Mitchell  1975)  with  the  characteristic  early  Visean  Mestognathus 
beckmani  and  it  was  given  a lower  Visean  age.  George  et  al.  (1976)  referred  the  Stone  Gill  Beds  and  the 
Coldbeck  Beds  to  the  late  Courceyan,  but  the  presence  of  Vla  foraminifera  (Ramsbottom  1977)  clearly 
indicates  an  early  Visean  age  for  these  sediments,  and  they  are  now  referred  to  the  Chadian  (Ramsbottom  1977). 


UPPER  MISSISSIPPI  VALLEY 

Collinson, Rexroad  * Thompson  1972 

S.  W.  MISSOURI 

Thompson  ♦ Fellows 
1970 

BELGIUM 

BRITAIN 

George  et  al  1976 

RAVENSTONEDALE 

formations 

conodont  zones 

conodont  zones 

assise 

stages 

formations 

conodont  zones 

Apot.  scalenus 

V3b 

Asbian 

St.  Louis 

Cavusgnathus 

V3a 

v2b 

Holkerian 

Ashfell  Limestone 

Cavusgnathus 

V2a 

Ashfell  Sandstone 

— 



Warsaw- Salem 

Taph  varians 
Apatognathus 

vjb 

Arundian 

M.  grandis  Beds 

B 

Cloghergnathus 

Keokuk 

Gnathodus  texanus -Taphrognathus 

Scandal  Beck 
Limestones 

A 

V 

Chadian 

G.  butbosus 

Coldbeck  Beds 

Bactrognathus 

B.  distortus 

Stone  Gill  Beds 

Taphrognathus 

Burlington 

Taphrognathus 

G.  cuneiformis 

Bactrognathus 

Bactrognathus 

Tn3 

Shap 

Fern  Glen 

P.  communis 

Ps.  muttistriatus 

Courceyan 

Conglomerate 

G.  semigtaber 

Meppen 

Ps.  muttistriatus 

/ G.  semigtaber 
/ P comm,  carina 

Pinskey  Gill 

Fauna  A 

S.  isosticha 

\ S.  cooperi  hassi 
\ 6 punctotus 

In2 

Chouteau 

S.  cooperi 

G.  det[catus 
S.  cooperi  cooperi 

text-fig.  7.  Correlation  of  the  Ravenstonedale  conodont  and  sedimentary  sequence  with  the  Belgian  Assises 
and  the  North  American  conodont  sequences. 


154 


PALAEONTOLOGY,  VOLUME  25 


Taphrognathus  sp.  also  occurs  in  the  Lower  Windsor  Group  of  Nova  Scotia  (von  Bitter  1976)  although  its 
occurrence  with  Cavusgnathus  spp.  probably  indicates  a younger  age  than  the  Taphrognathus  Zone  of 
Ravenstonedale. 

Cloghergnathus  Zone 

This  assemblage  zone  occurs  in  the  Scandal  Beck  Limestone  and  the  Michelinia  grandis  Beds.  Every  sample 
has  yielded  a fauna  although  none  of  them  are  abundant. 

The  characteristic  species  present  are  Cloghergnathus  carinatus,  Spathognathodus  scitulus,  Apatognathus 
asymmetricus,  and  A.  scandalensis.  The  appearance  of  Magnilaterella  robust  a and  Neoprioniodus  singular  is  at 
the  base  of  the  Michelinia  grandis  Beds  allows  the  subdivision  of  this  zone  into  two  subzones:  a lower  one 
with  Cloghergnathus,  Apatognathus  asymmetricus,  and  A.  scandalensis,  and  an  upper  one  with  Neoprioniodus 
singularis  and  Magnilaterella  robusta.  Both  of  the  latter  species  are  known  to  have  longer  ranges  elsewhere 
and  their  appearance  within  the  zone  may  well  be  due  to  environmental  change  at  the  base  of  the  Michelinia 
grandis  Beds.  The  fauna  of  the  zone  is  not  distinctive  and  is  merely  an  interregnum  between  the  disappearance 
of  Taphrognathus  and  the  appearance  of  Cavusgnathus. 

This  interval  compares  closely  to  the  Taphrognathus  varians- Apatognathus  interval  in  the  Upper  Mississippi 
Valley  (Collinson  et  al.  1972)  which  occupies  the  Warsaw,  Salem,  and  lower  St.  Louis  Formations.  In  Ireland 
(Austin  and  Mitchell  1975)  there  is  a broad  zone  between  the  top  of  the  Lower  Carboniferous  Shale  (C^) 
with  T.  varians,  and  the  upper  Calp  Limestone  (D,),  with  Cavusgnathus  sp.,  of  which  the  lower  part  would 
correspond  to  the  Cloghergnathus  Zone. 

George  et  al.  (1976)  dated  the  Scandal  Beck  Limestone  (Subzone  A)  as  Chadian  and  the  Michelinia  grandis 
Beds  (Subzone  B)  as  Arundian.  According  to  George  et  al.  (1976)  this  part  of  the  sequence  is  largely  missing 
in  the  Avon  Gorge  or  is  represented  by  non-conodont  bearing  strata  (Austin  1973). 

Cavusgnathus  Zone 

The  Cavusgnathus  Zone  includes  the  highest  beds  of  the  Ashfell  Sandstone  and  at  least  the  lowest  beds  of  the 
Ashfell  Limestone.  The  fauna  includes  Cavusgnathus  regularis  and  C.  unicornis,  Neoprioniodus  scitulus, 
Spathognathodus  scitulus,  Magnilaterella  robusta,  and  Apatognathus  libratus. 

This  fauna  correlates  with  that  of  the  Apatognathus  scalenus-Cavusgnathus  Zone  of  the  Upper  Mississippi 
Valley  (Collinson  et  al.  1972)  which  occurs  in  the  Upper  St.  Louis  Formation.  However,  as  Austin  (in  Austin 
and  Mitchell  1975)  has  pointed  out,  there  is  a breccia  between  the  Upper  and  Lower  St.  Louis  Formation 
with  some  condensation  of  faunas.  This  may  account  for  the  absence  of  the  major  gap  between  the  disappearance 
of  Taphrognathus  and  the  appearance  of  Cavusgnathus  which  occurs  in  the  Ravenstonedale  sequence.  This 
zone  compares  well  with  the  lower  part  of  the  Cavusgnathus- Apatognathus  Zone  of  the  Avon  Gorge  (Austin 
1973)  (text-fig.  8). 

The  Ashfell  Sandstone  is  included  in  the  upper  part  of  the  Arundian  and  the  Ashfell  Limestone  in  the 
Holkerian  by  George  et  al.  (1976). 


Influence  of  environment  on  the  conodont  faunas 

The  influence  of  the  depositional  environment  on  the  conodont  animal  in  the  Palaeozoic  generally  is 
readily  apparent  from  a perusal  of  the  symposium  volume  on  conodont  palaeoecology  (Barnes 
1976),  and  in  the  Lower  Carboniferous  in  particular  from  the  work  of  Austin  (1976)  and  von  Bitter 
(1976).  There  is  a clear  distinction  between  extremely  shallow-water  faunas  seen  in  the  Windsor 
Group  of  Nova  Scotia  (Globensky  1967;  von  Bitter  1976)  and  Ravenstonedale  and  the  basinal 
faunas  of  the  same  age  in  Spain  (Higgins  and  Wagner  Gentis,  in  press)  and  in  Germany  (Bischoff 
1957;  Voges  1959).  Other  faunas  show  mixing  of  both  shallow-water  and  basinal  types,  and  Austin 
(1976)  has  attempted  to  show  typical  associations  of  this  type.  Although  these  differences  are 
normally  ascribed  to  a depth  control  on  the  conodont  animal,  there  may  be  also  an  element  of 
provincialism  as  suggested  by  Austin  (1976),  for  the  type  of  fauna  seen  in  Ravenstonedale  appears 
to  have  a restricted  geographical  distribution.  In  view  of  the  growing  evidence  of  environmental 
control  of  the  conodont  animal  it  is  important  to  know  whether  the  faunal  changes  in  the 
Ravenstonedale  sequence  reflect  evolutionary  or  environmental  changes,  or  both. 

Clydagnathus,  occurring  in  dolomites  and  dolomitic  limestones  in  Pinskey  Gill  is,  according  to 
Austin  (1976),  found  in  dolomites  and  oolites  in  littoral/deltaic  and  offshore  neritic  facies  only. 


HIGGINS  AND  VARKER:  CONODONT  FAUNAS 


.55 


STAGES 

George  et  al  1976 

AVON 

GORGE 

RAVENSTONEDALE 
conodont  zones 
this  paper 

George  et  al  1976 
Kellaway  8 Welch  1955 

Rhodes,  Austin  8 Druce  1969 
Austin  1973 

Holkerian 

Upper  Clifton 
Down  Limestone 

Cavusgnathus 

Apatognathus 

Cavusanathus 

Arundian 

Lower  Clifton  Down  Limestone 
Upper  Clifton  Down  Mudstone 
Goblin  Coombe  Oolite 

Cavusgnathus  - Apatognathus 
no  conodonts 

B 

Cloghergnathus 

Lower  Clifton  Down  Mudstone 

no  conodonts 

Chadion 

Gully  Oolite 
Sub -Oolite  Bed 

Mestognathus  beckmanni 
Polygnathus  bischoffi 

A 

Taphrognathus 

\\VV\VV 

\\ V\\V\VVV^ 

G.  an  tet  exanus 
Ptacinatus 

Ptacinatus  - Ps.tongiposticus 

Courceyon 

Black  Rock  Limestone 

Bispathodus  costatus  costatus 
G.  del i cat  us 

Fauna  A 

Lower  Limestone  Shale 

poor  exposure 
S.  cf.  robustus 
B.  acuteatus 

Siphonodella  - Pinornatus 

Shirehampton  Beds 

Pa.  variabitis  - Pinornatus 

text-fig.  8.  Comparison  of  the  conodont  zonation  of  Ravenstonedale  with  that  of  the  Avon  Gorge. 


and  is  typically  supratidal.  Sandberg  (1976),  in  his  study  of  late  Devonian  biofacies,  concluded 
that  Clydagnathus  is  abundant  in  offshore  banks  and  lagoons  of  shallow  brackish  to  normally 
saline  banks  commonly  occurring  with  algae. 

Taphrognathus,  according  to  Austin  has  a similar  distribution  to  that  of  Clydagnathus , and  von 
Bitter  (1976)  came  to  a similar  conclusion  in  relating  Taphrognathus  to  his  Biofacies  II  which 
occurred  in  inner  shelf  and  reefoid  environments. 

Apatognathus  and  Spathognathodus  scitulus  generally  occur  together  and  were  suggested  by 
Austin  (1976)  to  be  an  association  which  occurred  in  the  littoral/delta  front,  offshore  neritic,  and 
back  reef  lagoonal  facies.  Von  Bitter  grouped  these  two  with  his  Biofacies  II  association. 

Cavusgnathus  has  a similar  shallow-water  origin  to  Taphrognathus  according  to  Austin  (1976), 
von  Bitter  (1972,  1976),  and  Merrill  and  Martin  (1976). 

Cloghergnathus  is  a relatively  new  genus  and  its  distribution  is  poorly  known,  but  it  does  occur 


156 


PALAEONTOLOGY,  VOLUME  25 


in  the  Windsor  Group  of  Nova  Scotia  (Globensky  1967;  von  Bitter  1976)  where  it  would  probably 
be  placed  in  von  Bitter’s  biofacies  II,  and  it  also  occurs  in  Ireland  (Austin  and  Mitchell  1975) 
where  it  occurs  in  association  with  Taphrognathus. 

The  majority  of  the  platform  conodonts  in  these  faunas  are  asymmetric:  Clydagnathus  unicornis 
and  all  species  of  Cavusgnathus  are  right-sided  whereas  Cloghergnathus  carinatus  is  both  left  and 
right  sided  and  Taphrognathus  varians  often  shows  a tendency  towards  asymmetry.  The  asymmetry 
appears  to  be  associated  with  shallow-water  environments,  but  is  not  restricted  to  these  environ- 
ments. Cavusgnathus , for  example,  may  occur  in  basinal  environments  (Higgins  1975),  and  so  may 
Apatognathus  but  in  small  numbers  only.  They  are  only  dominant  in  shallow-water  faunas.  Similarly, 
the  basinal  faunas  of  this  age,  dominated  by  the  symmetrical  or  highly  ornamented  platform 
elements,  may  occur  on  the  margins  of  the  shallow-water  zones  mixed  with  asymmetric  elements. 
Von  Bitter  recorded  the  presence  of  simple  gnathodids  in  his  biofacies  II,  and  more  complex 
gnathodids  occur  with  cavusgnathids  and  apatognathids  in  his  deep-water  biofacies  III.  In  the 
Ravenstonedale  sequence  such  mixing  occurs  rarely,  indicating  the  extreme  shallowness  of  the 
water.  Neoprioniodus  singularis,  which  is  a common  basinal  species,  occurs  as  an  uncommon 
component  of  the  Michelinia  grandis  Beds  but  its  origin  is  known  to  be  earlier  in  basinal  sequences 
and  it  must  be  regarded  as  a deeper-water  immigrant  into  a dominantly  shallow-water  environ- 
ment. 

Thus  Clydagnathus,  Cloghergnathus,  Taphrognathus,  Cavusgnathus,  and  Apatognathus  are 
environmentally  controlled  to  a high  degree,  occurring  typically  in  littoral  and  lagoonal  facies 
but  ranging  into  offshore  infratidal  facies.  The  faunas  in  which  they  occur  are  restricted  in  both 
variety  and  abundance.  Their  persistent  occurrence  throughout  the  Ravenstonedale  sequence 
is  a reflection  of  the  persistence  of  this  facies  in  the  area,  and  has  enabled  us  to  study  the 
development  of  these  unusual  faunas  through  a long  sequence.  The  proposed  zones,  although 
fundamentally  environmentally  controlled,  are  of  biostratigraphic  usage  within  this  type  of 
environment. 

Neoprioniodus  singularis,  on  the  other  hand,  is  clearly  of  only  local  value  as  a stratigraphic 
marker  and  merely  reflects  the  major  transgression  which  occurred  at  the  base  of  the  Michelinia 
grandis  Beds  (see  also  Leeder  in  discussion  of  George  1978)  and  brought  with  it  immigrant  faunas 
into  the  area. 

In  broad  outline  the  changes  in  the  conodont  faunas  occur  both  at  stage  boundaries  and  at  the 
cycle  boundaries  of  Ramsbottom  (1973,  1977),  although  the  two  do  not  exactly  coincide.  The 
boundary  between  the  Taphrognathus -Cloghergnathus  Zones  is  coincident  with  the  algal  horizon 
at  the  top  of  the  Coldbeck  Beds  which,  according  to  Ramsbottom  (1973,  1977),  is  the  boundary 
between  two  cycles.  Similarly  the  boundary  between  Subzones  A and  B of  the  Cloghergnathus 
Zone  coincides  with  the  Chadian-Arundian  boundary  although  the  exact  horizon  of  the  boundary 
is  thought  to  be  unexposed  (George  et  al.  1976).  The  beginning  of  the  Cavusgnathus  Zone  coincides 
with  the  appearance  of  the  first  limestone  above  the  sandstone  beds  of  the  Ashfell  Sandstone. 
These  occur  in  a few  metres  of  shales  immediately  below  the  base  of  the  Ashfell  Limestone  which 
is  taken  as  the  boundary  between  the  Arundian  and  Holkerian  Stages,  and  the  boundary  between 
cycles  3 and  4 of  Ramsbottom  (1973). 

With  the  exception  of  the  Subzone  B fauna,  all  the  conodont  faunas  are  of  a shallow-water  type 
and  they  occur  throughout  the  cycles  in  the  Chadian,  through  the  lower  part  of  the  Arundian,  and 
reappear  at  the  top  of  the  Arundian.  The  general  distribution  of  the  faunas  supports  the  attribution 
of  the  beds  to  the  stages  of  George  et  al.  (1976),  but  the  similarity  between  the  faunas  is  probably 
due  to  the  position  of  the  Ravenstonedale  area  during  Lower  Carboniferous  times  at  the  head  of 
a shallow  gulf  where  a change  of  sea  level  to  bring  in  sandstone  or  algal  horizons  led  to  the  absence 
of  conodonts  rather  than  a change  in  the  type  of  fauna. 


HIGGINS  AND  VARKER:  CONODONT  FAUNAS 


.57 


SYSTEMATIC  DESCRIPTIONS 

All  type  and  figured  material  has  been  lodged  in  the  micropalaeontological  collection.  Department 
of  Geology,  University  of  Sheffield. 

Genus  Apatognathus  Branson  and  Mehl 
Type  species.  Apatognathus  varians  Branson  and  Mehl,  1934 


Apatognathus  asymmetricus  sp.  nov. 

Plate  19,  figs.  7,  9 

Holotype.  PI.  19,  fig.  7.  R21. 

Horizon.  Scandal  Beck  Limestone,  sample  SB9. 

Diagnosis.  Anterior  process,  laterally  thickened,  half  the  height  of  the  posterior  process  which  is 
thin.  Denticles  on  posterior  process  fused  and  twice  the  height  of  those  on  the  anterior  process. 

Description.  Processes  unequal,  diverging  at  about  40°.  Strongly  inwardly  curved.  Anterior  process  low,  laterally 
thickened  with  a shelf  on  the  upper  edge,  and  maintaining  its  height  to  the  extremity.  Up  to  ten  denticles  on 
the  oral  surface  whose  inclination  increases  towards  the  cusp.  The  denticles  are  small,  laterally  compressed, 
and  are  incurved.  They  are  discrete  for  about  half  their  length. 

Posterior  process  thin,  but  high,  up  to  twice  the  height  of  the  anterior  process  and  the  inner  face  is  strongly 
convex.  The  denticles  are  twice  as  high  as  those  of  the  anterior  process  and  are  fused  for  most  of  their  length. 
They  are  strongly  laterally  compressed  and  decrease  in  height  towards  the  extremity  of  the  process.  The 
denticles  are  slightly  incurved. 

Cusp  is  twice  the  height  and  width  of  the  posterior  bar  denticles.  The  adjacent  posterior  bar  denticle  is 
fused  to  its  posterior  edge,  but  it  is  largely  free  on  its  anterior  edge  due  to  the  small  size  of  the  anterior 
process  denticles. 

The  cavity  is  triangular  and  is  situated  on  the  inner  face  of  unit  below  the  cusp.  The  processes  are  grooved 
on  the  aboral  side. 

Comparison.  This  species  most  closely  resembles  Apatognathus  cuspidatus  from  which  it  differs  in 
lacking  lateral  thickening  on  the  posterior  process  and  having  more  strongly  fused  denticles.  It 
differs  from  A.  scandalensis  in  having  lateral  thickening  on  the  anterior  process. 


Apatognathus  libratus  Varker 
Plate  19,  fig.  12 

1967  Apatognathus?  libratus  Varker  pp.  134,  135,  pi.  18  figs.  3,  6,  8,  9,  12,  13. 

Remarks.  This  is  an  extremely  large  specimen,  but  its  major  characteristics  conform  to  the  diagnosis 
of  Apatognathus  libratus.  The  main  difference  lies  in  the  width  of  the  process  beneath  the  cusp 
which  is  twice  as  broad  as  in  other  figured  specimens  of  this  species. 

Apatognathus  scandalensis  sp.  nov. 

Plate  19,  fig.  10 

Holotype.  PI.  19,  fig.  10,  R18. 

Horizon.  Scandal  Beck  Limestone,  sample  SB8. 

Diagnosis.  A wide  angled,  almost  symmetric,  Apatognathus  with  thin  processes  and  denticles  on 
the  anterior  process  which  are  highest  near,  but  not  at,  the  cusp.  The  denticles  on  the  anterior 
process  are  twice  the  height  of  those  on  the  posterior  process. 


158 


PALAEONTOLOGY,  VOLUME  25 


Description.  The  processes  diverge  at  about  50°  and  are  strongly  inwardly  curved.  The  unit  is  slender  and 
thickening  of  the  processes  only  occurs  close  to  the  cusp,  where  a slight  shelf  is  formed. 

The  anterior  limb  is  slightly  longer  than  the  posterior  and  is  strongly  curved  inwardly.  It  is  higher  than 
the  posterior  limb  and  bears  up  to  thirteen  laterally  compressed  denticles  which  rise  to  a point  about  one-third 
from  the  cusp.  The  denticles  are  sharply  pointed  and  are  flatter  on  the  outer  than  the  inner  side  and  are 
discrete  for  more  than  half  their  length.  The  posterior  process  is  about  two-thirds  the  length  of  the  anterior 
process  and  is  inwardly  curved  about  the  same  amount.  Only  slight  thickening  occurs  near  the  cusp.  Up  to 
ten  denticles  occur  on  its  oral  surface  and  these  are  only  half  the  height  of  the  largest  anterior  process  denticles 
and  are  subequal. 

The  cusp  is  twice  the  height  and  width  of  the  largest  denticle  on  the  processes,  and  is  laterally  compressed 
more  strongly  on  the  outer  than  the  inner  side.  It  is  discrete  for  more  than  half  its  length  and  is  sharply 
pointed.  It  is  slightly  curved  inwards. 

A shallow  triangular  pit  occurs  at  the  base  of  the  cusp  and  is  situated  on  the  inner  face  rather  than  aborally. 
It  is  continued  along  the  processes  as  barely  discernible  grooves. 

Comparison.  Apatognathus  scandalensis  most  closely  resembles  A.  cuspidatus  but  differs  in  having 
a less  strongly  inclined  and  a smaller  cusp  and  a larger  angle  between  the  processes. 

Genus  Cavusgnathus  Harris  and  Hollingworth,  1933 
Type  species.  Cavusgnathus  altus  Harris  and  Hollingworth,  1933 

Cavusgnathus  regularis  Youngquist  and  Miller 
Plate  18,  figs.  12,  13 

1949  Cavusgnathus  regularis  Youngquist  and  Miller,  p.  169,  pi.  101,  figs.  24,  25. 

Remarks.  The  generally  short  and  compact  form  and  the  regular  denticulation  of  the  blade  of  this 
species  are  present  in  the  Ravenstonedale  specimens,  but  there  are  some  differences  from  previously 
described  examples.  The  principal  difference  is  the  length  of  the  fixed  blade  which  extends  up  to 
half  the  length  of  the  platform  and  is  larger  than  that  normally  found  on  specimens  of  this  species. 


EXPLANATION  OF  PLATE  18 


All  figures  x 60 

Figs.  1-11.  Cloghergnathus  carinatus.  P.  element.  1,  oral  view  of  specimen  from  the  Stone  Gill  Beds,  sample 
SGI 9,  specimen  R29,  transitional  to  Taphrognathus  varians.  2,  7,  oral  and  outer  lateral  views  of  holotype, 
specimen  from  the  Scandal  Beck  Limestone,  sample  SB2,  specimen  R30.  3,  inner  lateral  view  of  a specimen 
from  the  Stone  Gill  Beds,  sample  SG19,  specimen  R32.  4,  oral  view  of  a young  specimen  from  the  Coldbeck 
Beds,  sample  SG32,  specimen  R31.  5,  outer  lateral  view  of  a young  specimen  from  the  Michelinia  grandis 
Beds,  sample  MB10,  specimen  R35.  6,  10,  outer  lateral  and  oral  views  of  a specimen  from  the  Scandal 
Beck  Limestone,  sample  SB9,  specimen  R33.  8,  9,  11,  aboral,  outer  lateral  and  oral  views  of  a specimen 
from  the  Coldbeck  Beds,  sample  SG32,  specimen  R34,  specimen  transitional  to  Taphrognathus  varians. 

Figs.  12,  13.  Cavusgnathus  regularis.  Oral  and  inner  lateral  views  of  a specimen  from  the  Ashfell  Sandstone, 
sample  AS1,  specimen  R36. 

Fig.  14.  Cavusgnathus  unicornis.  Inner  lateral  view  of  a specimen  from  the  Ashfell  Sandstone,  sample  AS1, 
specimen  R38. 

Figs.  15,  16.  Taphrognathus  varians.  Oral  and  inner  lateral  views  of  specimens  from  the  Stone  Gill  Beds, 
sample  SGI 9,  specimens  R37,  R39. 

Fig.  17.  Lonchodina  sp.  Inner  lateral  view  of  specimens  from  the  Coldbeck  Beds,  sample  SG33,  specimen  R27. 

Fig.  18.  O element  of  Cloghergnathus  carinatus.  Inner  lateral  view  of  a specimen  from  the  Coldbeck  Beds, 
sample  SG32,  specimen  R14. 

Fig.  19.  A,  element  of  Cloghergnathus  carinatus.  Inner  lateral  view  of  a specimen  from  the  Coldbeck  Beds, 
sample  SG31,  specimen  R3. 


PLATE  18 


HIGGINS  and  VARKAR,  Lower  Carboniferous  conodonts 


160 


PALAEONTOLOGY,  VOLUME  25 


Cavusgnathus  unicornis  Youngquist  and  Miller 
Plate  18,  fig.  14 

1949  Cavusgnathus  unicornis  Youngquist  and  Miller,  p.  619,  pi.  101,  figs.  18-23. 

Remarks.  The  median  carina  developed  only  in  the  posterior  part  of  the  platform  is  prominent 
only  in  large  specimens  and  is  often  completely  absent  in  small  specimens. 

Genus  Cloghergnathus  Austin  and  Mitchell 
Type  species.  Cloghergnathus  globenskii  Austin  and  Mitchell,  1975. 

Remarks.  Cloghergnathus  was  named  by  Austin  (in  Austin  and  Mitchell  1975)  for  specimens  with 
a lateral  blade  which  does  not  extend  on  to  the  platform  as  a high,  crested  structure  as  in 
Cavusgnathus.  The  blade  may  occupy  a sub-median  position  whilst  remaining  connected  to  one  of 
the  platform  sides  and  such  forms  are  probably  transitional  to  Taphrognathus.  Its  characteristics 
are  those  of  Clydagnathus  Rhodes,  Austin,  and  Druce  an  early  Courceyan  genus  but  a considerable 
stratigraphic  gap  occurs  between  the  last  occurrence  of  Clydagnathus  and  the  first  appearance  of 
Cloghergnathus  in  the  Chadian. 


Cloghergnathus  carinatus  sp.  nov. 

P element 
Plate  18,  figs.  1-11 

Holotype.  PI.  18,  figs.  2,  7.  R30. 

Horizon.  Scandal  Beck  Limestone,  sample  SB  2. 

Diagnosis.  A right-  and  left-sided  element  with  the  anterior  blade  developed  on  the  inner  side  of 
the  unit.  A central  trough  runs  the  length  of  the  platform  but  is  occupied  by  a short  carina  in  the 
posterior  quarter.  The  blade  is  short,  one-quarter  to  one-third  the  length  of  the  platform,  and  does 
not  extend  on  to  the  platform.  It  is  extended  above  the  platform  and  is  convexly  curved.  The  unit 
is  arched. 

Description.  The  anterior  blade  consists  of  up  to  five  denticles  of  which  the  largest  is  in  the  centre  giving  the 
blade  a convex  outline.  The  denticles  are  free  for  about  half  their  length.  The  platform  is  straight  to  strongly 
curved  and  bears  a deep  median  trough  which  is  occupied  by  a carina  for  up  to  the  last  third  of  the  platform. 
The  carina  is  very  prominent  in  small  specimens.  The  inner  parapet  is  curved  and  commonly  does  not  reach 
the  posterior  end  of  the  unit  and  is  continued  anteriorly  as  the  blade,  although  the  curvature  of  some  specimens 
is  difficult  to  determine  because  the  blade  may  be  outwardly  curved.  In  a few  specimens  the  blade  is  slightly 
offset  and,  whilst  it  originates  from  the  inner  side  of  the  platform,  it  occupies  a sub-central  position  on  both 
right  and  left  forms.  The  oral  surface  of  the  platform  is  covered  by  nodes  or  transverse  ridges.  The  platform 
is  widest  near  midlength  and  commonly  the  inner  side  is  wider  than  the  outer.  The  unit  is  arched  in  lateral 
view  both  the  oral  and  aboral  surfaces  of  the  platform  being  curved  and  the  aboral  surface  of  the  blade  being 
strongly  downturned.  The  curvature  is  less  marked  in  young  specimens  which  may  be  almost  straight.  The 
cavity  occupies  the  whole  of  the  aboral  surface  of  the  platform  and  is  widest  slightly  anterior  to  its  mid-point 
tapering  to  both  the  anterior  and  the  posterior. 

Comparisons.  Cloghergnathus  carinatus  differs  from  C.  globenskii  Austin  and  Mitchell  by  its  arched 
aboral  and  oral  surfaces,  its  convex  blade,  which  originates  from  the  inner  side,  and  the  presence 
of  the  posterior  carina.  C.  rhodesi  is  left-sided,  does  not  possess  a carina,  and  its  blade  shape  is 
unknown. 

Remarks.  There  is  a tendency  for  the  blade  of  C.  carinatus  to  be  median  in  position  and  it  may 
include  some  of  the  Taphrognathus-Cavusgnathus  transition  forms  of  Rexroad  and  Collinson  (1963; 
especially  fig.  24,  p.  111).  Austin  (in  Austin  and  Mitchell  1975)  has  argued  that  Cloghergnathus  is 


HIGGINS  AND  VARKER:  CONODONT  FAUNAS 


61 


not  intermediate  between  Taphrognathus  and  Cavusgnathus  but  is  an  offshoot  of  the  former  genus 
in  which  the  blade,  although  lateral,  does  not  extend  on  to  the  platform.  In  this  respect 
Cloghergnathus  mirrors  Adetognathus  which  is  probably  a younger  homeomorph,  or  which  may 
prove  to  be  the  same  genus  when  its  full  range  is  known.  All  figured  specimens  of  Cloghergnathus 
have  a low  blade  or  one  which  rises  gradually  as  in  C.  carinatus  none  of  them  have  the  typical 
cavusgnathoid  blade.  In  addition,  all  known  species  of  Cloghergnathus  are  either  right-  and  left-sided 
or  only  left-sided,  whereas  Cavusgnathus  species  are  all  right-sided.  C.  carinatus  is  also  similar  to 
Clydagnathus  darensis  but  differs  in  having  a larger  basal  cavity. 

Cloghergnathus  non-platform  Elements 
O element 

Plate  18,  fig.  18;  Plate  19,  fig.  4 

1957  Ozarkodina  compressa  Rexroad,  p.  36,  pi.  2,  figs.  1,  2. 

Remarks.  The  Ravenstonedale  specimens  have  the  prominent  apical  denticle  of  the  paratype  but 
have  a smaller  number  of  denticles  on  the  processes  than  is  usual  in  this  element.  However,  it  is 
thought  to  fall  within  the  variability  of  the  element. 

N element 
Plate  19,  fig.  18 

1941  Prioniodus  varians  Branson  and  Mehl,  p.  174,  pi.  5,  figs.  7,  8. 

1957  Prioniodina  varians  (Branson  and  Mehl),  Bischoff,  p.  49,  pi.  5,  fig.  35. 

1957  Neoprioniodus  varians  (Branson  and  Mehl),  Rexroad,  p.  35,  pi.  2,  fig.  10. 

Remarks.  Von  Bitter  (1976)  named  Neoprioniodus  camurus  Rexroad  as  the  Ne  element  in  his 
multi-element  reconstruction  of  Cavusgnathus  windsorensis.  He  also  figured  a similar  form  in  his 
reconstruction  of  the  non-P  elements  of  Cavusgnathus  sp.  from  the  Pennsylvanian  (1972,  pi.  9, 
fig.  5).  Baesemann  (1973)  figured  a similar  neoprioniodontid  in  his  reconstruction  of  Adetognathus 
(pi.  2,  figs.  26,  35).  No  form  comparable  to  N.  camurus  occurs  in  the  Ravenstonedale  samples 
where  the  only  neoprioniodontids  are  N.  varians , N.  acampylus,  and  N.  scitu/us. 

A:  element 

Plate  18,  fig.  19;  Plate  19,  fig.  20 

1953  Hindeodella  ensis  Hass,  p.  81,  pi.  16,  figs.  19-21. 

1960  Hindeodella  tenuis  Clarke,  p.  8,  pi.  1,  figs.  10,  11. 

Remarks.  There  are  two  variants  in  this  element.  One  is  typified  by  Hindeodella  tenuis  and  H.  ensis 
and  is  the  more  common  of  the  two.  This  has  a long  posterior  process  with  alternating  denticles 
and  a long,  often  inwardly  curved  anterior  bar  commonly  with  non-alternating  denticles  which 
are  as  large  as  the  large  posterior  bar  denticles.  The  other  variant  is  shorter  and  arched  with  a 
cusp  which  is  as  long  as  the  posterior  process,  and  has  non-alternating  denticles  of  which  the 
posterior  process  denticles  increase  in  size  posteriorly. 

A 3 element 
Plate  19,  figs.  5,  6,  8 

Description.  A robust  unit  consisting  of  two  massive  lateral  processes,  a large  cusp,  and  a slender  posterior 
process.  The  cusp  is  large,  slightly  curved  posteriorly,  and  has  a triangular  cross-section  with  sharp  posterior 
and  lateral  edges.  Its  surface  is  covered  with  fine  discontinuous  striations.  The  posterior  bar  is  incomplete, 
appears  to  be  narrow,  and  has  an  oval  cross-section. 

The  lateral  processes  are  massive,  bar-like  with  a square  cross-section,  and  almost  in  the  same  plane.  They 


162 


PALAEONTOLOGY,  VOLUME  25 


are  as  long  as  the  main  cusp  and  slightly  inclined  downwards.  The  denticles  on  the  oral  surface,  up  to  five 
in  number,  are  large,  increasing  in  size  towards  the  extremities  of  the  processes,  and  are  flattened  in  an 
antero-posterior  direction.  The  largest  denticle  is  approximately  half  the  size  of  the  cusp. 

The  aboral  side  of  the  unit  is  grooved,  and  the  grooves  meet  forming  a small  triangular  pit  beneath  the  cusp. 

Remarks.  This  form  most  closely  resembles  Hibbardella  ortha  Rexroad,  the  primary  difference 
being  the  shallow  angle  between  the  processes.  It  is  also  much  more  massive  than  H.  ortha  but 
this  is  a characteristic  shared  by  specimens  of  all  species  in  the  Ravenstonedale  faunas  and  is  most 
likely  due  to  the  shallow-water  environments  in  which  the  fauna  occurs. 

Genus  Magnilaterella  Rexroad  and  Collinson 
Type  species.  Magnilaterella  robust  a Rexroad  and  Collinson  1963. 

Magnilaterella  robusta  Rexroad  and  Collinson 

1940  Lonchodina  sp.  Branson  and  Mehl,  p.  171,  pi.  5,  fig.  10  only. 

1956  Metalonchodinal  sp.  Elias  p.  124,  pi.  4,  fig.  3. 

1963  Magnilaterella  robusta  Rexroad  and  Collinson,  pp.  14-17,  pi.  1,  figs.  4,  5,  9. 

Remarks.  Typically  Magnilaterella  robusta  has  a prominent  bevel  on  the  inner  side  of  the  lateral 
process.  However,  Rexroad  and  Collinson  (1965)  did  include  specimens  without  the  bevel  which 
have  the  characteristic  denticulation  on  the  lateral  process. 

At  the  present  time  the  multi-element  species  to  which  M.  robusta  belongs  is  unknown. 


EXPLANATION  OF  PLATE  19 


All  figures  x 60 

Figs.  1-3.  Lonchodina  sp.  Inner  lateral  views  of  specimens  from  the  Coldbeck  Beds.  Fig.  1 from  sample  SG33, 
specimen  Rl;  fig.  2 from  sample  SG31,  specimen  R2;  fig.  3 from  sample  SG33,  specimen  R40. 

Fig.  4.  O element  of  Cloghergnathus  carinatus.  Inner  lateral  view  of  a specimen  from  the  Coldbeck  Beds, 
sample  29,  specimen  R26. 

Figs.  5,  6,  8.  A 3 element  of  Cloghergnathus  carinatus.  Posterior  views  of  specimens  from  the  Coldbeck  Beds, 
sample  29,  specimens  R9-11. 

Figs.  7,  9.  Apatognathus  asymmetricus.  7,  inner  lateral  view  of  holotype  from  the  Scandal  Beck  Limestone, 
sample  SB9,  specimen  R21.  9,  inner  lateral  view  of  a specimen  from  the  Scandal  Beck  Limestone,  sample 
SB7,  specimen  R20. 

Fig.  10.  Apatognathus  scandalensis.  Inner  lateral  view  of  holotype  from  the  Scandal  Beck  Limestone,  sample 
SB8,  specimen  R18. 

Figs.  11,  13,  19.  Apatognathus  cuspidatus.  Inner  lateral  views  of  specimens  from  the  Scandal  Beck  Limestone, 
fig.  11  (SB7),  specimen  R16  and  fig.  19  (SB2),  specimen  R17,  and  Coldbeck  Beds  sample  SG33,  specimen 

R15. 

Fig.  12.  Apatognathus  libratus.  Inner  lateral  view  of  a specimen  from  the  Ashfell  Sandstone,  sample  AS1, 
specimen  R19. 

Fig.  14.  Spathognathodus  scitulus.  Outer  lateral  view  of  a specimen  from  the  Ashfell  Sandstone,  sample  AS1, 
specimen  R5. 

Fig.  15.  Neoprioniodus  singularis.  Inner  lateral  view  of  a specimen  from  the  Michelinia  grandis  Beds,  sample 
MB2,  specimen  R8. 

Fig.  16.  Neoprioniodus  cf.  acampylus.  Inner  lateral  view  of  a specimen  from  the  Michelinia  grandis  Beds, 
sample  MB2,  specimen  R13. 

Fig.  17.  Neoprioniodus  sp.  Inner  lateral  view  of  a specimen  from  the  Coldbeck  Beds,  sample  SG29,  speci- 
men R7. 

Fig.  18.  N element  of  Cloghergnathus  carinatus.  Inner  lateral  view  of  a specimen  from  the  Scandal  Beck 
Limestone,  sample  SB9,  specimen  R12. 

Fig.  20.  A,  element  of  Cloghergnathus  carinatus.  Inner  lateral  view  of  a specimen  from  the  Coldbeck  Beds, 
sample  SG29,  specimen  R24. 


PLATE  19 


HIGGINS  and  VARKAR,  Lower  Carboniferous  conodonts 


164 


PALAEONTOLOGY,  VOLUME  25 


Genus  Neoprioniodus  Rhodes  and  Muller 
Type  species:  Prioniodus  conjunctus  Gunnell  1931. 

Neoprioniodus  cf.  acampylus  Rexroad  and  Collinson 
Plate  19,  fig.  16 

1965  Neoprioniodus  acampylus  Rexroad  and  Collinson,  p.  11,  pi.  1,  figs.  25-27. 

1968  Neoprioniodus  acampylus  Rexroad  and  Collinson,  Thompson  and  Goebel,  p.  37,  pi.  3,  fig.  2. 
Remarks.  This  specimen  conforms  to  the  general  outline  and  description  of  Neoprioniodus  acampylus 
but  differs  in  possessing  a more  robust  posterior  process  which  is  laterally  thickened,  and  it  lacks 
the  flaring  inner  lip  of  that  species.  It  differs  from  N.  camurus  Rexroad  in  being  unbowed.  This 
species  has  been  recorded  only  from  the  Warsaw  and  Salem  Formations  of  the  Upper  Mississippi 
Valley  (Rexroad  and  Collinson  1965)  and  the  Meramacian  of  Indiana  (Thompson  and  Goebel 
1968).  It  occurs  at  a similar  position  in  Ravenstonedale  being  confined  to  the  Michelinia  grandis  Beds. 

Neoprioniodus  sp. 

Plate  19,  fig.  17 

Description.  Robust  unit  consisting  of  cusp  and  a short  posterior  bar.  The  posterior  bar  is  laterally  thickened, 
approximately  square  in  cross-section,  and  is  about  half  the  length  of  the  cusp,  but  is  not  complete.  Its  oral 
surface  bears  four  denticles  which  are  discrete  for  more  than  half  their  length  and  are  laterally  compressed 
but  more  strongly  so  on  the  inner  side.  They  are  incurved  and  of  varying  length,  the  highest  being  the 
penultimate  one  on  the  cusp.  The  bar  is  straight  and  is  bevelled  on  the  inner  face. 

The  cusp  is  four  times  the  length  of  the  posterior  process  denticles  and  is  strongly  incurved  and  slightly 
curved  posteriorly.  Its  outer  side  is  convex  but  the  inner  side  has  a lip  along  both  margins  which  continues 
down  to  the  anticusp.  The  anticusp  is  short  and  appears  to  be  pointed  but  the  tip  is  broken.  The  aboral  edge 
of  the  cusp  is  strongly  bevelled  but  the  bevelling  does  not  continue  down  the  anticusp. 

The  cavity  is  elliptical,  small,  and  is  continued  as  a faint  groove  beneath  the  posterior  process. 

Genus  Lonchodina  Ulrich  and  Bassler 
Type  species:  Lonchodina  typicalis  Ulrich  and  Bassler  1926. 

Lonchodina  sp. 

Plate  18,  fig.  17;  Plate  19,  figs.  1-3 

Description.  The  unit  includes  both  right  and  left  forms;  it  is  bowed.  The  cusp  is  large,  laterally  flattened 
with  sharp  edges,  strongly  inwardly  curved,  and  posteriorly  inclined  and  curved.  Typically  the  cusp  narrows 
gently  towards  the  oral  tip  but  in  some  specimens  the  base  is  very  wide.  The  processes  are  the  same  height 
and  length  and  both  are  laterally  thickened  but  the  height  is  twice  the  thickness.  The  posterior  process  is 
almost  at  right  angles  to  the  cusp  and  its  oral  surface  bears  three  to  six  laterally  compressed  denticles  of 
variable  size,  but  the  two  penultimate  ones  are  the  largest  and  one  of  them  may  be  as  large  as  the  cusp.  The 
denticles  are  inwardly  curved  and  are  in  the  same  plane  as  the  cusp,  but  are  more  strongly  posteriorly  inclined. 
The  anterior  process  bears  three  to  five  denticles  on  its  oral  surface  which  are  subequal  in  size  and  half  the 
size  of  the  cusp.  They  are  more  founded  than  the  posterior  process  denticles  and  are  posteriorly  and  inwardly 
curved.  Both  processes  are  strongly  bevelled  and  striated  on  the  inner  side  and  inclined  inwardly  on  the  upper 
side,  giving  the  processes  a triangular  cross-section.  The  inner  bevelled  side  is  markedly  striated  on  the  lower 
half.  There  is  a small  lip  over  the  basal  cavity.  The  aboral  side  is  narrow  with  a small  triangular  pit  beneath 
the  cusp  and  a narrow  groove  beneath  the  processes. 

Remarks.  The  multi-element  apparatus  to  which  this  distinctive  form  belongs  is  unknown. 

Genus  Spathognathus  Branson  and  Mehl 
Type  species:  Spathodus  primus  Branson  and  Mehl  1941. 

Spathognathodus  scitulus  (Hinde) 

Plate  19,  fig.  14 

1900  Polygnathus  scitulus  Hinde  (part),  p.  343,  pi.  9,  figs.  9,  11  only. 

1949  Spathognathodus  scitulus  (Hinde)  Youngquist  and  Miller,  p.  622,  pi.  101,  fig.  4. 


HIGGINS  AND  VARKER:  CONODONT  FAUNAS 


165 


Remarks.  Austin  and  Rhodes  recorded  a fused  cluster  consisting  of  one  specimen  of  Spathognathodus 
scitulus  and  four  specimens  of  Apatognathus  sp.  from  the  Lower  Carboniferous  of  the  Avon  Gorge. 
Undoubtedly  the  forms  have  a similar  range  and  commonly  occur  in  the  same  samples.  Baesemann 
(1973)  described  a multi-element  species,  Ozarkodina  minuta  which  includes  a P element,  S.  minutus 
which  is  very  similar  to  S.  scitulus.  Similarly,  von  Bitter  (1976)  suggested  that  S.  cristulus,  also 
morphologically  similar  to  S.  scitulus,  could  be  the  P element  of  his  Ellisonia  sp.  apparatus.  However, 
neither  Baesemann’s  nor  von  Bitters’s  accompanying  element  are  present  in  the  Ravenstonedale 
sequence,  and  the  Austin  and  Rhodes  model  may  be  the  more  likely  for  the  S.  scitulus  apparatus. 

Genus  Taphrognathus  Branson  and  Mehl 
Type  species:  Taphrognathus  varians  Branson  and  Mehl  1941. 

Taphrognathus  varians  Branson  and  Mehl 
Plate  18,  figs.  15,  16 

1941  Taphrognathus  varians  Branson  and  Mehl  p.  182,  p.  6,  figs.  27-33,  35-40. 

Remarks.  A short  carina  occurs  in  the  posterior  quarter  of  the  median  sulcus  which  occurs  on  a 
few  but  not  the  majority  of  previously  figured  specimens  of  this  species.  At  the  present  time  such 
specimens  are  included  within  Taphrognathus  varians. 


Acknowledgements.  We  thank  Dr.  W.  H.  C.  Ramsbottom  for  helpful  discussion  on  the  stratigraphy  of  the 
Ravenstonedale  area.  The  work  was  supported  by  a N.E.R.C.  research  grant  which  is  gratefully  acknowledged. 


REFERENCES 

Austin,  R.  l.  1968.  Five  Visean  conodont  horizons  in  the  north  of  England.  Geol.  Mag.  105  (4),  367-371. 

— 1973.  Modification  of  the  British  Avonian  conodont  zonation  and  a reappraisal  of  European  Dinantian 
conodont  zonation  and  correlation.  Ann.  Soc.  geol.  de  Belgique,  96,  523-532. 

— 1976.  Evidence  from  Great  Britain  and  Ireland  concerning  west  European  Dinantian  conodont 
paleoecology.  Geol.  Soc.  Canada  Sp.  Paper  No.  1 5,  202-224. 

— and  mitchell,  m.  1975.  Middle  Dinantian  platform  conodonts  from  County  Fermanagh  and  County 
Tyrone,  Northern  Ireland.  Bull.  geol.  Surv.  Gt.  Britain,  55,  43-54. 

— and  Rhodes,  f.  h.  t.  1969.  A conodont  assemblage  from  the  Carboniferous  of  the  Avon  Gorge. 
Palaeontology,  12,  400-405. 

baesemann,  j.  f.  1973.  Missourian  (Upper  Pennsylvanian)  conodonts  of  northeastern  Kansas.  J.  Paleont.  47, 
689-710. 

barnes,  c.  R.  1976  (ed.).  Conodont  paleoecology.  Geol.  Soc.  Canada  Sp.  Paper  No.  15,  1-324. 
bischoff,  G.  1957.  Die  Conodonten-Stratigraphie  des  rheno-herzynischen  Unterkarbons  mit  Berucksichtigung 
der  Wocklumeria-Stufe  und  der  Devon/Karbon  Grenze.  Abh.  hess.  Landesmt.  Bodenf,  19,  1-64. 
capewell,  J.  G.  1955.  The  Post-Silurian  pre-marine  Carboniferous  sedimentary  rocks  of  the  eastern  side  of  the 
English  Lake  District.  Q.  Jl  geol.  Soc.  Lond.  Ill,  23-46. 
collinson,  c.,  rexroad,  c.  b.  and  Thompson,  t.  l.  1971.  Conodont  zonation  of  the  North  American 
Mississippian.  In  sweet,  w.  c.  and  Bergstrom,  s.  (eds.).  Symposium  on  conodont  biostratigraphy.  Mem. 
geol.  Soc.  Amer.  127,  353-394. 

Garwood,  e.  J.,  1907.  Notes  on  the  faunal  succession  in  the  Carboniferous  Limestone  of  Westmorland  and 
neighbouring  portions  of  Lancashire  and  Yorkshire.  Geol.  Mag.  44,  70-74. 

— 1913.  The  Lower  Carboniferous  succession  in  the  northwest  of  England.  Q.  Jlgeol.  Soc.  Lond.  68, 449-596. 
george,  t.  N.  1978.  Eustasy  and  tectonics:  sedimentary  rhythms  and  stratigraphical  units  in  British  Dinantian 

correlation.  Proc.  Yorks,  geol.  Soc.  42  (13),  229-262. 

JOHNSON,  G.  A.  L.,  MITCHELL,  M.,  RAMSBOTTOM,  W.  H.  C.,  SEVASTOPULO,  G.  M.  and  WILSON,  R.  B.  1976.  A 

correlation  of  Dinantian  rocks  in  the  British  Isles.  Geol.  Soc.  London,  Special  Report  No.  7,  1 -87. 
globensky,  y.  1967.  Middle  and  Upper  Mississippian  conodonts  from  the  Windsor  Group  of  the  Atlantic 
Provinces  of  Canada.  J.  Paleont.  41,  432-448. 


166 


PALAEONTOLOGY,  VOLUME  25 


higgins,  a.  c.  1975.  Conodont  zonation  of  the  late  Visean — early  Westphalian  strata  of  the  south  and  central 
Pennines  of  northern  England.  Bull.  geol.  Surv.  Gt.  Britain.  No.  53,  1-99. 
higgins,  a.  c.  and  wagner-gentis,  c.  h.  t.  Conodont  and  Goniatite  biostratigraphy  and  stratigraphic  history 
of  the  Lower  Carboniferous  and  early  Namurian  of  the  central  Cantabrian  Mountains,  NW  Spain.  (In  press.) 
holliday,  d.  w.,  neves,  R.  and  Owens,  b.  1979.  Stratigraphy  and  palynology  of  early  Dinantian  (Carboniferous) 
strata  in  shallow  boreholes  near  Ravenstonedale,  Cumbria.  Proc.  Yorks,  geol.  Soc.  42  (19),  343-356. 
jenkins,  t.  b.  h.  1974.  Lower  Carboniferous  conodont  biostratigraphy  of  New  South  Wales.  Palaeontology, 
17,  909-924. 

Johnson,  G.  A.  l.  and  marshall,  a.  e.  1971.  Tournaisian  Beds  in  Ravenstonedale,  Westmorland.  Proc.  Yorks, 
geol.  Soc.  38,  261-280. 

kellaway,  G.  a.  and  welch,  f.  b.  a.  1955.  The  Upper  Old  Red  Sandstone  and  Lower  Carboniferous  rocks  of 
Bristol  and  the  Mendips  compared  with  those  of  Chepstow  and  the  Forest  of  Dean.  Bull.  geol.  Surv.  Gt. 
Britain,  9,  1-21. 

Merrill,  G.  k.  and  martin,  m.  d.  1976.  Environmental  control  of  conodont  distribution  in  the  Boyd  and 
Mattoon  Formations  (Pennsylvanian  Missourian)  Northern  Illinois.  Geol.  Soc.  Canada  Sp.  Paper  No.  15, 
243-271. 

ramsbottom,  w.  h.  c.  1973.  Transgressions  and  regression  in  the  Dinantian:  a new  synthesis  of  British 
Dinantian  stratigraphy.  Proc.  Yorks,  geol.  Soc.  39  (28),  567-607. 

— 1974.  Dinantian.  In  rayner,  d.  h.  and  Hemingway,  j.  e.  (eds)  The  Geology  and  Mineral  Resources  of 
Yorkshire:  Occ.  Publ.  Yorks.  Geol.  Soc.  47-73. 

— 1977.  Major  cycles  of  transgression  and  regression  (mesothems)  in  the  Nanurian.  Ibid.  41  (24),  261-291. 
rexroad,  c.  b.  1957.  Conodonts  from  the  Chester  Series  in  the  type  area  of  southwestern  Illinois.  Illinois 

geol.  Survey  Rept.  Inv.  199,  1-43. 

1958.  Conodonts  from  the  Glen  Dean  Formation  (Chester)  of  the  Illinois  Basin.  Ibid.  209,  1-27. 

— 1965.  Conodonts  from  the  Keokuk,  Warsaw  and  Salem  Formations  (Mississippian)  of  Illinois.  Ibid.  circ. 
No.  388,  1-26. 

— and  collinson,  c.,  1963.  Conodonts  from  the  St.  Louis  Formation  (Valmeyeran  Series)  of  Illinois, 
Indiana  and  Missouri.  Illinois  geol.  Surv.,  circ.  No.  355,  1-28. 

Rhodes,  f.  h.  t.,  Austin,  r.  l.  and  druce,  e.  c.  1969.  British  Avonian  (Carboniferous)  conodont  faunas,  and 
their  value  in  local  and  intercontinental  correlation.  Bull.  British  Museum  (Nat.  Hist.)  Geol.  Supplement 
No.  5,  1-313. 

Sandberg,  c.  a.  1976.  Conodont  biofacies  of  late  Devonian  Polygnathus  styriacus  zone  in  western  United 
States.  Geol.  Soc.  of  Canada,  Spec.  Paper  No.  15,  171-186. 
turner,  J.  s.  1950.  Notes  on  the  Carboniferous  Limestone  of  Ravenstonedale,  Westmorland.  Trans.  Leeds 
geol.  Assoc.  6 (3),  27-37. 

varker,  w.  j.  1967.  Conodonts  of  the  genus  Apatognathus  Branson  and  Mehl  from  the  Yoredale  Series  of 
the  North  of  England.  Palaeontology,  10,  124-141. 

— and  higgins,  a.  c.  1979.  Conodont  evidence  for  the  age  of  the  Pinskey  Gill  Beds  of  Ravenstonedale, 
north-west  England.  Proc.  Yorks,  geol.  Soc.  42  (20),  357-369. 

vaughan,  a.  1905.  The  palaeontological  sequence  in  the  Carboniferous  limestone  of  the  Bristol  area.  Q.  Jl 
geol.  Soc.  Lond.  61,  181-307. 

voges,  a.  1959.  Conodonten  aus  dem  Untercarbon  I und  II  (Gattendorfia  und  Pericyclus-Stufe)  des 
Sauerlandes.  Palaont.  Z.  33,  266-314. 

von  bitter,  p.  h.  1976.  Palaeoecology  and  distribution  of  Windsor  Group  (Visean-Early  Namurian)  conodonts. 
Port  Hood  Island,  Nova  Scotia,  Canada.  Geol.  Soc.  Canada,  Sp.  Paper  No.  15,  225-241. 

a.  c.  higgins 
Department  of  Geology 
University  of  Sheffield 
Mappin  Street 
Sheffield  SI  3JD 

W.  J.  VARKER 

Department  of  Earth  Sciences 
University  of  Leeds 
Leeds  LS2  9JT 


Typescript  received  1 May  1980 
Revised  typescript  received  4 July  1980 


SOMASTEROIDEA,  ASTEROIDEA,  AND  THE 
AFFINITIES  OF  LUIDIA  (PLATASTERIAS ) 
LATIRADIA  TA 

by  DANIEL  B.  BLAKE 


Abstract.  Important  changes  in  the  taxonomy  and  phylogenetic  interpretation  of  stellate  echinoderms  were 
proposed  during  the  1960s  by  H.  B.  Fell;  certain  of  this  author’s  ideas  are  re-evaluated.  Fell  argued  that 
the  extant  west  American  sea  star  Platasterias  latiradiata  Gray  is  a surviving  member  of  the  otherwise 
Palaeozoic  Somasteroidea.  The  extant  family  Luidiidae  was  considered  primitive  among  true  asteroids  and 
it  was  included  with  the  Palaeozoic  family  Palasteriscidae  in  the  order  Platyasterida.  The  skeletal  arrangement 
of  Platasterias  and  Luidia  was  interpreted  as  having  been  derived  with  relatively  limited  change  from  a 
currently  unknown  Cambrian  crinoid  ancestry.  It  is  argued  here  that  Platasterias  is  not  a somasteroid  but  a 
subgenus  of  Luidia,  and  the  Luidiidae  is  returned  to  the  large  living  order  Paxillosida.  The  origin  of  the 
morphology  of  Luidia,  including  Platasterias,  is  related  to  sea  star  behaviour  and  habitat  rather  than  a crinoid 
ancestry.  The  Luidiidae  is  not  considered  to  be  of  major  importance  in  delineating  asteroid  phylogeny. 

Several  papers  published  during  the  1960s  by  H.  B.  Fell  (1962a,  b;  1963  a-c;  1967)  developed  a 
series  of  stimulating  and  intriguing  hypotheses  on  stellate  echinoderm  phylogeny  that  significantly 
altered  viewpoints  of  the  history  of  these  organisms.  Certain  of  these  ideas  are  reconsidered  in  this 
paper. 

Prior  to  Fell’s  work,  three  classes  or  subclasses  were  recognized  among  stellate  echinoderms; 
the  Asteroidea  (sea  stars  or  starfish  ),  Ophiuroidea  (brittle  stars  or  serpent  stars  and  basket  stars), 
and  the  Somasteroidea,  the  last  a taxon  of  primitive  echinoderms  then  considered  to  be  restricted 
to  lower  and  middle  Palaeozoic  rocks.  Relationships  among  the  groups  were  unclear.  Because  of 
disparate  early  development,  many  biologists  (e.g.  Hyman  1955)  believed  the  ophiuroids  and 
asteroids  had  independent  origins,  and  were  only  secondarily  similar  in  certain  features.  These 
workers  found  (and  still  find)  it  undesirable  to  combine  ophiuroids  and  asteroids  in  formal  higher 
taxa  below  the  phylum  level,  e.g.  the  class  Stelleroidea  and  subphylum  Asterozoa  sensu  Spencer 
and  Wright  1966. 

Other  workers,  stressing  the  fossil  record,  were  of  the  opinion  that  the  two  living  groups  shared 
a common  origin,  and  that  their  ancestors  or  near  ancestors  could  be  recognized  among  the  fossil 
somasteroids.  Further,  Fell  (e.g.  1948)  considered  that  the  developmental  arguments  for  separating 
the  asteroids  from  ophiuroids  were  unconvincing.  These  workers  tend  to  believe  that  the  two  should 
| be  combined  in  a phylogenetically  unified  higher  taxon  below  the  phylum  level  (e.g.  Spencer  and 
Wright  1966). 

In  his  work  on  phylogeny.  Fell  developed  a number  of  interrelated  topics:  (1)  the  nature  of  the 
Somasteroidea;  (2)  the  relationships  among  the  extant  sea  stars  Platasterias,  Luidia,  the  remaining 
asteroids  and  the  Palaeozoic  Somasteroidea;  (3)  the  relationship  between  the  extant  ophiuroid 
Ophiocanops  and  the  Palaeozoic  ophiuroids;  and  (4)  the  phylogenetic  relationships  among 
somasteroids,  asteroids,  and  ophiuroids.  Fell  proposed  a hypothesis  for  the  origins  of  stelleroids 
from  an  inferred  crinoid  ancestry,  suggested  a sequence  of  evolutionary  events  leading  from  the 
Somasteriodea  to  the  Ophiuroidea  and  Asteroidea,  and  transferred  a number  of  living  taxa  to 
groups  previously  considered  to  be  of  exclusively  Palaeozoic  occurrence.  These  included  the  family 
Luidiidae  from  the  Paxillosida  to  the  Platyasterida,  both  within  the  Asteroidea;  the  genus 

IPalaeontology,  Vol.  25,  Part  1,  1982,  pp.  167-191,  pis.  20-22.| 


168 


PALAEONTOLOGY,  VOLUME  25 


Platasterias  from  the  Asteroidea  (Paxillosida)  to  the  Somasteroidea;  and  the  ophiuroid  Ophiocanops 
fugiens  to  the  Oegophiurida  from  the  Phrynophiurida.  Fell’s  re-evaluations  were  enthusiastically 
received  by  some  workers,  but  a number  of  his  ideas  were  challenged,  including  the  arguments  for 
the  derivation  of  asterozoans  from  crinoids,  by  Philip  (1965),  answered  by  Fell  (1965);  the  affinities 
of  Ophiocanops,  by  Hotchkiss  (1977);  and  the  affinities  of  Platasterias  by  several  workers  (e.g. 
Madsen  1966;  Blake  1967,  1972,  1973;  Pearse  1969;  Algor  1971). 

A survey  of  textbooks  and  papers  published  during  the  later  1960s  and  1970s  shows  a continuing 
uncertainty  as  to  how  Platasterias  latiradiata  should  be  treated,  yet  the  question  of  affinities  of  the 
species  is  an  important  one.  Not  only  is  a living  fossil  intriguing  simply  as  a survivor,  but  its 
existence  raises  questions  on  the  nature  of  evolutionary  processes.  Further,  once  ranked  as  a ‘living 
fossil’,  a species  will  become  the  basis  for  reconstruction  of  the  biology  of  its  presumed  close  fossil 
relatives,  a fact  illustrated  by  Fell’s  (19626,  p.  2)  explanation  of  purpose  for  one  of  his  papers: 
‘This  contribution  is  limited  to  brief  discussion  of  the  major  features  of  somasteroid  anatomy,  as 
illustrated  by  Platasterias .’  Platasterias  as  a somasteroid  clearly  will  strongly  influence  interpreta- 
tions of  stellate  echinoderm  biology  and  history. 

The  concern  of  this  paper  is  primarily  with  the  second  of  the  four  topics  cited  above,  the  nature 
of  Platasterias  and  the  Luidiidae.  Other  problems  in  the  nature  of  relationships  among  stellate 
echinoderm  groups  are  in  need  of  restudy,  but  these  ideas  are  beyond  the  purposes  of  this  paper. 

I will  argue  that  P.  latiradiata  should  be  considered  a subgenus  of  Luidia,  itself  a genus 
unequivocally  included  in  the  Asteroidea  (rather  than  the  Somasteroidea)  by  Fell  (1963a).  Assigning 
Platasterias  as  a subgenus  of  Luidia  does  not  in  itself  challenge  Fell’s  ideas  on  somasteroid/asteroid 
phylogeny;  Luidia  (including  Platasterias)  might  still  be  the  primitive  extant  asteroid  with  many 
features,  as  outlined  by  Fell,  derived  with  little  modification  from  the  somasteroids.  In  re-evaluating 
the  taxonomic  and  phylogenetic  position  of  the  monogeneric  Luidiidae,  I will  next  argue  that  there 
is  no  clear  connection  between  Palaeozoic  somasteroids  and  the  luidiids  (including  Platasterias), 
and  that  the  Luidiidae  should  be  transferred  from  the  order  Platyasterida  to  the  order  Paxillosida. 
A functional,  rather  than  phylogenetic,  hypothesis  for  the  origin  of  luidiid  morphology  is  proposed. 


SOMASTEROIDEA 

The  concept  of  the  Somasteroidea  originated  with  Spencer  (1951)  and  evolved  with  the  work  of 
Fell  (1962a,  b;  1963 a-c;  1967),  Spencer  and  Wright  (1966),  and  McKnight  (1975).  In  spite  of  this 
effort,  the  group  is  difficult  to  characterize  and  therefore  comparisons  between  somasteroids  and 
other  organisms  are  difficult  as  well.  Relevant  fossils  are  in  need  of  restudy  and  reillustration.  I 
have  based  the  following  diagnosis  on  the  literature  (and  not  original  study  of  the  fossils),  and  I 
believe  it  represents  the  fossil  somasteroids  as  pictured  by  those  workers  who  include  Platasterias 
in  the  taxon.  In  keeping  with  the  arguments  presented  here,  however,  I have  removed  characters 
derived  from  Platasterias  only.  Although  such  a diagnosis  logically  follows  arguments  for  transfer 
of  Platasterias,  a concept  of  the  Somasteroidea  is  necessary  for  the  comparisons  that  follow. 

Subclass  Somasteroidea  Spencer,  1951 

Asterozoans  in  which  the  axial  skeleton  consists  of  ambulacrals  in  a double  series,  usually  paired  but  apparently 
alternating  in  some  species.  Each  ambulacral  gives  rise  to  a transverse  series  of  ossicles;  in  apparently  primitive 
species,  these  ossicles  are  similar  and  rod-like  elements  termed  virgals,  but  in  more  advanced  species  the 
virgals  are  differentiated  into  adambulacrals,  marginals,  and  related  ossicles.  A permanent  ambulacral  furrow 
or  groove  is  lacking  so  that  the  long  axis  of  the  ambulacral  is  approximately  linear  and  horizontal,  and  the 
ossicles  essentially  lie  in  the  plane  of  the  remaining  ossicles  of  the  oral  surface.  The  adambulacral  (or  first 
virgal)  (different  spellings  of ‘virgal’  and  its  plural  form  have  been  used;  I have  followed  Spencer  and  Wright 
1966)  generally  abuts  the  lateral  (abradial)  margin  of  the  ambulacral,  or  overlap  is  slight.  In  some  species 
the  ambulacrals  probably  could  be  raised  to  form  temporary  ambulacral  furrows.  A large  or  small  radial 
channel  for  the  radial  water  canal  is  present  along  the  oral  margin  of  the  line  of  juncture  of  pairs  of  ambulacrals, 
or  it  is  more  or  less  enclosed  by  the  ambulacrals.  The  tube-feet  are  seated  in  broad  basins  which  in  some 
species  communicate  to  the  body  cavity.  Jaw  ossicles  are  differentiated,  but  odontophores  are  lacking.  Open 


BLAKE:  SEA  STAR  PHYLOGENY 


169 


text-fig.  1 . Morphology  of  the  Somasteroidea,  modified  slightly 
after  Spencer  (1951)  and  Fell  (1963a).  a,  abactinal  arrangement 
of  Sturzaster  marstoni  (Salter),  morphology  considered  by  Spencer 
to  be  close  to  that  of  Chinianaster;  Spencer  (1951,  p.  95).  b.  ambu- 
lacrals  of  Archegonaster  pentagonus;  Spencer  (1951,  p.  104). 
c,  metapinnules  with  cover  plates,  Chinianaster  levyi\  Fell  (1963a, 
p.  394).  d,  interpretation  of  arm  of  Ampullaster  ubaghasi,  view  of 
oral  surface;  Fell  (1963a,  p.  394).  e,  reconstruction  of  a part  of 
the  arm  of  Chinianaster,  water  vascular  system  darkened;  Fell 
1963a,  p.  402).  f,  interpretation  of  arm,  near  tip,  of  Villebrunaster 
thoralr.  Fell  (1963a,  p.  394).  Key:  1 , ambulacrals;  2,  adambulacrals; 
3,  virgals;  4,  marginals.  Illustrations  courtesy  of  H.  Barraclough 
Fell  and  the  Royal  Society. 


buccal  slits  might  be  present  at  the  mouth  frame  in  at  least  some  species.  Aboral  ossicles  are  paxilliform, 
with  delicate  tetradiate  bases.  Encrusting  ossicles  or  spinelets  are  present  at  least  on  some  ossicles  of  the 
ventral  surface.  Over  all,  the  skeleton  beyond  the  ambulacrals  is  quite  delicate  (text-fig.  1). 

Discussion.  Most  of  these  characters  are  also  expressed  among  species  assigned  to  the  Asteroidea, 
but  critical  to  the  concept  of  the  Somasteroidea  is  an  ambulacral/adambulacral  arrangement  in 
which  an  ambulacral  furrow  or  groove  is  lacking  (text-fig.  2).  Spencer  (1951,  p.  88)  says  in  reference 
to  somasteroids  that  ‘.  . . these  first  stages  show  no  sign  of  an  ambulacral  furrow’,  and  Fell  (1963a, 
p.  389)  used  the  development  of  a furrow  or  groove  as  the  basic  character  separating  somasteroids 
from  asteroids.  The  ambulacral  groove  or  furrow  is  different  from  the  ambulacral  channel  for  the 
radial  water  canal;  for  further  discussion,  see  under  ambulacral  column  arrangement. 


170 


PALAEONTOLOGY,  VOLUME  25 


In  somasteroids  the  abradial  margin  of  the  ambulacral  either  abuts  the  adradial  side  of  the 
adambulacral  (first  virgal),  or  the  ambulacral  can  overlap  the  adambulacral  to  a very  limited  extent 
(text-fig.  2).  In  his  discussion  of  somasteroids,  Fell  (1963a,  p.  393)  says  ‘Each  metapinnule  arises 
from  the  abradial  margin  of  one  of  the  paired  ambulacral  (or  brachial)  elements  . . .’.  The  long 
axis  of  the  ambulacral  is  approximately  horizontal  and  the  basin  for  the  tube-foot  lies  close  to  the 
plane  of  the  oral  surface. 

In  true  asteroids  the  ambulacral  rests  on  the  aboral  surface  of  the  adambulacral,  rather  than 
laterally  adjacent  to  it  (text-fig.  2).  In  addition,  the  axis  of  the  adradial  end  of  the  ambulacral  is 
oblique  to  that  of  the  general  ambulacral/adambulacral  articulation  surfaces.  These  arrangements 
produce  a permanent  furrow  with  the  basin  for  the  tube-foot  elevated. 

McKnight  (1975)  argued  that  the  lack  of  an  odontophore  is  an  important  character  unifying 
somasteroids.  The  odontophore  is  a small,  unpaired  typically  T-shaped  ossicle  present  between 
members  of  a jaw  ossicle  pair.  It  appears  to  brace  and  thus  strengthen  the  jaw  apparatus,  and  its 
phylogenetic  development  might  reflect  the  changing  food  habits  suggested  by  Spencer  (1951), 
away  from  the  small  particle  feeding  inferred  for  somasteroids  to  the  large  particle  feeding  present 
in  many  living  asteroids. 


A 


text-fig.  2.  Stylized  diagrams  oriented  transverse  to  arm  axes, 
showing  arrangement  of  1 , ambulacrals,  and  2,  adambulacrals 
in  a,  somasteroids  ( Chinianaster  arrangement  to  left,  Ampul- 
laster  to  right),  and  b,  asteroids.  The  overlap  of  the  ambulacral 
on  the  adambulacral  in  Ampullaster  is  very  small;  compare  to 
text-fig.  Id.  3,  radial  water  channel.  Based  on  work  of  Spencer 
(1951)  and  Fell  (1963a),  especially  the  latter’s  fig.  7,  p.  395. 


COMPARISONS  BETWEEN  PLATASTERIAS,  FOSSIL  SOMASTEROIDS 
AND  TRUE  ASTEROIDS 

Platasterias  is  compared  to  fossil  somasteroids  and  living  asteroids,  especially  Luidia  as  traditionally 
recognized,  in  the  following  somewhat  overlapping  sequence:  (1)  over-all  body  form;  (2)  nature  of 
the  skeleton;  (3)  ambulacral  column  arrangement;  (4)  growth  gradients;  (5)  other  morphologic 
features;  (6)  feeding.  Fell’s  view  of  Platasterias  as  a somasteroid  was  derived  in  large  part  from 
his  evaluation  of  ambulacral  column  arrangement  and  growth  gradients.  A discussion  of 
the  taxonomic  positions  of  Platasterias  and  Luidia  sensu  Fell  therefore  must  wait  until  section  4. 
Each  section  begins  with  a summary  of  the  ideas  of  Spencer  (1951),  Fell  (1962a  et  seq.),  and  Spencer 
and  Wright  (1966)  and  continues  with  my  comments.  I have  tried  to  include  all  major  arguments 
on  Platasterias,  but  the  discussion  is  not  comprehensive  on  other  topics,  such  as  the  morphology 


BLAKE:  SEA  STAR  PHYLOGENY 


table  1.  Changes  in  the  distribution  of  important  characters  in  the  inferred  somasteroid/asteroid  phylogeny, 
as  envisaged  by  H.  B.  Fell.  The  diagram  was  prepared  by  the  writer,  based  on  his  reading  of  Fell  (1963a, 
pp.  389  ff.;  19636)  and  it  is  intended  as  an  aid  to  understanding  Fell’s  ideas  of  the  major  events  in  his  inferred 
phylogeny  (represented  by  the  sequence  1 through  6)  rather  than  as  an  exact  character  distribution  summary. 
Dotted  lines  mean  a character  is  present  in  only  some  members  of  a division,  and  minor  exceptions  occur. 
For  example,  although  the  family  is  not  entered  separately,  the  Porcellanasteridae  lack  an  anus  and  suctorial 
tube-feet,  whereas  other  sea  stars,  such  as  certain  Astropecten  and  Asterias  species,  have  subpetaloid  arms. 
Notes:  (1)  virgals  similar  in  Division  1,  dissimilar  in  Division  2;  (2)  development  of  actinals  becomes  stronger 
through  the  sequence;  (3)  no  entry  in  Fell  (1963a)  for  this  division.  The  terminal  is  the  unpaired  ossicle  at 
the  distal  tip  of  the  arm. 

Among  somasteroids,  the  monogeneric  Chinianasteridae  was  considered  primitive,  based  largely  on  the 
presence  of  undifferentiated  virgals  and  robust  ambulacrals,  and  the  absence  of  communication  pores  between 
ambulacrals  for  ampullae.  Next  in  the  sequence,  but  not  separated  from  the  Chinianasteridae  in  Fell  1963a, 
came  the  Villebrunasteridae,  including  Villebrunaster  and  Ampullaster.  Here,  virgals  are  differentiated  and 
pores,  inferred  for  internal  ampullae,  were  present.  Platasterias  (Platasteriidae)  was  considered  to  have  common 
features  with  both  the  Chinianasteridae  and  the  Villebrunasteridae  but  appeared  to  be  about  at  the 
villebrunasterid  grade  based  on  the  inferred  differentiation  of  the  virgals,  and  the  presence  of  internal  ampullae 
(19636,  p.  144).  The  monogeneric  Archegonasteridae,  not  included  in  the  table,  was  considered  to  be  an 
advanced  somasteroid,  specialized  in  body  shape  and  lack  of  interradial  slits,  metapinnule  reduction, 
and  development  of  the  robust  marginals.  True  asteroids  originated  with  development  of  a permanently  erect 
ambulacral  furrow. 


< 


w Group  1 

co  1 . Chinianasteridae 
< 

2.  Platasteriidae 
Group  2 

< 3.  Platanasteridae 

L±J 

Q 

O 4.  Luidiidae 

cr 

tu 

H Group  3 

*■*  5.  Astropectinidae 

6.  Most  other 
asteroids 


of  the  fossils.  Fell’s  (1963a,  b)  major  ideas  on  somasteroid/asteroid  phylogeny  are  summarized  in 
Table  1. 

McKnight  (1975)  modified  the  concept  of  the  Somasteroidea  and  transferred  five  Palaeozoic 
genera  to  this  taxon.  His  modification  is  included  here,  but  I did  not  attempt  to  re-evaluate 
somasteroids  in  light  of  the  transferred  taxa  because  properties  of  these  genera  were  not  incorporated 
in,  and  do  not  appear  to  bear  directly  on.  Fell’s  ideas  on  the  position  of  Platasterias  and  Luidia. 

1.  Over-all  body  form 

Spencer  (1951,  p.  91)  described  somasteroids  as  having  a large  central  body  in  which  the  arms 
are  parts  of  the  oral  surface,  i.e.  . . the  arms  are  just  beginning  to  be  differentiated’.  Fell  (1963a) 
described  somasteroids  as  extremely  flattened  but,  more  important,  as  asterozoans  with  petaloid 
arms.  Fell  (1963a,  fig.  5)  presented  a sequence  of  diagrams  illustrating  the  inferred  transformation 
of  arm  shape  beginning  with  the  petaloid  arm  of  a monoserial  crinoid,  and  ending  with  the  triangular 


172 


PALAEONTOLOGY,  VOLUME  25 


arm  of  the  extant  astropectinid  Plutonaster.  The  petaloid  arm  of  Platasterias  was  placed  by  Fell 
between  the  petaloid  arms  of  a chiniansterid  somasteroid  and  a somewhat  weakly  petaloid 
platanasterid  asteroid.  Petaloid  arms  thus  were  considered  to  be  subdued  in  the  most  primitive  of 
Asteroidea,  the  fossil  Platyasterida,  and  they  are  absent  from  later  groups. 

Spencer  (1951)  did  not  consider  the  somasteroids  to  be  flattened,  rather  he  saw  them  as  flexible, 
noting  (p.  93):  ‘At  one  time  the  body  is  strongly  compressed,  at  another  the  body  of  the  same 
species  is  well  rounded.’ 

Discussion.  P.  latiradiata  is  a relatively  flat  species,  but  such  a body  shape  is  not  restricted  to 
somasteroids,  e.g.  see  such  extremely  flattened  living  asterinids  as  Anseropoda  and  Stegnaster. 

Although  the  arms  of  Platasterias  are  strongly  petaloid  and  the  disc  deeply  notched  interbrachially 
(PI.  20,  fig.  1),  comparable  shapes  are  known  elsewhere,  albeit  more  weakly  developed.  Examples 
include  Astropecten  regalis  (PI.  20,  fig.  2)  (Paxillosida)  and  Asterias  forbesi  (Forcipulatida)  (PI.  20, 
fig.  8).  Inferomarginals  in  Astropecten  regalis,  like  those  in  Platasterias,  are  transversely  elongate 
(Blake  1973,  pi.  14)  and  the  body  is  flattened.  These  similarities  among  taxonomically  widely 
separated  species  suggest  convergence  as  an  alternative  hypothesis  for  origin  of  body  shape,  perhaps 
under  conditions  such  as  those  suggested  by  Madsen  (1966):  ‘I  assume  the  petaloid  arms  in 
Platasterias  (brought  about  by  the  transverse  elongation  of  the  adambulcralia  and  inferomargin- 
alia)  to  be  secondary  adjustment  to  a life  on  a shifting  sandy  bottom  (and  perhaps  a primarily 
ciliary  method  of  feeding).’  Fell  (1962a,  p.  634)  noted  that  the  extreme  narrowness  of  the  ambulacral 
furrow  in  Platasterias  is  suggestive  of  the  fossil  somasteroids,  although  he  observed  that  this  could 
be  secondary.  Narrow  furrows  occur  in  other  modern  taxa,  for  example,  the  Ophidiasteridae 
(Valvatida)  and  the  Echinasteridae  (Spinulosida). 

2.  Nature  of  the  skeleton 

Spencer  (1951)  considered  Chinianaster  and  Villebrunaster  to  be  the  primitive  somasteroids.  In  these 
genera  he  recognized  three  types  of  ventral  ossicles:  the  ambulacrals  arranged  in  a double  row,  the 
mouth  angle  ossicles,  and  the  rod-like  intermarginals  or  virgals  arranged  in  transverse  series 
extending  from  each  ambulacral  ossicle.  Fell  used  the  term  ‘metapinnules’  for  the  rows  of  virgals, 
after  the  inferred  homologous  pinnules  of  crinoids.  In  more  advanced  somasteroids  the  morpho- 
logically simple  virgals  became  differentiated  to  form  marginals,  adambulacrals,  and  other  ossicles. 


EXPLANATION  OF  PLATE  20 

Figs.  1,  5,  6,  10,  12.  L.  ( Platasterias ) latiradiata  (Gray).  1,  over-all  aboral  view  showing  petaloid 
arms,  alignment  of  abactinals,  and  similarity  in  marking  to  that  of  L.  clathrata,  see  PI.  22,  x 1.  5,  lateral 
view  of  inferomarginal,  furrow  left,  row  of  processes  are  articular  facets  linking  successive  inferomarginals, 
x 9.  6,  adambulacral,  proximal  view,  furrow  right;  one  muscle  groove  (upper,  light  arrow)  and  one 
articular  facet  (lower,  dark  arrow)  marked;  these  and  other  structures  correspond  in  positions  with  similar 
structures  in  L.  clathrata,  fig.  4 (see  Blake  1972,  1973;  Heddle  1967  for  further  discussion).  10,  aboral  view 
of  cleared  arm  showing  enlarged  superomarginal  row  (arrow)  arising  at  terminal,  as  in  L.  clathrata,  see 
fig.  9,  x 3.  12,  aboral  view  of  uncleared  arm  showing  abactinal  ossicle  and  granule  development,  compare 
to  L.  clathrata,  fig.  11,  x 3. 

Fig.  2.  Astropecten  regalis  Gray.  Oral  view  showing  petaloid  arms  and  well  defined  marginal  spines  (arrow),  x 1 . 

Figs.  3,  4,  9,  11.  Luidia  clathrata  (Say).  3,  lateral  view  of  inferomarginal,  furrow  left,  x9.  4,  adambulacral, 
inclined  proximal  view,  furrow  right,  see  discussion  for  fig.  6,  above,  x 9.  9,  inclined  aboral  view  of  cleared 
arm,  see  discussion  for  fig.  10,  above,  x 3.  11,  aboral  view  of  uncleared  arm,  see  discussion  for  fig.  12, 
above,  x 3. 

Fig.  7.  Luidia  neozelanica  Mortensen.  Lateral  view  of  inferomarginal,  compare  with  figs.  3,  6;  note  outline 
and  lack  of  articular  facet  row  in  this  species;  see  text  for  further  discussion,  x 9. 

Fig.  8.  Asterias  sp.  Aboral  view  showing  petaloid  arms  in  which  an  interbrachial  notch  (arrow)  nearly  reaches 
the  madreporite,  x 


PLATE  20 


BLAKE,  living  sea  stars 


174 


PALAEONTOLOGY,  VOLUME  25 


The  change  in  terminology  marks  this  differentiation.  Some  virgals  were  partially  (Archegon- 
asteridae)  or  completely  lost  (Archophiactinidae)  (Spencer  1951). 

Spencer  (1951)  did  not  describe  marginals  in  Villebrunaster  from  his  relatively  incomplete  material, 
but  Fell  ( 1 963 6,  p.  1 44)  pointed  out  that  these  ossicles  can  be  recognized  in  more  complete  specimens. 
Fell  (19626,  p.  66;  1963a,  p.  396)  further  noted  that  there  are  no  adambulacrals,  superambulacrals, 
or  inferomarginals  in  Chinianaster . The  metapinnules  of  Chinianaster  terminate  in  a free  radiole, 
as  in  Villebrunaster  (Fell  19626,  p.  16).  These  patterns  are  important  in  interpreting  Chinianaster  as 
the  primitive  somasteroid,  and  Villebrunaster  as  a second  step  in  somasteroid  evolution. 

The  ambulacrals  of  somasteroids  (text-fig.  1)  were  described  as  stout  and  block-like,  or  a lateral 
wing  is  developed;  they  form  a sheath  or  channel  for  the  radial  water  vessel.  The  virgals  are  rod-like 
and  form  the  walls  of  channels  separating  successive  metapinnules.  The  virgals  of  Villebrunaster 
were  described  (Fell  19626,  p.  16)  as  having  a flattened  base  and  a flanged  keel  (text-fig.  lc,  e) 
and  thus  were  seen  as  similar  to  those  of  Platasterias , but  Fell  (1963a,  p.  422)  cautioned  that  these 
ossicles  proved  morphologically  plastic  in  later  evolution. 

The  abactinal  skeleton  of  fossil  somasteroids  (text-fig.  1a)  consists  of  an  open  meshwork  of 
paxilliform  ossicles,  each  consisting  of  a slender,  vertically  oriented  axial  stalk  bearing  a number 
of  elongate  basal  flanges.  Fell  (19626,  p.  14)  noted  that  Chinianaster  abactinals  are  quite  similar 
to  those  of  Platasterias  and  Luidia  but  less  similar  to  those  of  the  astropectinids  in  that  in  the 
latter  the  base  is  disc-like  and  lacks  basal  projections,  and  the  stalk  is  relatively  stout. 

In  living  sea  stars,  an  unpaired  ossicle  termed  the  odontophore  is  found  between  members  of 
an  oral  ossicle  pair.  In  certain  Palaeozoic  asteroids,  Spencer  (1919  in  1914-1940)  concluded  that 
this  ossicle  was  derived  from  an  occluded  inferomarginal,  but  because  of  the  distribution  of  ossicles 
about  the  jaw  region  in  Chinianaster  and  Villebrunaster,  Fell  (1963a,  p.  401,  fig.  8)  suggested  that 
in  Platasterias  an  analogous  T-shaped  ossicle  was  derived  from  an  occluded,  non-metapinnular 
tegminal  ossicle.  McKnight  (1975),  however,  argued  that  an  odontophore  (and,  presumably,  the 
analogous  T-shaped  ossicle  as  well)  is  lacking  from  adult  somasteroids.  Stressing  the  significance  of 
the  absence  of  this  ossicle  McKnight  (1975)  transferred  the  Flelianthasteridae  and  three  genera  of 
the  Taeniactinidae  to  the  Somasteroidea  from  the  Spinulosida  sensu  Spencer  and  Wright  (1966). 
The  somasteroid  skeleton  beyond  the  ambulacral  column  was  described  by  Spencer  (1951,  p.  93) 
as  ‘slightly  built’. 

Discussion.  The  primary  similarity  between  Platasterias  and  fossil  somasteroid  ossicle  form  seems 
to  be  in  transverse  elongation,  and,  perhaps,  the  development  of  keel-  and  flange-shaped  ossicles, 
as  in  Chinianaster  and,  to  some  extent,  Platasterias. 

Virgals  differentiated  as  marginals  are  absent  from  Chinianaster,  the  inferred  primitive  somasteroid 
(Fell,  19636,  p.  144).  Those  of  Villebrunaster  (text-fig.  If),  representing  the  inferred  next  phylogenetic 
step  (Fell,  19636),  are  rather  simple  cylinders  elongate  parallel  to  the  arm  axis.  They  are  not 
transversely  elongate  as  is  the  case  in  Platasterias  (PI.  20,  fig.  5;  PI.  21,  figs.  2,  5)  and  as  presumably 
would  be  true  in  a primitive  somasteroid  under  strong  influence  of  transverse  gradients  (see  below). 
Illustrated  ambulacrals  (Fell,  1963a,  figs.  5,  6;  text-fig.  1,  herein)  of  fossil  somasteroids  are  relatively 
robust  with  a broadened  adradial  head  and  an  abradial  wing  deflected  distally,  whereas  those  of 
Platasterias  (PI.  21,  fig.  2)  are  wide  and  short,  lacking  the  broadened  head.  In  over-all  outline,  the 
ambulacrals  of  Platasterias  thus  seem  to  be  more  primitive  even  than  those  of  Chinianaster  (although 
ambulacrals  were  not  considered  to  be  derived  from  virgals).  Chinianaster  lacks  differentiated 
adambulacrals,  whereas  those  of  Villebrunaster  (Fell  1963a,  fig.  6;  text-fig.  1,  herein)  apparently 
are  simple  ossicles  elongate  parallel  to  the  arm  axis,  rather  than  transversely  elongate  as  is  the  case 
in  Platasterias  (PI.  20,  fig.  5;  PI.  21,  fig.  5).  There  are  no  known  fossils  of  adambulacral  shape 
similar  to  those  of  Platasterias. 

In  contrast,  ossicle  morphology  of  Platasterias  is  very  close  to  that  of  Luidia  clathrata,  differing 
significantly  only  in  degree  of  transverse  elongation  (PI.  21,  figs.  1,  2;  text-fig.  3).  The  abactinals, 
marginals  (PI.  20,  figs.  3,  5)  and  adambulacrals  (PI.  20,  figs.  4,  6)  are  particularly  similar.  This  is 
true  not  only  of  the  form  itself,  but  also  of  the  complex  articulation  structure  arrangement  of  the 


BLAKE:  SEA  STAR  PHYLOGENY 


.75 


ambulacral  column,  involving  both  ambulacrals  and  adambulacrals  (Blake  1972,  1973).  The 
abactinal  surface  of  the  adradial  end  of  the  ambulacral  is  truncated  in  Platasterias,  but  the  basic 
form  of  these  ossicles  is  the  same  as  in  Luidia.  Fell  (19626)  stressed  the  presence  of  a Y-shaped 
groove  on  the  actinal  surface  of  the  ambulacral  of  Platasterias,  pointing  out  that  the  structure 
also  occurs  on  fossil  somasteroids,  and  he  suggested  that  the  associated  muscle  is  used  to  temporarily 
elevate  the  ambulacral  furrow.  This  muscle  depression,  however,  occurs  on  all  modern  asteroids. 

Fell  (19626,  p.  7)  also  noted  in  Platasterias  that  the  tube-feet  emerge  from  a broad,  basin-like 
depression.  The  transverse  section  of  the  middle  of  the  long  axis  of  the  ambulacrals  of  most  asteroid 
species  (except,  for  example,  the  compressed  ossicles  of  certain  asteriids  and  asterinids  and  the 
delicate  ossicles  of  the  pterasterids)  is  cylindrical,  hence  a broad,  basin-like  depression  is  typical 
of  most  species. 

Algor  (1971)  pointed  out  that  Platasterias  has  the  ambulacral  system  characteristic  of  modern 
asteroids,  and  he  concluded  that  Platasterias  is  in  no  way  primitive.  Algor  (1971)  differentiated 
between  ancient  and  modern  asteroids  on  the  basis  of  articulation  structures  across  the  furrow, 
arguing  these  muscles  and  facets  were  weakly  developed  in  Palaeozoic  taxa  compared  to  the  sturdy 
patterns  seen  in  living  species.  Algor’s  sample  of  the  Palaeozoic  species  was  too  small,  for  sturdy 
articulation  structures  comparable  to  that  of  modern  species  do  occur,  e.g.  in  Promopalaester 
magnificus  (Miller)  and  P.  dyeri  Meek.  Important,  however,  is  that  Platasterias  is  constructed  in 
the  same  pattern  as  Luidia,  rather  than  in  the  weakly  articulated  manner  described  for  the 
somasteroids. 

Superambulacrals  are  present  in  both  Platasterias  and  Luidia  but  they  have  not  been  reported 
from  fossil  somasteroids.  Inferomarginals  of  both  Platasterias  latiradiata  and  L.  clathrata  are 
step-shaped  with  overlapping,  multifaceted  contact  points  (PI.  20,  figs.  3,  5)  (Blake  1973).  The 
superomarginals  are  relatively  large  in  Platasterias  but  their  basic  form  is  the  same  as  those  of  L. 
clathrata.  Arrangement  and  basic  morphology  of  mouth  frame  ossicles  are  essentially  the  same 
between  P.  latiradiata  and  L.  clathrata,  although  the  odontophore  is  absent  from  fossil  somasteroids 
(McKnight  1975).  Distal  arm  ossicles,  which  are  not  transversely  elongate,  are  essentially  identical 
between  P.  latiradiata  and  L.  clathrata,  and  much  closer  to  each  other  than  they  are  to  those  of 
species  belonging  to  any  other  genus. 

Fell  (19626,  p.  15)  suggested  that  the  abactinals  of  P.  latiradiata  are  similar  to  those  of 
L.  neozelanica,  however,  paxilliform  abactinals  occur  in  all  of  the  living  sea  star  orders  except  the 
Forcipulatida.  A survey  of  the  diagrams  of  Fisher  (1911)  shows  that  basal  projections  are  present 


text-fig.  3.  Cross-sections  of  arms  of  A,  Luidia  ( Platasterias ) latiradiata-,  and  b,  Luidia  ( Pe  talas  ter ) 
clathrata.  In  life,  arms  are  flexible  and  ossicle  orientations  vary  with  behaviour;  inferomarginals  can  be 
more  steeply  inclined  to  the  horizontal  than  shown  here  or  in  PI.  21,  fig.  1.  Key:  1,  abactinals;  2, 
superomarginals;  3,  inferomarginals;  4,  superambulacrals;  5,  actinals;  6,  ambulacrals;  7,  adambulacrals. 


176 


PALAEONTOLOGY,  VOLUME  25 


in  abactinals  of  taxonomically  diverse  species  and  that  the  slender  abactinals  illustrated  for  several 
of  the  spinulosidan  families  (e.g.  Solasteridae,  Korethrasteridae,  Pterasteridae)  are  closer  to 
Spencer’s  somasteroid  illustration  (1951,  fig.  5)  than  are  those  of  Platasterias  or  Luidia.  The 
abactinals  of  P.  latiradiata  and  L.  clathrata  are  very  similar  to  each  other,  however. 

Similarities  extend  to  exterior  morphology  as  well  (PI.  20,  figs.  1,  11,  12;  PI.  22),  as  was 
demonstrated  by  Madsen  (1966)  for  P.  latiradiata  and  L.  marginata. 

I have  argued  on  the  basis  of  skeletal  morphology  that  P.  latiradiata  clearly  is  very  close  to  L.  clathrata , but 
shows  only  relatively  minor  similarities  to  fossil  somasteroids.  Fell  dissected  specimens  of  Luidia;  if  the 
similarities  are  this  strong,  why  was  he  not  struck  by  them  as  well?  I believe  the  answer  lies  in  the  species  of 
Luidia  available  to  Fell  for  dissection.  For  forty-three  species  of  Luidia,  Doderlein  (1920)  recognized  ten 
subgenera  and  four  supra-subgeneric  taxa  he  termed  ‘groups’,  each  group  named  for  a representative  species. 
Doderlein  (1920,  p.  223)  presented  a phylogenetic  interpretation  of  Luidia  in  which  he  suggested  that  the 
Clathrata  Group  was  primitive  and  gave  rise  to  both  the  Quinaria  and  Alternata  Groups;  the  Ciliaris  Group 
was  in  turn  derived  from  the  Quinaria  Group. 

In  external  appearance,  species  of  the  four  groups  are  superficially  similar  to  one  another,  whereas  the 
ossicle  morphology  of  the  members  of  the  Ciliaris  Group  is  quite  distinct  from  that  of  the  other  three  (Blake 
1973);  I suspect  that  if  the  external  morphology  of  the  Ciliaris  Group  were  as  distinct  as  the  skeletal  morphology, 
this  group  would  be  recognized  as  a separate  genus. 

In  his  discussions  Fell  placed  little  emphasis  on  the  species  available  to  him  for  study,  but  in  an  illustration 
of  abactinals  (19626,  p.  15)  and  an  arm  tip  in  cross-section  (1963a,  p.  386)  he  does  note  that  his  drawings 
were  based  on  L.  neozelanica.  This  species  had  not  been  described  at  the  time  of  Doderlein  (1920),  but  it  has 
since  been  referred  to  the  Ciliaris  Group  (Clark  1953;  Fell  1963a,  p.  433).  Further,  Fell’s  (1963a,  p.  395) 
drawing  of  the  cross-section  of  a Luidia  arm  shows  a crescentic  inferomarginal  outline  typical  of  the  Ciliaris 
Group,  rather  than  the  step-shape  typical  of  the  marginals  of  the  other  three  groups  (PI.  20,  figs.  3,  5,  7). 
Madsen  (1966)  assigned  Platasterias  to  Petalaster,  a subgenus  of  the  Clathrata  Group;  although  I recognized 
Platasterias  at  the  subgeneric  level  because  of  its  distinctive  shape,  I agree  with  Madsen’s  inclusion  of  the 
species  in  the  Clathrata  Group. 

Fell  thus  appears  to  have  based  his  comparisons  between  Luidia  and  Platasterias  on  a species  inferred  to 
be  well  removed  in  phylogenetic  position  (Doderlein  1920)  and  distinct  in  ossicle  morphology  (Blake  1973) 
from  those  Luidia  species  which  are  suggested  to  be  closest  to  P.  latiradiata  (Madsen  1966). 

3.  Ambulacral  column  arrangement 

Although  not  in  his  diagnosis,  Spencer  (1951,  p.  88)  noted  that  ‘.  . . these  first  stages  . . .’  (in 
reference  to  somasteroids  in  general)  \ . . show  no  sign  of  an  ambulacral  groove  . . .’.  Spencer  did 
describe  a \ . . shallow  channel  . . .’  (p.  102)  along  the  midradius  of  Archegonaster,  and  ‘.  . . a 
deep  channel  . . .’  (p.  98)  in  Chinianaster.  Spencer  (1951)  cited  his  1914  paper  for  illustration  of 
the  channel.  In  a drawing  of  an  asteroid  arm  in  cross-section,  Spencer  (1914  in  1914-1940)  showed 
that  the  ambulacral  channel  refers  to  the  axial  notch  for  the  radial  water  canal.  The  channel  is 
thus  a structure  distinct  from  the  ambulacral  furrow  or  groove.  Although  the  ambulacral  channel 
is  deep  in  Chinianaster,  the  basins  for  the  tube-feet  are  \ . . placed  almost  in  the  oral  plane’.  If  an 
ambulacral  furrow  is  present,  as  in  asteroids,  the  ambulacrals  are  arched  and  the  basins  for  the 
tube-feet  are  raised  above  the  oral  surface  of  the  animal. 

In  Fell’s  discussions,  whether  or  not  the  ambulacral  furrow  is  at  most  temporarily  erect 
(somasteroids)  or  permanently  erect  (asteroids)  is  essential  to  the  notion  of  somasteroids  and 
Platasterias  as  a somasteroid.  The  change  from  temporarily  to  permanently  erect  was  selected  as  the 
point  at  which  the  pre-asteroid  phase  gave  way  to  the  asteroid  phase  (1963a,  p.  391). 

Fell  described  a well-developed  ‘lateral  wing’  on  the  ambulacral  of  Platasterias.  This  ‘wing’ 
provides  attachment  for  the  musculature  extending  to  the  first  virgal  (i.e.  the  adambulacral)  that 
permits  temporary  erection  of  the  ambulacral  ossicle  to  form  an  inverted  V-shaped  groove  considered 
homologus  with  the  asteroid  furrow.  The  ossicle  arrangement  in  Platasterias  was  considered  to  be 
asteroid-like  and  leading  toward  the  asteroid  grade  of  organization  (19626,  p.  10).  Although 
recognizing  that  in  P.  latiradiata,  the  wing  extends  over  the  adambulacral  in  the  manner  found  in 
asteroids,  Fell  (1963a,  p.  393)  considered  ‘.  . . the  major  [i.e.  transverse]  axis  of  the  ambulacral 


BLAKE:  SEA  STAR  PHYLOGENY 


177 


lies  almost  horizontally,  in  the  same  axis  as  the  metapinnule  which  it  bears,  as  in  the  Villebrun- 
asteridae  . . In  asteroids,  the  adambulacrals  are  below  the  abradial  end  of  the  ambulacrals  and 
the  basins  for  the  tube-feet  permanently  raised  above  the  substrate. 

The  channel  for  the  water  canal  in  Platasterias  was  considered  partially  enclosed  by  the  ambulacral 
ossicles,  much  as  in  the  Palaeozoic  fossils. 

Discussion.  Interpretation  of  the  ambulacral  ossicles  is  important  to  the  idea  of  a temporarily 
erectable  furrow  (PI.  21,  figs.  1,  2;  text-fig.  3).  In  Platasterias , the  impression  of  a linear,  nearly 
horizontal  ambulacral  ossicle  is  conveyed  by  the  broad  ‘lateral  wing’  of  the  ambulacral,  but 
important  to  the  nature  of  the  furrow  is  the  angle  between  the  axis  of  the  ambulacral/adambulacral 
articulation  and  the  axis  of  the  adradial  portion  of  the  ossicle.  The  ambulacral/adambulacral 
articulation  axis  is  approximately  horizontal  in  the  relaxed  living  sea  star  (although  capable  of 
broad  adjustment;  see  Heddle  1967)  and  therefore  the  angle  the  adradial  end  of  the  ossicle  makes 
to  the  articulation  axis  reflects  the  relaxed,  permanent  furrow  development.  From  a number  of 
medial  arm  ossicles  of  mature  specimens  that  I measured  on  a binocular  microscope  stage,  this 
angle  is  approximately  20°  in  L.  clathrata,  30°  in  P.  latiradiata.  The  size  of  the  angle  depends  on 
the  precise  points  selected  for  measurement,  but  the  critical  idea  is  that  in  Platasterias,  as  in  Luidia, 
these  two  axes  are  not  parallel;  a permanent  furrow  is  present  in  both.  Fell’s  sketches  (1963a, 
p.  395),  although  presumably  not  intended  to  be  precise,  do  accurately  reflect  approximate 
relationships,  and  the  presence  of  a furrow  in  both. 

In  a figure  description.  Fell  (19626,  p.  13)  suggested  that  the  outer,  oral  surfaces  of  the 
adambulacrals,  as  seen  in  Platasterias,  are  erected  into  the  furrow  at  the  asteroid  grade.  Although 
these  faces  are  pulled  toward  one  another  as  any  sea  star  closes  the  furrow,  the  surfaces  are  not 
erect  in  the  furrow  in  asteroids  and  remain  directed  toward  the  substrate  in  the  relaxed  animal. 
As  is  true  of  other  surface  ossicles,  the  adambulacrals  bear  spines  on  their  surficial  faces.  Both 
Platasterias  and  other  sea  stars  have  approximately  vertically  oriented  adradial  side  faces  on  the 
adambulacrals. 

Equally  important  to  ambulacral  ossicle  orientation  is  the  nature  of  the  ambulacral/adambulacral 
articulation  structures,  for,  as  noted  above,  the  same  facets  and  muscle  depressions,  arranged  in 
approximately  the  same  proportions,  can  be  recognized  in  both  Platasterias  and  Luidia  (PI.  20, 
figs.  4,  6)  (Blake  1972,  1973),  as  well  as  in  other  asteroids.  Both  sea  stars  must  be  capable  of 
approximately  the  same  movements;  Heddle  (1967)  described  how  this  musculature  and  articulation 
can  be  used  in  locomotion  and  digging,  and  in  a different  genus  I (Blake  1981)  have  argued  that 
these  structures  can  be  used  in  righting,  interpretations  that  seem  appropriate  for  Platasterias  as 
well  as  Luidia.  The  broad  lateral  wing,  as  suggested  by  Madsen  (1966)  for  the  inferomarginals  and 
adambulacrals,  is  a part  of  the  morphology  of  Platasterias  adapted  to  its  habitat  of  a shifting, 
sandy  bottom. 

As  noted  by  Fell  (19626,  p.  7),  the  radial  water  canal  in  Platasterias  occupies  a channel  along 
the  arm  axis  (PI.  21,  fig.  1)  much  as  in  certain  somasteroids.  A similar  channel,  however,  occurs 
in  L.  clathrata  (PI.  21,  fig.  2)  as  well  as  in  other  asteroids  (e.g.  Asterias),  hence  the  feature  is  not 
of  taxonomic  value  among  living  sea  stars. 

4.  Growth  gradients 

Spencer  (1951,  p.  91)  described  the  interambulacrals,  or  virgals,  as  arranged  in  linear  series  at  an 
angle  to  the  ambulacral  row  with  a single  series  arising  at  the  abradial  edge  of  each  ambulacral. 
The  development  of  virgals  provided  the  basis  for  the  families  recognized  by  Spencer:  in  the 
Chinianasteridae,  ossicles  occupy  the  entire  oral  surface  apart  from  the  ambulacrals  and  mouth;  in 
the  Archegonasteridae,  virgals  occur  only  distally  on  the  arms,  and  virgals  are  lacking  from  the 
Archophiactinidae. 

Fell  (1963a)  developed  his  ideas  of  growth  gradients  in  asterozoans  about  this  transverse  alignment 
of  oral  surface  ossicles.  Growth  gradients  were  considered  to  be  parallel  or  weakly  convergent 
lines  along  which  morphologic  structures  were  aligned.  Growth  gradients  were  envisioned  as 


178 


PALAEONTOLOGY,  VOLUME  25 


arranged  in  two  series,  one  set  parallel  to  the  arm  radius  and  a second,  lateral  set  subperpendicular 
to  the  first.  The  ambulacral  ossicles,  nerve  ganglia,  and  radial  water  vessel  were  aligned  along  the 
main  longitudinal  gradient,  the  adambulacrals  along  the  first  lateral  gradient,  and  so  on. 

Each  row  of  virgals  was  termed  a metapinnule,  and  each  represents  a transverse  gradient.  The 
influence  of  either  longitudinal  or  transverse  gradients  could  be  stronger;  where  transverse  gradients 
dominated,  metapinnule  ossicles  were  aligned,  whereas  these  ossicles  were  offset  transversely  but 
aligned  longitudinally  as  the  longitudinal  gradients  became  dominant.  The  differentation  of  the 
typical  asteroid  oral  surface  ossicles  from  the  morphologically  similar  virgals  was  associated  with 
the  inferred  declining  influence  of  transverse  gradients:  the  adambulacral  from  virgal  1,  the 
superambulacral  from  virgal  2,  the  marginal  from  virgal  3,  and  the  marginal  radiole,  from  virgal  4 
(Fell,  1963a,  pp.  392,  405). 

Transverse  gradients  dominate  in  primitive  somasteroids  (e.g.  Chinianaster)  but  gradually  lose 
their  influence;  longitudinal  gradients  dominated  in  Astropecten  and  phylogenetically  later  asteroids. 
The  transverse  gradients  were  ultimately  derived  from  the  inferred  crinoid  ancestor,  hence  the 
metapinnule  from  the  crinoid  pinnule. 

Members  of  ambulacral  pairs  apparently  alternate  across  the  arm  axis  in  certain  fossil 
somasteroids  (Spencer  1951,  p.  102)  much  as  the  brachials  alternate  in  biserial  crinoids.  This 
similarity  of  arrangement  contributed  to  Fell’s  view  of  a close  relationship  between  crinoids  and 
stelleroids  (Fell  1963a,  p.  415).  In  Platasterias  the  transverse  gradients  were  considered  clearly 
recognizable  (PI.  21,  fig.  5),  although  the  ossicles  are  equally  clearly  equated  with  their  inferred 
homologues  in  asteroids. 

Ossicle  alignment  was  the  sole  criterion  provided  for  recognition  of  gradient  type.  Most 
post-Palaeozoic  sea  stars  have  few  to  many  ossicles  on  the  oral  surface  between  the  adambulacral 
and  inferomarginal  series:  these  are  the  so-called  actinal  intermediates  of  Fell’s  terminology,  or 
more  simply,  actinals  (text-fig.  3).  Actinals  typically  form  a more  or  less  tightly  packed  mosaic, 
they  are  similar  to  one  another  in  shape,  and  although  a broad  size  range  can  be  present,  size 
changes  are  gradational  across  the  surface.  The  ossicles  thus  are  perceived  as  forming  parallel  rows 
of  a single  orientation,  or,  frequently,  intersecting  rows,  one  of  which  is  oriented  more  or  less 
parallel  to  the  arm  axis,  the  other  radiating  obliquely  from  the  furrow  toward  the  animal  margin. 
Dominant  orientation  of  rows  can  be  consistent  within  taxa;  Hotchkiss  and  Clark  (1976)  used  the 
arrangement  of  these  ossicles  to  separate  the  Asteropseidae  from  the  Poraniidae. 


EXPLANATION  OF  PLATE  21 

Figs.  1,  4,  6.  Luidia  clathrata  (Say).  1,  transverse  view  of  arm,  ossicles  identified  in  text-fig.  2;  see  note  with 
text-fig.  2,  x 3.  4,  inclined  aboral  view  of  arm  showing  alignment  of  lateral  abactinals  with  superomarginal 
row  (arrow);  alignment  is  lost  in  midarm  area;  some  spinelets  are  still  in  place,  x 3.  6,  oral  surface  of  arm 
showing  alignment  of  inferomarginals  (arrows),  actinals,  and  adambulacrals.  Tube-feet  are  visible  along  arm 
axis,  x 3. 

Figs.  2,  3,  5.  L.  ( Platasterias ) latiradiata  (Gray).  2,  transverse  view  of  arm,  ossicles  identified  in  text-fig.  2, 
x 4.  3,  aboral  view  of  arm  showing  alignment  of  lateral  abactinals  with  superomarginal  (arrow)  row; 
alignment  is  lost  toward  midarm  region,  x 3.  5,  oral  surface  of  arm  showing  alignment  of  ossicles, 
enlarged  marginal  radioles;  arrow  indicates  inferomarginals,  x 3. 

Fig.  7.  Asterias  sp.  Aboral  view  of  interior  of  oral  surface  of  arm  showing  alignment  of  ossicles  in  radial 
series,  see  discussion  in  text;  arrows  indicate  ambulacrals,  x 3. 

Fig.  8.  Archaster  typicus  Muller  and  Troschel.  Oral  surface  of  arm  showing  well-defined  fascioles  (arrow) 
between  inferomarginals.  Pits  in  some  adambulacrals  are  for  pedicellariae.  Note  similarity  of  lateral  spines 
with  those  of  L.  (P.)  latiradiata  and  Astropecten  regalis  (PI.  20,  fig.  2),  x 3. 

Fig.  9.  Dermasterias  imbricata  (Grube).  Oral  surface  of  arm  showing  ambulacral  spines  (arrows)  crossing  over 
furrow  axis,  in  function  forming  furrow  cover  plates,  x 3. 


PLATE  21 


mi 

7 


BLAKE,  living  sea  stars 


180 


PALAEONTOLOGY,  VOLUME  25 


Certain  authors  have  interpreted  the  oblique  arrangement  of  actinals  in  true  asteroids  as  reflecting 
transverse  growth  gradients  sensu  Fell,  but  Fell  himself  equivocated  on  the  origins  and  significance 
of  these  ossicles.  In  reference  to  the  astropectinid  and  later  stages  of  evolution.  Fell  (1963a,  p.  389) 
said  ‘Actinal  intermediate  plates  are  usually  present,  and  are  arranged  in  longitudinal  rows;  they 
are  sometimes  also  arranged  in  oblique  series,  but  these  latter  are  unrelated  to  the  transverse 
gradients  which  produce  the  ambulacrals  and  adambulacrals.’  The  inferred  phylogenetic  history 
of  these  ossicles  was  described  as  well  (Fell  1963a,  p.  391):  ‘Actinal  intermediate  plates  are  lacking 
from  division  1 (i.e.  Chinianasteridae,  see  Table  1,  herein),  appear  as  minute  and  irregular  rudiments 
in  division  2 (i.e.  Platasteriidae),  persist  as  somewhat  larger  and  more  irregular  elements  in  the 
luidiids,  become  progressively  more  conspicuous  in  various  genera  of  the  Astropectinidae,  and 
extremely  conspicuous  in  post-astropectinid  groups.’  In  suggesting  that  actinals  were  absent  at  the 
chinianasterid  grade  but  appeared  at  the  platasteriid  grade,  Fell  implied  that  these  ossicles  are 
derived  but  appeared  very  early.  In  addition.  Fell  (19636,  p.  144)  stated  that  the  virgals  in  Platasterias 
had  stabilized  at  four  elements  (identified  above)  not  including  actinals. 

Fell  (1963a,  p.  387),  however,  also  stated,  ‘Ossicles  are  differentiated  along  the  transverse  gradients 
in  the  following  sequence,  from  within  outwards:  adambulacral,  superambulacral  (if  present), 
actinal  intermediate  plates,  marginals.’  Elsewhere,  Fell  (1962a,  p.  635)  referred  to  actinals  as 
‘accessory  virgalia’;  in  the  figure  description  (Fell  19626,  fig.  4a),  the  actinal  was  labeled  ‘intercalary 
virgalium  (actinal  intermediate)’  and  again  (19636,  abstract)  the  metapinnule  was  referred  to  as 
differentiated  into  the  actinal  intermediate,  as  well  as  the  other  ossicle  types.  Although  he  apparently 
was  uncertain  as  to  how  to  treat  these  ossicles  in  Platasterias  and  Luidia,  Fell’s  strongest  statements 
seem  to  be  that  the  actinals  were  secondary,  and  not  derived  from  virgals. 

In  order  to  understand  Fell’s  interpretation  of  Platasterias  and  the  nature  of  the  Somasteroidea,  it  is  necessary 
to  consider  the  several  steps  Fell  (1963a)  used  to  select  the  groups  of  extant  sea  stars  he  inferred  to  be 
primitive,  and  the  phylogenetic  sequence  selected  to  connect  fossil  somasteroids  to  living  species.  Among 
living  sea  stars,  Fell  limited  his  search  for  the  primitive  group  to  the  Luidiidae,  Astropectinidae,  and 
Porcellanasteridae  because  ‘It  has  been  generally  agreed  that  the  more  primitive  extant  asteroids  are  the  three 
families  in  which  the  tube-feet  lack  suckers’  (1963a,  p.  285).  Of  the  three,  the  porcellanasterids  were  tentatively 
eliminated  because  they  lack  superambulacrals,  one  of  the  ossicles  inferred  to  be  homologous  with  the  virgals 
of  somasteroids. 

More  important  to  Fell’s  (1963a,  p.  385)  phylogenetic  ideas  was  the  concept  of  growth  gradients.  Among 
living  sea  stars,  transverse  rows  are  most  clearly  defined  in  the  Luidiidae,  less  so  (because  the  two  systems 
of  gradients  were  seen  as  changing  influence  gradually)  in  the  Astropectinidae  and  Porcellanasteridae.  As 
Fell  had  recognized  the  strongly  developed  transverse  gradients  in  the  lower  Paleozoic  somasteroids,  their 
development  in  the  Luidiidae  permitted  recognition  of  this  group  as  primitive  among  living  families.  The 
gradients  are  weaker  but  present  in  the  Astropectinidae  (Fell,  1963c,  p.  467),  which  was  then  next  in  the 
phylogenetic  sequence.  The  Porcellanasteridae  is  specialized  in  a number  of  features,  and  was  therefore  third 
in  the  sequence. 

Isolation  of  Platasterias  as  a somasteroid  began  with  the  realization  that  this  genus  ‘.  . . exhibits  growth 
gradients  identical  with  those  of  Ordovician  somasteroids’  (Fell,  1963a,  p.  383).  Fell  then  obtained  specimens 
of  Platasterias  for  dissections;  this  work  led  to  the  assignment  of  Platasterias  to  the  Somasteroidea,  as  well 
as  to  the  refinement  of  a phylogenetic  sequence  extending  from  somasteroids  through  several  steps  of  asteroid 
evolution  (see  Table  1).  The  phylogenetic  sequence  was  based  on  a variety  of  characters  and  emphasized 
morphology  of  the  fossils  but  not  their  stratigraphic  position  because  the  fossils  appeared  through  an  interval 
of  strata  inferred  to  be  too  brief  to  permit  recognition  of  phylogenetic  sequence  based  on  stratigraphic  sequence 
(1963a,  p.  385).  Fell  (1967,  p.  580)  pointed  out,  however,  that  his  phylogeny  is  consistent  with  what  is  known 
of  the  stratigraphic  sequence. 

In  summary,  gradient  recognition  depends  upon  ossicle  alignment.  A continuing,  gradual  decline 
in  transverse  gradient  influence  and  a concomitant  increase  in  the  influence  of  logitudinal  gradients 
was  suggested  to  have  occurred  during  somasteroid/asteroid  phylogeny. 

Discussion.  As  pointed  out  by  Fell,  the  ambulacrals,  adambulacrals,  superambulacrals,  and 
marginals  are  aligned  in  both  Platasterias  and  Luidia  (PI.  21,  figs.  5,  6).  Although  Luidia  typically 
has  a cluster  of  larger  spines  at  the  abradial  ends  of  the  marginals  (PI.  22,  figs.  1,  9)  rather  than 


BLAKE:  SEA  STAR  PHYLOGENY 


18 


a single  larger  radiole,  as  in  Platasterias  (PI.  21,  fig.  5),  the  positions  of  the  spines  on  the  marginals 
are  the  same.  Although  only  a single  large  spine  occurs  on  each  Platasterias  marginal,  numerous 
smaller  spines  are  present,  and  further,  number  of  spines  apparently  is  not  important,  as  is  shown 
by  the  variation  found  among  Luidia  species.  Presence  of  a single  marginal  radiole  in  the  petaloid 
species  Astropecten  regalis  (PI.  20,  fig.  2)  argues  for  the  convergent  evolution  of  marginal  spine 
reduction.  Because  there  is  no  difference  in  ossicle  alignment  between  Platasterias  and  Luidia,  this 
character  cannot  be  used  for  taxonomic  separation  (although  ossicle  proportions  do  differ). 

A more  serious  problem  arises  from  the  arrangement  of  the  actinals  and  abactinals.  These  ossicles 
in  Luidia  and  Platasterias  are  relatively  stout,  and  the  actinals  and  the  lateral  two  to  four  rows  of 
abactinals  in  Platasterias  and  the  Clathrata  Group  species  of  Luidia  are  aligned  with  the  four 
ossicles  (Fell,  1963a,  pp.  392,  405)  of  the  transverse  series  (PI.  20,  figs.  1,  9-12;  PI.  21,  figs.  3-6). 
Abactinal  alignment  with  the  marginals  is  weaker  toward  the  middle  of  the  abactinal  surface  and 
among  the  other  Luidia  groups  ( sensu  Doderlein  1920)  in  general.  A single  row  of  actinals  is 
present  in  most  species  of  Luidia,  including  members  of  the  apparently  primitive  Clathrata  Group 
(Doderlein  1920),  but  extra  series  are  present  among  the  Alternata  Group. 

In  some  manner,  the  arrangement  of  the  actinals  and  abactinals  must  be  accommodated  with 
ideas  on  transverse  gradients  because  the  alignment  satisfies  the  criterion  of  transverse  gradient 
recognition.  Fell  (1963a,  p.  415)  recognized  the  problem  presented  by  the  abactinals  and  noted 
that  these  ossicles  were  not  aligned  in  the  primitive  Chinianasteridae.  He  suggested  (p.  416)  that 
in  Platasterias  the  gradients  would  seem  to  be  carried  into  the  soft  tissues  in  their  path,  and  that 
caused  \ . . the  paxillae  to  differentiate  as  if  they  were  virgalia’.  The  alignment  was  thus  considered 
secondary.  Fell  also  (1963a,  p.  389)  pointed  out  that  actinal  intermediates  (i.e.  actinals)  are  aligned 
in  transverse  series  in  various  astropectinid  and  phylogenetically  {sensu  Fell  1963a)  later  sea  stars. 

For  discussion  purposes,  I have  accepted  the  following  hypotheses  from  Spencer’s  and  Fell’s 
papers:  (1)  that  asterozoans  were  derived  from  a crinoid  ancestry,  so  that  the  primitive  asterozoan 
had  radial  rows  of  virgals;  (2)  that  somasteroids  are  the  primitive  asterozoans  and  gave  rise  to  the 
asteroids;  (3)  that  abactinals  were  present  in  the  primitive  somasteroid  (because  Spencer  1951, 
described  them  in  the  earliest  known  taxa).  If  (1)  or  (2),  as  now  understood,  were  rejected,  then 
the  phylogenetic  hypothesis  using  growth  gradients  could  not  be  maintained.  If  (3)  were  rejected, 
then  the  fossil  record  is  misleading  and  where  it  is  misleading  can  be  interpreted  only  using  external 
criteria. 

Beginning  with  these  assumptions,  abactinal  alignment  seen  in  Platasterias  and  Luidia  can  be 
considered  either  primitive  and  present  in  the  first  somasteroid,  or  derived  and  appearing  after  the 
first  somasteroid  grade.  The  actinals  of  Platasterias  and  Luidia  can  be  considered  either  as  primitive 
and  homologous  with  virgals,  or  not.  If  primitive,  then  their  alignment  must  be  primitive,  i.e.  they 
are  a part  of  the  primitive  transverse  gradient.  If  secondary,  their  alignment  must  be  secondary 
as  well  (i.e.  their  alignment  could  not  be  part  of  a primitive  transverse  gradient  if  the  ossicles 
themselves  did  not  occur  in  the  primitive  species).  Where  I treat  the  actinals  as  primitive,  I will 
treat  the  specific  actinal  arrangement  seen  in  Platasterias  and  Luidia  (i.e.  a single  actinal  row)  as 
secondary  because  of  the  distribution  of  these  ossicles  in  the  accepted  (see  above)  primitive 
somasteroid. 

First,  if  we  accept  both  abactinal  arrangement  and  actinal  ossicles  as  seen  in  Luidia  as  primitive,  then  the 
fossil  record  is  incomplete  and  misleading.  The  abactinal  pattern  reported  in  somasteroids,  including 
Chinianaster,  the  inferred  primitive  genus,  is  an  open  meshwork  of  light  paxilliform  ossicles  (Spencer  1951, 
p.  91).  If  a hypothesis  of  declining  influence  of  transverse  gradients  is  to  be  followed,  then  the  abactinal 
arrangement  seen  in  Platasterias  and  Luidia  must  have  been  derived  from  some  unknown  pre-chinianasterid 
somasteroid,  and  retained  while  other  characters  (including  ossicle  differentiation)  were  evolving  toward  the 
asteroid  grade.  This  means  that  known  fossil  somasteroids  had  to  be  off  the  main  line  of  somasteroid  evolution 
because  their  abactinal  fields  had  already  attained  a derived,  open  meshwork  state  (text-fig.  1a;  PI.  21,  fig.  3), 
a condition  not  reached  in  surviving  asteroid  evolution  ( sensu  Fell)  until  after  the  level  of  the  Paxillosida. 
Following  the  implications  further  because  the  abactinals  of  Luidia  and  Platasterias  are  sturdy,  the  primitive 
abactinal  condition  in  somasteroids  likely  was  sturdy  as  well,  and  the  relatively  fragile  ossicles  of  known 


.82 


PALAEONTOLOGY,  VOLUME  25 


fossils  represent  derived  states.  Thus,  taxa  with  the  relatively  sturdy  arrangement  are  unknown  from  the  fossil 
record,  whereas  the  relatively  fragile  ones  were  preserved.  Finally,  the  fossil  record  itself  becomes  unreliable 
in  that  although  most  oral  surface  ossicle  distributions  in  known  fossils  are  considered  to  represent  primitive 
conditions,  the  abactinal  arrangement  is  considered  secondary  for  reasons  external  to  the  fossils  themselves. 

As  to  the  actinal  surface  ossicles,  their  reduction  sequence  in  known  somasteroids  began  with  a full  oral 
surface  field  (Chinianasteridae),  continued  through  an  intermediate  step  in  which  ossicles  were  present  only 
distally  on  the  arms  (Archegonasteridae),  and  ended  with  complete  elimination  of  these  ossicles  (Archophi- 
actinidae)  (Spencer  1951).  Thus  the  presence  in  Platasterias  and  most  Luidia  of  a single  row  extending  the 
entire  arm  length  must  represent  a lineage  separate  from  known  fossils  because  they  match  neither  any  known 
fossils,  nor  the  known  reduction  sequence. 

The  second  approach  is  to  assume,  as  Fell  did,  that  abactinal  alignment  and  actinal  ossicles  are  secondary. 
This  path,  however,  also  yields  difficulties  because  it  requires  extension  of  the  transverse  gradients  on  to  the 
aboral  surface  at  the  time  the  influence  of  these  gradients  is  hypothesized  to  be  declining  and  the  longitudinal 
gradient  influence  increasing.  Because  the  row  of  actinals  also  becomes  aligned  with  adjacent  adambulacrals 
and  marginals,  we  must  hypothesize  alternating  rows  of  increasing  and  decreasing  influence  of  transverse 
gradients  (i.e.  adambulacrals,  decreasing;  actinals,  increasing;  marginals  and  marginal  radioles,  decreasing; 
abactinals,  increasing).  Even  if  we  were  to  argue  that  abactinals  represent  an  ossicle  system  separate  from 
crinoids  and  not  derived  from  them,  and  therefore  not  primitive  in  somasteroids,  the  problem  of  alignment 
is  not  resolved.  Abactinals  do  occur  in  an  open  meshwork  pattern  in  the  known  primitive  somasteroid 
(Chinianaster),  and  Platasterias  was  considered  to  be  approximately  at  the  Villebrunaster  level  of  organization, 
at  a post -Chinianaster  pre-Archegonaster  somasteroid  grade.  Transverse  gradients  are  hypothesized  to  be 
declining  in  influence  during  this  sequence,  yet  in  Platasterias  these  gradients  had  to  be  extending  their 
influence  on  to  a new  area  of  the  body. 

Further,  because  transverse  alignment  can  be  secondary  among  some  ossicle  systems,  it  might  also  be 
secondary  among  all  others.  Platasterias  therefore  does  not  readily  fit  into  a somasteroid/asteroid  phylogeny 
based  on  growth  gradients.  This  does  not  challenge  the  hypothesis  as  applied  only  to  known  fossils. 

Primitive  and  derived  states  could  be  rearranged  so  that  abactinal  alignment  is  considered  primitive  and 
actinals  derived,  or  vice  versa,  but  this  only  combines  inherent  difficulties  in  different  ways. 

It  would  be  possible  to  argue  the  large  actinal  field  was  not  lost  in  the  connecting  links  between  somasteroids, 
luidiids,  and  remaining  asteroids,  but  this  alternative  removes  Platasterias  and  Luidia  from  the  main  line  of 
asteroid  evolution,  for  there  is  no  known  platasteriid  or  luidiid,  fossil  or  living,  with  the  appropriate  morphology 
(i.e.  well-defined  fascioles,  and  a large  actinal  field).  Tethyaster  is  an  astropectinid  that  comes  quite  close  to 
this  morphology,  and  indeed  Jangoux  (1975)  argued  that  this  genus  is  in  many  ways  intermediate  between 
Luidia  and  Astropecten.  The  marginal  fascioles  in  Tethyaster,  however,  are  relatively  small,  and  the  development 
of  fascioles  also  was  important  in  Fell’s  hypothesis,  serving  as  feeding  and  respiration  channels. 

In  any  event  beginning  with  the  assumptions  of  Spencer  (1951)  and  Fell  (1962a  et  seq.),  and 
assuming  either  that  abactinal  alignment  and  actinal  occurrence  is  primitive  or  secondary,  known 
fossil  and  modern  taxa  permit  no  application  of  the  hypothesis  of  growth  gradients  to  somasteroid/ 
asteroid  phylogeny  that  does  not  require  ignoring  important  aspects  or  morphology.  In  essence, 
alignment  is  too  extensive  in  Platasterias,  more  extensive  than  in  the  fossil  somasteroids,  and  this 
calls  for  explanation  other  than  phylogenetic  inheritance  from  a crinoid  ancestry.  Such  an  origin 
is  suggested  below,  under  a functional  explanation  for  alignment  in  Luidia. 

As  noted  earlier  here,  individual  ambulacral  column  and  marginal  ossicles  of  Platasterias  are 
transversely  more  elongate  than  are  those  of  the  fossil  somasteroids.  Again,  Platasterias  is  showing 
an  inferred  primitive  trait  more  strongly  than  it  is  seen  in  the  earliest  known  somasteroids,  and 
therefore  an  explanation  other  than  inheritance  of  a primitive  condition  is  demanded. 

5.  Other  morphologic  characters 

The  nature  of  the  mouth  frame.  In  modern  sea  stars,  all  ambulacral  pairs  are  united  across  the 
furrow  by  muscles  and  articular  structures.  In  the  somasteroids,  Spencer  argued  that  the  proximal 
ambulacrals  (three  pair  in  Chinianaster ) were  not  linked  across  the  furrow,  resulting  in  broad, 
V-shaped  openings  termed  buccal  slits.  These  slits,  however,  are  likely  to  have  resulted  from 
post-mortem  events.  Spencer  (1951,  p.  100)  argued  that  ambulacrals  of  certain  small  Chinianaster 
were  pulled  apart  after  death,  and  Fell  (1963a,  p.  403)  suggested  that  the  buccal  slits  of  Spencer 


BLAKE:  SEA  STAR  PHYLOGENY 


183 


resulted  from  such  changes.  The  ossicles  of  the  buccal  slits  in  Archegonaster  apparently  show  no 
significant,  distinctive  morphologic  characters  (Spencer,  1951,  fig.  9)  as  might  be  expected  of  ossicles 
that  form  the  margin  of  an  opening  rather  than  an  articulated  double  column.  Minor  differences 
are  illustrated  in  these  ossicles,  but  near-oral  ambulacral  column  ossicles  also  are  somewhat 
distinctive  in  certain  living  sea  stars  (e.g.  the  Astropectinidae). 

Arrangement  of  spinelets,  along  the  fasciolar  margins.  In  Platasterias , a series  of  small  spinelets, 
termed  cover  plates  by  Fell,  line  the  edges  of  the  adoral  surfaces  of  the  marginals  and  adambulacrals. 
These  spinelets  were  described  as  being  held  in  a web  and,  when  depressed,  they  cover  the  fasciolar 
grooves  between  subsequent  metapinnules  or  transverse  gradient  series  (1963a,  p.  397).  The 
arrangement  was  inferred  derived  from  the  crinoid  ancestor,  but  differs  in  closing  over  the 
interpinnular  groove  rather  than  over  the  pinnule  surface,  as  in  crinoids.  The  cover  plates  in 
Platasterias  were  seen  as  protecting  the  fasciolar  or  feeding  grooves. 

A similar  arrangement  is  found  in  Luidia,  Astropecten,  and  other  sea  stars,  although  the  ossicles 
in  these  taxa  typically  are  more  slender,  i.e.  spine-like,  and  much  more  numerous  than  in  Platasterias. 
As  in  Platasterias , these  ossicles  can  be  depressed  over  the  furrows.  I was  unable  to  recognize  true 
webbing  in  preserved  specimens  of  Luidia  (although  such  a covering  is  present  in  Goniopecten, 
suborder  Cribellina,  order  Paxillosida),  but  a mucus  covering  is  present  that  in  dried  specimens 
available  to  me  can  extend  to  the  tips  as  well  as  between  adjacent  spinelets.  The  dense,  overlapping 
arrangement  of  many  spinelets  provides  an  effective  cover  for  the  fasciolar  furrow.  Various  sea 
stars,  e.g.  Dermasterias  imbricata  (PI.  21,  fig.  9),  are  capable  of  pulling  the  sides  of  the  furrow  (i.e. 
the  adambulacrals)  together  and  arching  the  furrow  spines  over  the  furrow.  Functionally,  these 
are  cover  plates  and  such  plates  are  widely  distributed  among  the  Asteroidea. 

The  arrangement  found  in  Platasterias  could  be  readily  derived  from  that  of  Luidia  by  reduction 
of  spine  number  and  minor  changes  of  spine  shape.  Arrangement  of  fasciolar  spinelets  therefore 
provides  little  justification  for  significant  taxonomic  separation  of  Luidia  and  Platasterias. 

Super ambulacrals.  Fell  (1963a)  argued  that  the  superambulacrals  of  Platasterias  and  Luidia  are 
occluded  second  virgals.  This  explanation  of  origins  is  plausible,  however,  superambulacrals  are 
unknown  in  fossil  somasteroids  but  they  are  present  throughout  the  Luidiidae  and  Astropectinidae. 
This  distribution  can  only  serve  to  isolate  the  fossil  from  the  extant  sea  stars,  not  link  them  in  a 
single  phylogenetic  sequence. 

Soft  parts.  Platasterias  has  a blind  gut  with  caeca  extending  into  the  arms,  an  arrangement  Fell 
(1963a,  p.  396)  thought  probably  existed  in  fossil  somasteroids.  In  addition,  Platasterias  has  small, 
simple  non-suctorial  tube-feet  considered  similar  to  those  in  Chinianaster  (Fell  1962c,  p.  474)  and 
small,  double  internal  ampullae  also  inferred  to  be  similar  to  those  in  Ordovician  somasteroids. 
Luidia,  like  Platasterias,  has  simple,  non-suctorial  tube-feet,  double  ampullae,  and  a blind  gut  with 
caeca  extending  into  the  arms. 

6.  Feeding  habits 

Spencer  (1951,  p.  87)  emphasized  feeding  behaviour  in  his  tripartite  division  of  stelleroids:  The 
grouping  of  starfish  adopted  here  is  based  upon  the  activities  of  the  arms,  especially  during  feeding.’ 
The  asteroids  were  seen  as  carnivores  on  larger  organisms,  primitive  ophiuroids  were  believed 
adapted  for  feeding  on  small  particles  on  or  in  the  bottom,  and  the  somasteroids  inferred  to  have 
lived  on  planktic  particles  (in  the  suggested  relatively  primitive  genus,  Villebrunaster)  or  particles 
from  the  surface  sediment  layers  (in  the  suggested  advanced  somasteroid  Archegonaster). 

Fell  also  stressed  observed  and  inferred  feeding  behaviour  in  his  phylogenetic  sequence.  He 
(1962a,  p.  14)  suggested  two  types  of  feeding  are  present  in  Platasterias,  ‘microphagous  ciliary 
feeding’,  and  ‘selective  detrital  feeding’.  In  the  former,  particles  of  food  were  believed  to  be  conveyed 
in  water  currents  along  the  fascioles  to  the  arm  radius,  then  proximally  toward  the  mouth.  In 
specimens  of  Platasterias  available  to  him.  Fell  observed  amphipods  in  the  mouth  and  in  a food 
groove;  he  suggested  that  the  amphipods  were  collected  by  the  tube-feet  and  passed  to  the  mouth. 


184 


PALAEONTOLOGY,  VOLUME  25 


He  inferred  that  this  would  be  the  limit  of  carnivorous  feeding,  because  of  the  relatively  small 
mouth  of  Platasterias.  A specimen  of  Platasterias  in  the  U.S.N.M.  collections  has  a foraminiferan, 
small  snail,  and  arthropod  fragments  in  the  mouth  area;  the  presence  of  this  small  prey  does  not 
appear  fortuitous. 

As  observed  by  Fell  (1963a),  Astropecten  and  especially  Luidia  (PI.  21,  fig.  6)  are  living  species 
with  quite  similar  fasciolar  furrows  extending  between  the  marginals  and  ambulacrals,  yet  these 
are  voracious  predators  of  other  echinoderms,  molluscs,  and  arthropods.  Although  suspension 
feeding  has  been  suggested  for  both  (Fenchel  1965,  for  Luidia  sarsi;  Gemmill  1915  and  Gislen  1924 
for  Astropecten),  Feder  and  Christensen  (1966)  doubt  this  behaviour  occurs  in  these  genera.  The 
furrows  in  both,  as  Fell  (1963a,  p.  391)  observed  for  the  Luidiidae,  are  probably  respiratory  in 
nature.  In  contrast,  taxonomically  diverse  genera  in  which  suspension  feeding  has  been  observed 
do  not  possess  fasciolar  furrows  (e.g.  Henricia,  Spinulosida,  Rassmussen  1965,  ossicles  are  aligned 
in  this  genus,  yielding  unobstructed  channels  between  ossicle  rows;  Oreaster,  Valvatida,  Halpern 
quoted  in  Anderson  1978;  unidentified  Brisingidae,  Forcipulatida,  Pawson  1978).  Fell  recognized 
the  differences  in  function  between  the  fascioles  of  the  modern  sea  stars  and  that  inferred  for  the 
somasteroids  and  assumed  the  function  changed  with  the  evolution  of  the  asteroid  grade. 

Suspension  feeding  is  difficult  to  document  without  direct  observation.  Gislen  (1924)  found 
surface  ciliary  currents  to  be  widespread  in  sea  stars  and  therefore  the  potential  for  ciliary  feeding 
as  well,  although  the  currents  flow  away  from  the  oral  area  in  some  species.  Gislen  (1924),  however, 
considered  the  functions  of  the  currents  to  be  largely  for  respiration  and  cleaning,  although  he  did 
observe  some  capture  of  particles. 

The  size  of  the  mouth  does  not  provide  a useful  guide  to  potential  prey  size  limits,  as  suggested 
by  Fell  for  Platasterias  (1962 b,  p.  14),  for  both  Astropecten  and  typical  Luidia  species,  although 


EXPLANATION  OF  PLATE  22 

Fig.  1.  Luidia  alternata  (Say).  USNM  7528.  Inclined  view  of  disc  region  of  specimen  not  distorted  by  food; 
specimen  in  alcohol,  x 1. 

Figs.  2,  3.  Luidia  alternata  (Say).  USNM  7528.  2,  inclined  aboral,  x and  3,  oral  view,  x 1,  of  individual 
containing  unbroken  corona  of  Lytechnius  varegatus  (Lamarck).  The  peristome  of  the  echinoid  (arrow, 
pointing  to  gill  slit)  is  centred  on  that  of  the  sea  star.  The  sea  star  major  radius  is  140  mm,  relaxed  minor 
radius  20  mm,  and  height  1 5 mm.  The  echinoid  corona  diameter  about  50  mm,  height  25  mm.  All  dimensions 
approximate.  Spines  and  organic  materials  are  missing  from  the  echinoid  but  pieces  of  the  Aristotle’s  Lantern 
are  present.  The  echinoid  thus  apparently  was  dead  or  nearly  so  at  time  of  ingestion  and  it  had  been 
consumed  and  presumably  was  soon  to  be  expelled  at  time  of  collection.  In  alcohol. 

Figs.  4,  5.  Luidia  clathrata  (Say).  USNM  E17599.  4,  view  into  disc;  and  5,  lateral  view  of  fragmentary 
individual,  radius  about  100  mm,  containing  a specimen  of  Moira  atropos  (Lamarck),  length  about  40  mm. 
The  echinoid  is  now  incomplete  (arrows  point  to  remaining  coronal  plates)  but  the  aboral  surface  of  the 
sea  star  retains  the  form  of  the  sea  urchin,  hence  the  echinoid  was  complete  when  ingested.  Fine  lines  beyond 
the  echinoid  plates  are  its  spines;  the  echinoid  apparently  was  ingested  alive,  or  very  recently  dead;  specimen 
is  dry,  x 1 . 

Fig.  6.  Luidia  clathrata  (Say).  USNM  8442.  Inclined  aboral  view  of  specimen,  radius  about  1 10  mm,  containing 
Moira  atropos,  length  about  40  mm.  Spines  are  retained  on  prey,  x 1. 

Figs.  7,  8.  Luidia  clathrata  (Say).  USNM  E3268.  7,  aboral  view;  and  8,  oral  view  of  sea  star  containing  a 
broken  but  not  distorted  Mellita  quinquiesperforatus  (Leske).  The  spines  are  missing  from  the  echinoid. 
Sea  star  radius  about  150  mm,  that  of  the  echinoid  is  about  50  mm.  Aboral  ossicles  of  the  sea  star  are 
distorted  from  life  position,  beyond  the  edge  of  the  sea  star  in  a relaxed  state  (arrow);  the  aboral  surface 
of  the  sea  star  was  partially  opened  after  collection.  Compare  colour  marking  with  that  of  L.  (P.) 
latiradiata,  PI.  20,  fig.  1;  specimen  is  dry,  x 

Fig.  9.  Luidia  clathrata  (Say).  Oral  view  of  mouth  area  of  specimen  not  distorted  by  food;  specimen  is  dry, 
x 1 . Luidia  can  consume  particles  of  dimensions  many  times  greater  than  the  2 or  3 mm  mouth  diameter, 
and  of  volumes  greater  than  that  of  the  relaxed  disc.  Much  caution  is  needed  in  interpreting  possible  food 
particle  size  in  fossil  organisms  with  flexible  bodies. 


PLATE  22 


1 


BLAKE,  sea  star  Luidia 


PALAEONTOLOGY,  VOLUME  25 


effective  predators,  also  have  relatively  small  oral  frames.  The  lack  of  importance  of  mouth  size 
is  clearly  seen  in  a series  of  Luidia  specimens  in  the  U.S.N.M.  collections  (PI.  22).  In  each  example 
illustrated,  the  size  of  the  ingested  echinoid  exceeds  that  of  the  relaxed  diameter  of  the  disc. 
Peristome  size,  measured  in  related  Luidia  specimens  of  comparable  sizes,  is  2 to  3 mm;  oral  spines 
mounted  around  the  mouth  of  the  sea  star  overlap  over  the  peristomial  opening.  Occurrence  of 
large  prey  in  a number  of  specimens  clearly  demonstrates  the  behaviour  is  not  unusual.  Astropecten, 
too,  is  capable  of  feeding  on  relatively  large  prey  organisms  (Clark  1962,  pi.  xiii). 

Thus,  Luidia  has  fasciolar  grooves  similar  to  those  of  both  Platasterias  and  presumably  of  the 
ancient  somasteroids,  but  there  is  no  clear  evidence  that  either  Platasterias  or  Luidia  make  use  of 
suspension  feeding  and  therefore  the  presence  of  fascioles  in  Palaeozoic  somasteroids  provides  no 
morphologic  evidence  of  suspension  feeding  in  these  organisms.  The  presence  of  very  large  prey 
in  Luidia  means  that  mouth  and  body  size  provide  no  clear  guide  to  food  particle  size  in  echinoderms 
with  a loosely  articulated  skeleton,  including  the  Palaeozoic  somasteroids. 

Summary 

Platasterias  is  very  close  to  typical  species  of  Luidia  in  all  characters  except  those  related  to  the 
transverse  elongation  of  ossicles,  and  similar  to  Palaeozoic  somasteroids  only  in  ways  it  is  also 
similar  to  Luidia.  Greater  differences  in  ossicle  morphology,  except  for  this  transverse  elongation, 
are  seen  between  species  of  the  Ciliaris  Group  of  Luidia  {sensu  Doderlein  1920)  and  Clathrata  Group 
( sensu  Doderlein  1920)  than  are  seen  between  Platasterias  and  the  Clathrata  Group  of  Doderlein. 
I therefore  consider  (Blake  1972)  Platasterias  to  be  a subgenus  of  Luidia  assignable  to  the  Clathrata 
Group,  and  not  a somasteroid  as  suggested  by  Fell.  As  noted  above,  Madsen  (1966)  carried  this 
interpretation  one  step  further,  for  he  did  not  believe  Platasterias  warranted  subgeneric  recognition. 
Summary  comparison  among  sea  star  taxa  is  provided  in  Table  2,  with  Porania  included  as  an 
example  of  an  asteroid  well  removed  from  the  Luidiidae. 

If  Luidia  were  to  be  extensively  subdivided,  perhaps  along  the  lines  suggested  by  Doderlein 
(1920),  then  Platasterias  could  be  recognized  at  the  generic  level,  but  on  a morphologic  basis,  it  still 
would  have  to  be  included  in  the  same  family  as  other  Luidia  species.  Luidia,  as  noted  by  Fell 
(1963a),  clearly  is  a member  of  the  Asteroidea. 

THE  NATURE  OF  LUIDIA 


The  affinities  of  the  Luidiidae 

In  his  work.  Fell  assumed  the  sea  stars  with  nonsuctorial  tube-feet,  and  in  particular  the 
astropectinids  and  luidiids  are  primitive  among  extant  sea  stars.  I agree  with  this  interpretation 
but,  although  consensus  might  lean  toward  the  families  cited,  no  general  agreement  has  ever  been 
reached. 

Mortensen  (1922),  noting  that  few  authors  had  commented  directly  on  the  question  of  primitive 
position,  queried  the  then-active  sea  star  workers  as  to  their  opinions.  W.  K.  Fisher,  H.  L.  Clark, 
and  R.  Koehler  all  considered  the  astropectinids  to  be  primitive,  as  did  Mortensen  himself,  whereas 
Doderlein  thought  the  asterinids  were  primitive.  Earlier,  Perrier  (1884)  argued  the  forcipulates 
were  primitive  on  inferred  directions  of  pedicellariae  evolution. 

Most  active  sea  star  taxonomists  list  the  Paxillosida  first  in  their  faunal  lists,  seemingly  thereby 
implying  an  inferred  primitive  position,  but  as  A.  M.  Clark  pointed  out  to  me  (pers.  comm.),  this 
is  simply  because  they  are  following  Fisher  (1911). 

In  this  diagnosis  of  the  Platyasterida  (including  the  Palasteriscidae  and  Luidiidae),  the  only 
character  listed  by  Fell  (1963a,  p.  392)  was  the  development  of  transverse  growth  gradients.  Spencer 
and  Wright  (1966)  expanded  upon  the  diagnosis,  but  most  features  listed  by  these  authors  also 
apply  to  members  of  the  Paxillosida.  They  did  include  the  presence  of  a single  row  of  marginals 
in  the  Platyasterida.  Although  relatively  small  in  Luidia  compared  to  those  in,  for  example, 
Astropecten,  superomarginals  are  present  in  Luidia  (PI.  20,  figs.  9,  10),  recognizable  on  the  basis 
of  position  of  origin  at  the  terminal  (see  criteria  listed  by  Blake  1978),  and  by  size  among  primitive 


BLAKE:  SEA  STAR  PHYLOGENY 


187 


( sensu  Doderlein  1920)  Luidia  species.  This  leaves  only  the  transverse  gradients  to  unite  the  Luidiidae 
with  the  Palaeozoic  Palasteriscidae,  but  this  character,  as  discussed  above,  appears  unreliable. 
Based  on  illustrations  provided  by  Spencer  (1919  in  1914-1940),  however,  Luidia  and  the 
Palasteriscidae  are  distinct  in  ossicle  morphology,  and  fascioles  apparently  are  lacking  in  the  fossil 
family.  In  addition,  they  are  separated  by  a Devonian  to  Miocene  interval. 

The  Luidiidae,  however  is  similar  to  the  Astropectinidae  in  many  soft  and  hard  part  characters, 
as  summarized  by  Fisher  (1911),  Fell  (1963a),  and  other  workers;  included  here  are  the  nature  of 
the  paxillae  and  other  ossicles,  and  the  presence  of  non-suctorial  tube-feet.  Following  McKnight 
(1977),  I therefore  have  herein  returned  the  Luidiidae  to  the  Paxillosida. 

A partial  revised  classification  of  the  stellate  echinoderms  based  on  Spencer  and  Wright  (1966), 
McKnight  (1975)  and  the  arguments  presented  here  is  provided  below.  For  comparative  purposes, 
the  equivalent  classification  of  Spencer  and  Wright  (1966)  is  also  included. 


Revised  classification: 

Class  Somasteroidea 
Order  Goniactinida 

Family  Chinianasteridae 
Family  Villebrunasteridae 
Family  Archegonasteridae 
Family  Archophiactinidae 
Family  Helianthasteridae 
Class  Asteroidea 
Order  Platyasterida 
Family  Palasteriscidae 
Order  Paxillosida 

Suborder  Flemizonina 
Suborder  Diplozonina 

Family  Luidiidae  (includes  Luidia 
( Platasterias ) ladradiata) 
Family  Astropectinidae 
Family  Porcellanasteridae 
Suborder  Cribellina 
Class  Ophiuroidea 


Spencer  and  Wright  (1966): 

Class  Stelleroidea 
Subclass  Somasteroidea 
Order  Goniactinida 
Family  Chinianasteridae 
Family  Villebrunasteridae 
Family  Platasteriidae  (includes  Platasterias 
ladradiata) 

Family  Archegonasteridae 
Family  Archophiactinidae 
Subclass  Asteroidea 
Order  Platyasterida 
Family  Palasteriscidae 
Family  Luidiidae 
Order  Paxillosida 

Suborder  Hemizonina 
Suborder  Diplozonina 
Family  Astropectinidae 
Family  Porcellanasteridae 
Suborder  Cribellina 
Subclass  Ophiuroidea 


The  gap  in  the  fossil  record  of  the  Platyasterida  sensu  Fell  1963a 

The  fossil  record  does  not  support  an  inference  that  the  lineage  leading  to  Luidia  has  endured 
nearly  unchanged  from  the  Palaeozoic.  The  order  Platyasterida  sensu  Fell  (and  Spencer  and  Wright 
1966)  includes  three  genera  assigned  to  two  families:  Platanaster  (M.  Ord.)  and  Palasteriscus  (L. 
Dev.),  both  belonging  to  the  Palasteriscidae,  and  Luidia  (Mio.-Rec.),  belonging  to  the  Luidiidae. 
There  is  thus  a gap  from  Lower  Devonian  to  Miocene  in  the  inferred  record  of  the  order.  At  first 
inspection,  this  appears  trivial  because  of  the  sketchy  record  of  sea  stars,  but  I believe  it  becomes 
more  important  when  this  record  is  carefully  considered. 

Among  post-Palaeozoic  sea  stars,  Astropecten  and  its  close  allies  have  a relatively  good  fossil 
record.  This  record  seems  to  result  from  body  structure,  habitat,  and  habits.  The  sea  stars  have 
relatively  stout  marginal  frames  and  at  least  Astropecten  frequently  lives  in  shallow  waters  on 
unconsolidated  substrates,  and  burrows  beneath  the  surface.  The  relative  abundance  of  Astropecten 
as  fossils  does  not  necessarily  result  from  a greater,  enduring  abundance  (although  they  are  very 
common  today),  but  from  where  and  how  individuals  live.  Luidia  lives  in  similar  environments, 
and  it  too  is  a burrower;  although  the  marginal  frame  is  not  as  stout  as  in  Astropecten,  the 
inferomarginals  of  many  species  are  comparable  to  those  in  Astropecten,  and  sturdier  than  those 
of  the  Asteriidae,  a family  known  from  the  Jurassic  and  whose  representatives  are  also  sometimes 
found  in  similar  habitats. 


PALAEONTOLOGY,  VOLUME  25 


table  2.  Comparison  of  somasteroids  and  asteroids.  Criteria  available  do  not  permit  separation  of 
Platasterias  as  a somasteroid.  Porania  (Poraniidae)  is  taxonomically  distant  from  the  somasteroids 
but  still  has  many  features  in  common  with  them.  Data  in  part  from  Fell  (1963a). 


ambulacra  1 

ossicle  alignment  furrow 

super- 

ambulacral 

terminal 

tube  feet 

somasteroids 

primarily  transverse  in  not  erect 
some,  others  longitudinal 

absent 

present  or 
absent 

? 

L.  (Platasterias)  both  transverse  and  erect 

longitudinal 

present 

present 

non-suctorial 

Luidia  clathrata  both  transverse  and  erect 

longitudinal 

present 

present 

non-suctorial 

Porania 

primarily  longitudinal  erect 

absent 

present 

suctorial 

other  asteroids 

primarily  longitudinal  erect 

present 

or 

absent 

present 

non-suctorial 
and  suctorial 

anus 

body  shape 

feeding  habits 

somasteroids 

? 

petaloid  to  large 
oral  disc 

inferred  suspension  or 
small  particle  bottom 
feeder 

L.  (Platasterias) 

absent 

petaloid 

carnivore 

Luidia  clathrata 
Porania 

absent 

present 

small  disc,  strap- 
shape  arms 
large  oral  disc 

carnivore 

suspension  feeding, other 

other  asteroids 

present 

or 

absent 

varied 

varied 

Luidia  is  a common  genus,  with  sixty  or  more  species  (many  described  since  Doderlein  1920) 
widely  distributed  in  temperate  to  tropical  seas;  further,  individuals  commonly  are  abundant. 
Individuals  of  most  species  of  Luidia  are  relatively  large  and  would  not  be  readily  overlooked  in 
the  fossil  record.  Luidia  superba  specimens  have  been  reported  from  the  Galapagos  Islands  with 
major  radii  up  to  415  mm  and  an  arm  breadth  at  the  disc  edge  of  60  mm  (Downey  and  Wellington 
1978).  The  absence  of  a fossil  record  for  platyasterids  ( sensu  Fell)  from  Devonian  to  Miocene 
certainly  is  not  conclusive  evidence  that  they  were  not  present,  but  because  structurally  comparable 
sea  stars  known  from  similar  environments  do  have  long  records,  the  absence  of  Luidia  or  its 
relatives  requires  explanation. 

A functional  explanation  for  the  transverse  alignment  of  ossicles  in  Luidia 

Fell  argued  that  the  body  plan  of  Luidia  is  the  direct  phylogenetic  heritage  of  Cambrian  crinoids; 
in  effect,  that  functional  explanations  of  the  morphology  of  these  contemporary,  mobile  carnivores, 
living  with  their  oral  surfaces  toward  the  substrate,  are  to  be  sought  in  ancient,  attached 
suspension-feeding  organisms  living  with  their  oral  surfaces  directed  into  the  water  column.  1 
believe  an  explanation  for  the  Luidia  body  plan  can  be  developed  that  is  more  closely  linked  to 
the  life  needs  of  sea  stars. 


BLAKE:  SEA  STAR  PHYLOGENY 


189 


The  arms  of  Luidia  are  very  flexible  in  part  because  they  are  narrow  and  low  and  in  part  because 
the  alignment  of  ossicles  yields  essentially  a segmented  pattern  in  which  each  radial  row  of  ossicles 
is  relatively  weakly  connected  to  adjacent  rows.  This  arrangement  weakens  toward  the  middle  of 
the  aboral  surface  of  the  arm,  where  the  abactinals  are  too  small  to  interfere  with  flexibility. 

The  value  of  flexibility  to  Luidia  can  be  correlated  with  preferred  food  and  habitat.  Luidia  is  a 
predator  on  larger  solitary  organisms-molluscs,  arthropods,  and  other  echinoderms— in  which 
long,  flexible  arms  are  useful  for  the  manipulation  of  prey.  Luidia  is  commonly  an  inhabitant  of 
shallow,  often  agitated,  bottoms  and,  further,  it  frequently  burrows  into  the  substrate.  Arm  flexibility 
is  useful  for  burrowing  (Heddle  1967)  and  also  for  righting  (Blake  1981). 

L.  ( P .)  latiradiata  shares  the  ossicle  arrangement  of  traditional  Luidia  species,  and  it,  too,  is  an 
active  predator  (see  above),  but  in  it  the  demands  of  environment  (Madsen  1966)  might  have 
dominated  the  need  for  maximum  arm  flexibility. 

A superficially  similar  transverse  alignment  is  seen  in  Asterias  and  other  asteriids  in  the  rows 
of  arm  ossicles  immediately  lateral  to  the  furrow  columns  (PI.  21,  fig.  7).  The  ancestry  of  the 
Asteriidae  is  unknown,  and  therefore  it  is  not  known  when  this  alignment  evolved,  but  asteriids, 
like  Luidia,  are  active  predators  with  a need  for  relatively  flexible  arms. 

Another  hypothesis  for  ossicle  alignment  in  Luidia  and  Astropecten  can  be  based  on  the  burrowing 
habits  of  these  sea  stars  combined  with  the  respiratory  current  flow  described  by  Gislen  (1924). 
Although  surface  currents  appear  typical  of  sea  stars,  deep  fascioles  seemingly  are  restricted  to 
burrowers.  These  channels,  combined  with  their  cover  spinelets,  should  provide  protection  from 
sediment  interference  for  the  current  flow.  Alignment  of  ossicles  and  their  intervening  channels  in 
turn  would  permit  a more  efficient  water  flow  than  that  possible  in  a species  with  an  irregular 
arrangement  of  ossicles.  Surface-dwelling  species  presumably  would  suffer  less  from  current 
disruption  and,  in  them,  furrows  generally  are  lacking. 

Archaster  is  an  extant  genus  superficially  very  similar  to  Astropecten  in  both  form  and  habit.  It 
too  has  relatively  deep  fascioles  (PI.  21,  fig.  8),  but  only  between  the  infero-,  and  not  the 
superomarginals.  Archaster,  however,  is  an  atypical  member  of  the  Valvatida  rather  than  the 
Paxillosida,  as  is  Astropecten.  It  is  possible  that  Archaster  was  derived  directly  from  an 
Astropecten- like  source  but  the  skeletal  morphology  in  this  genus  does  not  seem  to  support  such 
an  idea;  convergence  resulting  from  habitat  similarities  is  a preferable  hypothesis. 

Thus,  both  the  alignment  of  ossicles  in  the  arm  and  the  grooves  between  the  aligned  ossicles 
seem  subject  to  convergent  evolution;  use  in  any  phylogenetic  scheme  must  be  made  with  great  care. 

SUMMARY 

Platasterias  latiradiata  Gray  is  removed  from  the  Somasteroidea  and  assigned  at  subgeneric  rank 
to  the  genus  Luidia  of  the  monogeneric  asteroid  family  Luidiidae;  Luidia  ( Platasterias ) therefore 
should  not  be  singled  out  as  a model  for  the  reconstruction  of  the  biology  of  early  stellate 
echinoderms.  Reasons  for  the  transfer  of  L.  ( Platasterias ) are:  (1)  The  single  distinctive  feature  of 
the  Palaeozoic  somasteroids  seems  to  be  the  absence  of  a permanent  ambulacral  furrow.  In  contrast 
the  adambulacral/ambulacral  arrangement  in  L.  (P.)  latiradiata  is  the  same  as  that  found  in  L. 
clathrata  and  essentially  as  in  all  other  living  asteroids;  a true  furrow  is  present  in  L.  (P.)  latiradiata. 

(2)  Individual  ossicle  morphology  and  ossicle  and  muscle  arrangement  of  L.  ( Platasterias ) are 
easily  within  a reasonable  range  of  variation  for  Luidia.  Although  many  (but  not  all)  ossicles  of 
L.  (P.)  latiradiata  are  proportionately  broad,  they  are  morphologically  closer  to  the  Clathrata 
Group  species  of  Luidia  than  these  ossicles  are  in  turn  to  the  Ciliaris  Group  species  of  Luidia. 

(3)  A reconstruction  of  phylogenetic  history  based  on  ossicle  alignment  (growth  gradients)  requires 
important  reversals  of  direction  of  evolution  in  order  to  enable  fossil  somasteroids  and  extant  L. 

( Platasterias ) and  other  luidiids  to  fit  the  hypothesis.  (4)  Two  important  ossicle  types,  the 
odontophore  and  superambulacral,  are  present  in  L.  ( Platasterias ) and  other  extant  sea  stars,  but 
they  apparently  are  absent  from  the  fossil  somasteroids.  (5)  Although  L.  ( Platasterias ) has  a petaloid 
arm  shape  suggestive  of  that  of  certain  crinoids,  weakly  developed  petaloid  arms  are  present  in 


190 


PALAEONTOLOGY,  VOLUME  25 


other  taxonomically  widely  separated  sea  star  taxa;  the  character  is  subject  to  convergence.  (6) 
Known  food  habits  of  L.  ( Platasterias ) suggest  similarities  to  those  of  typical  Luidia  species  and 
not  to  those  habits  inferred  by  Spencer  (1951)  for  the  fossil  somasteroids.  (7)  Available  data  suggests 
soft-part  morphology  of  L.  ( Platasterias ) is  essentially  that  of  a typical  Luidia. 

The  Luidiidae  is  transferred  from  the  otherwise  Paleozoic  order  Platyasterida  to  the  common 
post-Palaeozoic  Paxillosida. 

The  origin  of  the  transverse  alignment  of  ossicles  in  Luidia  (including  Platasterias ) is  ascribed 
to  habits  and  habitats  of  the  sea  star.  This  ossicle  arrangement  enhances  arm  flexibility,  useful  to 
an  active,  burrowing  predator  living  in  shallow,  often  turbulent,  environments.  The  presence  of 
deep  furrows  between  aligned  ossicle  series  also  provides  an  open  channel,  unimpeded  by  sediment, 
for  the  flow  of  respiratory  water  currents. 

Acknowledgements.  I thank  Ailsa  M.  Clark,  Maureen  E.  Downey,  and  David  L.  Pawson  for  very  helpful 
reviews  of  an  earlier  version  of  the  manuscript;  Porter  M.  Kier  for  identification  of  the  echinoids;  Richard 
L.  Turner  for  discussion  of  certain  points;  the  authorities  of  the  U.S.  National  Museum  for  loan  of  specimens; 
and  H.  Barraclough  Fell  and  the  authorities  of  the  Royal  Society  for  permission  to  reproduce  certain  drawings. 


REFERENCES 

algor,  J.  r.  1971.  Structure  of  the  ambulacral  system  in  some  modern  and  early  Paleozoic  starfish.  Abstr. 
Prog.  Geol.  Soc.  Am.,  Wash.  D.C.  3,  489-490. 

anderson,  j.  m.  1978.  Studies  on  functional  morphology  in  the  digestive  system  of  Oreaster  reticulatus  (L.) 

(Asteroidea).  Biol.  Bull.  mar.  biol.  Lab.,  Woods  Hole.  154,  1-14. 
blake,  D.  B.  1967.  Skeletal  elements  in  asteroids.  Ab.  1967,  Spec.  Pap.  Geol.  Soc.  Am.  115,  15-16. 

— 1972.  Sea  star  Platasterias : ossicle  morphology  and  taxonomic  position.  Science,  N.Y.  176,  306-307. 

— 1973.  Ossicle  morphology  of  some  recent  asteroids  and  description  of  some  west  American  fossil  asteroids. 
Univ.  Calif.  Pubis,  geol.  Sci.  104,  60  pp. 

1978.  The  taxonomic  position  of  the  modern  sea-star  Cistina  Gray,  1840.  Proc.  biol.  Soc.  Wash.  91, 
234-241. 

— 1981.  The  new  Jurassic  sea  star  genus  Eokainaster  and  comments  on  life  habits  and  the  origins  of  the 
modern  Asteroidea.  J.  Paleont.  55,  33-46. 

Clark,  A.  m.  1953.  Notes  on  asteroids  in  the  British  Museum  (Natural  History)  III.  Luidia.  Bull.  Br.  Mus. 
nat.  Hist.,  zool.  1,  379-412. 

— 1962.  Starfishes  and  their  relations.  Brit.  Mus.  (nat.  Hist.),  Pub.  377.  118  pp. 
doderlein,  l.  1920.  Die  Asteriden  der  Siboga-Expedition  II:  Die  Gattung  Luidia  und  ihre  Stammesgeschichte. 
Siboga-Expeditie  Mono.  46b,  193-294. 

downey,  m.  e.  and  Wellington,  G.  m.  1978.  Rediscovery  of  the  giant  sea  star  Luidia  superba  A.  H.  Clark  in 
the  Galapagos  Islands.  Bull.  mar.  Sci.  28,  375-376. 

feder,  h.  m.  and  Christensen,  a.  m.  1966.  Aspects  of  asteroid  biology.  In  boolootian,  r.  a.  (ed.).  Physiology 
of  echinodermata.  Interscience  Pub.,  John  Wiley  & Sons,  New  York.  822  pp. 
fell,  h.  b.  1948.  Echinoderm  embryology  and  the  origin  of  chordates.  Biol.  Rev.  23,  81-107. 

— 1962a.  A surviving  somasteroid  from  the  eastern  Pacific  Ocean.  Science,  N.Y.  136,  633-636. 

— 19626.  A living  somasteroid,  Platasterias  latiradiata  Gray.  Paleont.  Contr.  Univ.  Kans.  Artie.  6,  Echino- 
dermata. 16  pp. 

— 1963a.  The  phylogeny  of  sea-stars.  Phil.  Trans.  R.  Soc.,  Lond.,  ser.  B 246,  381-435. 

— 19636.  A new  family  and  genus  of  Somasteroidea.  Trans.  R.  Soc.  N.Z.  3,  143-146. 

— 1963c.  The  evolution  of  the  echinoderms.  Rep.  Smithson.  Instn.  4559,  457-490. 

— 1965.  Reply  to  ancestry  of  sea-stars  by  G.  M.  Philip.  Nature,  208,  768-769. 

— 1967.  Echinoderm  ontogeny.  In  moore,  r.  c.  (ed.).  Treatise  on  invertebrate  paleontology  pt.  S, 
Echinodermata  1,  1,  S1-S296;  2,  S297-S650.  [Lawrence,  Kansas,  and  Geol.  Soc.  Am.] 
fenchel,  t.  1965.  Feeding  biology  of  the  sea  star  Luidia  sarsi  Diiben  and  Koren.  Ophelia,  2,  223-236. 
fisher,  w.  k.  1911.  Asteroidea  of  the  North  Pacific  and  adjacent  waters.  Bull.  U.S.  natn.  Mus.  76.  420  pp. 
gemmill,  J.  F.  1915.  On  the  ciliation  of  asterids,  and  on  the  question  of  ciliary  nutrition  in  certain  species. 
Proc.  zool.  Soc.  Lond.  1915,  1-19. 


BLAKE:  SEA  STAR  PHYLOGENY 


191 


gislen,  t.  1924.  Echinoderm  studies.  Zool.  Bidr.  Upps.  9,  1-316. 

heddle,  d.  1967.  Versatility  of  movement  and  the  origins  of  asteroids.  In  millott,  n.  Echinoderm  Biology. 
Symp.  zool.  Soc.  Lond.  20.  240  pp. 

hotchkiss,  f.  h.  c.  1977.  Ophiuroid  Ophiocanops  (Echinodermata)  not  a living  fossil.  J.  nat.  Hist.  11,  377-380. 
and  clark,  A.  M.  1976.  Restriction  of  the  family  Poraniidea  sensu  Spencer  and  Wright,  1966  (Echino- 
dermata: Asteroidea).  Bull.  Br.  Mus.  nat.  Hist.,  zool.  30,  263-268. 
hyman,  L.  H.  1955.  The  invertebrates:  Echinodermata.  McGraw-Hill,  New  York.  764  pp. 
jangoux,  M.  1975.  Note  on  the  genus  Tethyaster  Sladen.  Rev.  zool.  Afr.  89,  761-768. 
mcknight,  d.  G.  1975.  Classification  of  somasteroids  and  asteroids  (Asterozoa:  Echinodermata).  J.  Roy.  Soc. 
N.Z.  5,  13-19. 

— 1977.  Classification  of  recent  paxillosid  sea-stars  (Asterozoa:  Echinodermata).  N.Z.  Ocean.  Inst.  Rec.  3, 
113-119. 

madsen,  F.  j.  1966.  The  Recent  sea-star  Platasterias  and  the  fossil  Somasteroidea.  Nature,  209,  1367. 
mortensen,  t.  1922.  Echinoderm  larvae  and  their  bearing  on  classification.  Ibid.  110,  806. 

— 1931.  Contributions  to  the  study  of  the  development  and  larval  forms  of  echinoderms.  K.  danske 
Vidensk.  Selsk.  Skr.,  natur.  math.  9th  ser.,  4,  (1).  32  pp. 
pawson,  d.  l.  1978.  Some  aspects  of  the  biology  of  deep-sea  echinoderms.  Thalassia  jugosl.  12,  287-293. 
pearse,  j.  s.  1969.  Treatise  on  invertebrate  paleontology.  Part  S:  Echinodermata  1 (a  review).  Q.  Rev.  Biol. 
44,  298. 

perrier,  e.  1884.  Memoire  sur  les  Etoiles  de  mer  recueillies  dans  la  mer  des  Antilles  et  le  golfe  du  Mexique 
par  M.  Agassiz.  Nouv.  Archs  Mus.  Hist,  nat.,  ser.  2,  6,  127-276. 
philip,  g.  m.  1965.  Ancestry  of  sea-stars.  Nature  208,  766-768. 

Rasmussen,  H.  w.  1965.  Asteroids  du  Tertiaire  inferieur  de  Libye  (Afrique  du  Nord).  Ann.  Paleont.  Invert. 
52,  1-15. 

spencer,  w.  k.  1914-1940.  British  Palaeozoic  Asterozoa.  Palaeontogr.  Soc.  ( Monogr .)  540  pp. 

— 1951.  Early  Palaeozoic  starfish.  Phil.  Trans.  R.  Soc.,  Lond.,  ser.  B 235,  87-129. 

— and  wright,  c.  w.  1966.  Asterozoans.  In  moore,  r.  c.  (ed.).  Treatise  on  invertebrate  paleontology, 
pt.  U,  Echinodermata  3,  1,  Ul-U366a;  2,  U367-U696.  Lawrence,  Kansas,  and  Geol.  Soc.  Am.  Press. 

D.  B.  BLAKE 

Department  of  Geology 
University  of  Illinois 
245  Natural  History  Building 
1301  W.  Green  St. 
Urbana,  Illinois  61801 
U.S.A. 


Typescript  received  30  September  1980 
Revised  typescript  received  10  January  1981 


A LOWER  CARBONIFEROUS  AlSTOPOD 
AMPHIBIAN  FROM  SCOTLAND 

by  CARL  F.  WELLSTEAD 


Abstract.  "Ophiderpeton' , a mid-Visean  ai'stopod  from  the  Wardie  Shales  near  Edinburgh  is  described. 
Although  ‘ Ophiderpeton ’ possesses  many  of  the  attributes  characteristic  of  ai'stopods  and  seems  generally  more 
closely  comparable  to  ophiderpetontids  than  to  phlegethontiids,  details  of  cranial  anatomy  such  as  possession  of 
a relatively  short  skull,  short  parietals,  and  the  absence  of  a tabular-parietal  contact,  as  well  as  the  presence  of 
spinal-nerve  foramina  in  only  a portion  of  the  vertebral  column  and  the  absence  of  tetra-radiate  ribs, 
postcranially,  distinguishes  ‘ Ophiderpeton ’ from  both  of  the  currently  recognized  ai'stopod  families.  On  the  basis 
of  these  differences,  ‘ Ophiderpeton ’ is  renamed,  Lethiscus  stocki,  and  a new  family,  Lethiscidae,  is  erected  to  hold 
the  new  species. 

Of  all  Palaeozoic  amphibians,  the  ai'stopods  are  the  most  specialized  and  phylogenetically  isolated. 
The  genera  known  from  the  Upper  Carboniferous  are  totally  without  limbs  and  have  up  to  230  trunk 
vertebrae.  The  vertebrae  are  simple  cylinders,  without  a trace  of  trunk  intercentra  or  caudal  haemal 
arches.  The  skull  has  lost  much  of  the  dermal  cover  common  to  labyrinthodonts  and  rhipidistians. 
These  snake-like  ai'stopod  genera  have  been  classified  among  the  lepospondyls  along  with  nectrideans 
and  microsaurs,  sharing  with  them  such  features  as  holospondylous  vertebrae,  lack  of  an  otic  notch, 
absence  of  labyrinthine  infolding  of  teeth,  and  absence  of  palatine  teeth.  It  is  not  certain  that  these 
features  indicate  common  ancestry. 

In  addition  to  the  familiar  Upper  Carboniferous  genera,  a single  specimen  from  the  Lower 
Carboniferous,  originally  identified  by  Thomas  Stock  (1882)  as  an  ai'stopod,  has  lain  undescribed  for 
nearly  a century.  The  latest  discussion  of  this  specimen  was  that  of  Baird  ( 1 964),  who  accepted  Stock’s 
identification,  and  cited  a number  of  broad  similarities  it  held  with  the  Upper  Carboniferous  genera. 
Baird  also  mentioned  the  difficulty  in  studying  the  specimen  which  had  contributed  to  its  neglect.  To 
gain  more  information  about  the  early  differentiation  of  Palaeozoic  amphibians,  a further  attempt 
has  been  made  to  study  the  specimen. 

The  specimen  (text-fig.  1)  is  49  cm  long  and  is  preserved  in  a very  elongate  nodule,  which  was 
originally  broken  into  dozens  of  pieces,  each  of  which  was  split  through  the  middle  to  give  a series  of 
sections  through  the  skull  and  vertebrae.  No  attempt  was  made  by  Stock  to  prepare  the  specimen 
further.  Preliminary  efforts  to  remove  the  matrix  from  the  bone,  mechanically  and  chemically,  have 
not  been  successful,  although  X-ray  photography  and  tomography  have  revealed  some  detail  of  the 
skull  and  postcranial  skeleton  otherwise  concealed  in  matrix. 

Presently,  enough  detail  can  be  seen  in  the  sections  to  give  a general  description  of  the  animal  and  to 
provide  sufficient  information  about  its  anatomy  to  allow  discussion  of  its  relationships  to  other 
lepospondyls.  The  only  further  ‘preparation’  has  been  to  clean  glue  and  a century’s  accumulation  of 
grime  from  the  specimen.  This  cleaning  revealed  with  startling  clarity  vertebrae,  ribs,  scales,  and 
sections  through  the  skull.  The  bone  is  preserved,  but  is  outlined  with  a thin  coating  of  pyrite.  The 
neural  and  notochordal  canals  and  large  lacunae  within  the  bones  and  skull  are  filled  with  calcite. 
Details  of  bone  histology  are  evident,  and  the  random  breakage  of  the  nodule  has  produced  a series  of 
i sections  in  many  planes.  Information  provided  by  these  sections  would  be  lost  were  the  bone  to  be 
removed  for  production  of  latex  casts  following  Baird’s  technique  (Baird  1955);  therefore,  this 
otherwise  very  effective  method  of  preparing  specimens  was  not  attempted.  Normal  external  views  of 
the  bone  are  rarely  evident,  but  those  available  reveal  sufficient  detail  along  the  column  to  allow 

[Palaeontology,  Vol.  25,  Part  1,  1982,  pp.  193-208.| 


194 


PALAEONTOLOGY,  VOLUME  25 


text-fig.  1.  Lethiscus  stocki,  MCZ  2185.  a,  whole  specimen,  x^.  b,  skull,  x 3. 


WELLSTEAD:  LOWER  CARBONIFEROUS  AMPHIBIAN 


95 


description  of  some  regional  differentiation.  Revealed  in  the  cleaned  specimen  were  several 
characteristics  which  ally  it  with  the  Ai'stopoda,  but  also  others  which  distinguish  it  from  ai'stopod 
families  Phlegethontiidae  and  Ophiderpetontidae. 

MATERIALS  AND  METHODS 

A standard  medical  X-ray  machine  and  the  Stratomatic  tomography  X-ray  machine  with  tri-spiral  movement 
and  0-6  focal  spot  were  used  to  produce  the  X-ray  photographs  critical  to  the  description  of  this  specimen.  The 
film  used  was  Ilford  X-ray  film. 

Text-fig.  3 is  an  X-ray  of  the  half-nodule  containing  the  skull  less  the  posterior  skull  roof.  X-rays  of  the  half- 
nodule bearing  the  posterior  skull  roof  revealed  no  more  than  can  be  seen  with  the  naked  eye.  The  matrix  of  the 
nodule  is  too  dense  for  successful  application  of  X-rays  to  the  assembled  skull-bearing  nodule  halves. 

Abbreviations  used  in  figures 

bo— basioccipital;  bs— basisphenoid;  c— coronoid;  cap— capitulum;  ect — ectopterygoid;  ep— epipterygoid; 
f — frontal;  it — intertemporal;  j — jugal;  1 — lacrimal;  m — maxilla;  mnd — mandible;  n — nasal;  ot — otic  capsules; 
p — parietal;  pf — postfrontal;  pm — premaxilla;  po — postorbital;  poz — postzygapophysis;  pp — postparietal; 
prf — prefrontal;  prz — prezygapophysis;  ps — parasphenoid;  pt — pterygoid;  q — quadrate;  qj — quadratojugal; 
sq— squamosal;  st— supratemporal;  t— tabular;  tub— tuberculum;  v— vomer. 


SYSTEMATIC  PALAEONTOLOGY 

Class  AMPHIBIA 
Sub-claSS  LEPOSPONDYLI 
Order  aistopoda 
Family  lethiscidae  fam.  nov. 

Diagnosis.  Same  as  for  the  only  known  genus. 

Lethiscus  gen.  nov. 

Type  species:  Lethiscus  stocki  sp.  nov. 

Diagnosis.  A small  elongate  amphibian  with  holospondylous  vertebrae  and  short-snouted  skull 
bearing  lateral  temporal  fenestrae.  Orbits  placed  in  the  anterior  third  of  the  skull  and  separated  from 
the  temporal  fenestrae  by  the  postorbital  bones.  The  bones  of  the  prefrontal-postfrontal-postorbital 
series  increase  in  size  posteriorly.  Intertemporal  absent.  Parietals  roughly  equivalent  in  length  to  the 
frontals  and  surrounding  a large  parietal  opening.  Tabulars  do  not  contact  the  parietal  bones. 
Postparietals  relatively  large.  Mandibles  deep  and  long,  approximating  the  length  of  the  skull. 

The  trunk  is  very  long.  Differentiation  along  the  vertebral  column  is  expressed  by  the  presence  in 
the  posterior  portion  of  the  column  of  spinal-nerve  foramina,  serrated  neural  spines  and  transverse 
processes  rising,  in  part,  from  the  centra.  Anteriorly,  transverse  processes  arise  solely  from  the  neural 
arches  and  spinal-nerve  foramina  are  absent.  Ribs  are  bicipital  and  robust. 

Etymology.  The  generic  name  extends  Cope’s  practice  of  naming  serpen  tiform  ‘lepospondyls’  for  rivers  in  Hades. 
In  this  case,  Lethe  is  a stream  named  for  the  Greek  god  of  forgetfulness. 

Lethiscus  stocki  sp.  nov. 

Diagnosis.  The  same  as  for  genus.  Specific  name  honours  the  discoverer  of  the  specimen. 

Holotype.  MCZ  2185.  Museum  of  Comparative  Zoology,  Harvard  University,  Cambridge,  Massachusetts, 
U.S.A.  Skull  and  postcranial  skeleton.  This  is  the  only  known  specimen. 

Locality  and  horizon.  Stock  (1882)  discovered  the  specimen  in  the  shales  of  the  Wardie  shore,  north  of 
Edinburgh,  Scotland  (Wood  1977,  gives  further  locality  data).  These  beds,  the  Wardie  Shales,  lie  in  the  middle  of 


196 


PALAEONTOLOGY,  VOLUME  25 


the  Lower  Oil  Shale  Group  (Mitchell  and  Mykura  1962)  and  have  a mid-Visean  age  (text-fig.  2d).  The  base  of  the 
Arthur’s  Seat  Volcanic  Beds  defines  the  base  of  the  group  in  the  Edinburgh  vicinity,  though  they  cannot  be  traced 
regionally.  Fitch,  Miller,  and  Williams  (1970)  report  a potassium-argon  date  of  347  ± 5 my  B.P.  from  these 
volcanics.  George,  Johnson,  Mitchell,  Prentice,  Ramsbottom,  Sevastopulos,  and  Wilson  (1976)  place  volcanic 
rocks  dated  at  338  + 4 my  B.P.  (Fitch  et  al.  1970)  in  the  lower  portion  of  the  Upper  Oil  Shale  Group.  The  Wardie 
Shales  can,  therefore,  be  estimated  as  approximately  340  million  to  345  million  years  old.  George  et  al.  (1976) 
correlate  the  mid-Visean  with  the  Middle  Mississippian  (Meramec)  of  North  America. 


DESCRIPTION 


Skull 

The  skull  (text-fig.  1)  is  exposed  in  an  irregular  para-frontal  fracture  which  in  one  half-nodule  yields  an 
unobscured  view  of  a portion  of  the  left  suspensorium  and  of  the  ventral  surface  of  the  skull  table  posterior  to  the 
fronto-parietal  suture.  In  the  other  half-nodule  only  extremely  irregular  sections  of  the  remainder  of  the  skull  can 
be  seen.  A second  fracture  yields  an  oblique  section  of  the  left  lateral  aspect  of  the  skull.  The  circum-orbital  bones 
have  been  interpreted  through  study  of  the  specimen,  an  X-ray  of  the  second  half-nodule  (text-fig.  3)  and 
tomographs.  The  remainder  of  the  skull,  including  the  snout,  lateral  and  ventral  margins,  portions  of  the  palate 
and  mandibles  have  been  interpreted  almost  entirely  from  this  X-ray  and  from  the  tomographs. 

The  skull  is  preserved  in  three  dimensions  with  some  distortion  and  fracturing  of  the  snout,  ventral  skull 
margins,  and  tabular-suspensorium  regions.  As  is  the  case  with  the  post-cranial  portion  of  the  specimen,  the 


text-fig.  2.  a,  correlation  chart  of  Lower  Carboniferous  stratigraphy  in  the  Scottish  Midland  Valley  (after 
George  et  al.  1976).  b,  range  chart  of  the  ‘lepospondyls’  (Carroll  1977;  Olson  1972;  Thomson  and  Bossy  1970). 


WELLSTEAD:  LOWER  CARBONIFEROUS  AMPHIBIAN 


197 


bone  is  well  preserved,  but  much  softer  than  the  enclosing  matrix,  rendering  mechanical  preparation  hazardous. 
Bones  are  coated  with  pyrite,  which  accentuates  sutures  as  well  as  cracks,  but  adheres  closely  to  the  bone.  This 
coating  obscures  any  ornamentation  and  evidence  of  lateral  line  canals  which  might  exist.  The  skull  is  triangular 
in  shape,  widest  at  its  posterior  extreme  and  is  approximately  twice  as  wide  as  it  is  high.  Openings  are  present  in 
the  skull  roof  for  external  nares,  orbits,  the  parietal  opening,  and  temporal  fenestrae.  The  orbits  are  in  the 
anterior  third  of  the  skull. 


text-fig.  3.  X-ray  photograph  of  skull,  Lethiscus  stocki,  MCZ2185,  less  the  skull  roof 
posterior  to  fronto-parietal  suture,  x 3. 


198 


PALAEONTOLOGY,  VOLUME  25 


The  relationships  of  the  bones  in  the  skull  table  can  be  viewed  directly  and  are,  therefore,  more  confidently 
interpreted  than  are  those  seen  in  X-ray.  The  bones  surrounding  the  parietal  opening  are  assumed  to  be  the 
parietals  and  further  homologies  within  the  skull  table  follow  the  discussion  of  the  skull  bones  in  Ophiderpeton 
by  Thomson  and  Bossy  (1970,  p.  24). 

The  parietals  bear  dentate  sutures  with  surrounding  bones  and  are  fused  to  one  another  posterior  to  the 
parietal  opening.  The  postparietals  are  also  fused  and  together  with  the  parietals  comprise  two-thirds  of  the  skull 
table  area.  The  postparietals  are  broader  posteriorly  than  at  the  parietal  border  and  extend  to  the  rear  margin  of 
the  skull. 

The  postparietals  are  excluded  from  the  temporal  fenestrae  by  the  supratemporal  and  tabular  bones. 
Intertemporal  bones  are  absent.  The  supratemporals  are  rectangular  bones  approximately  three  times  as  long  as 
they  are  wide.  Their  medial  sutures  with  the  postparietals  and  parietals  are  smoothly  sinuous  while  their  anterior 
margins  have  dentate  sutures  with  the  parietal  bones.  The  posterior  margins  are  poorly  defined.  Ventrally,  the 
supratemporals  appear  to  bear  sutures  with  the  pterygoids,  although  the  contact  is  obscured  by  sediment  and 
breakage  along  the  parafrontal  fracture. 

A suture-like  lineation,  seen  in  the  specimen  intersecting  the  posterior  margin  of  the  left  postparietal,  taken  in 
conjunction  with  the  posterolateral  borders  of  the  skull  table  and  the  posterior  regions  of  the  supratemporals, 
provides  the  evidence  for  the  tabular  bones.  An  element  between  the  right  supratemporal  and  the  fused 
postparietals  appearing  to  bear  a contact  with  the  parietals  may  be  a medial  process  of  the  right  tabular. 
However,  the  mate  of  this  ‘tabular’  process  is  not  found  on  the  left  side  of  the  skull  table,  suggesting  that  the 
feature  is  more  probably  an  irregular  fracture  within  the  postparietals  and  that  the  tabulars  have  no  contact  with 
the  parietals.  The  ventral  margins  of  the  tabulars  are  obscured  by  sediment  and  breakage,  but  would  seem  to 
have  had  a short  suture  with  the  squamosal  posterior  to  the  squamosal-pterygoid  contact. 

The  relationships  of  the  frontals,  parietals,  and  the  right  orbital  series  can  be  observed  directly.  The  frontals 
are  paired,  narrow,  and  equal  in  length  to  the  parietal  and  postparietal.  They  share  an  interdigitating  suture  with 
the  parietals,  but  have  a sinuous  suture  with  pre-  and  postfrontals  and  postorbitals. 

The  fragment  anterolateral  to  the  left  frontal  is  taken  questionably  as  the  small  left  nasal. 

The  postfrontal  and  postorbital  separate  the  orbit  from  the  temporal  fenestrae.  The  prefrontals  are  small 
elements  of  indistinct  shape.  The  postfrontals  are  approximately  twice  as  long  as  the  prefrontals  and  are  wedge- 
shaped.  While  the  posterior  portion  of  the  right  postorbital  is  missing,  the  left  one  seems  to  be  complete  and  is 
large,  approximately  2-5  times  the  size  of  the  postfrontal.  The  exact  nature  of  the  relationship  of  postorbital  to 
the  lateral  edge  of  the  parietal  is  obscured  by  sediment. 

The  lacrimals  are  elongate  bones  which  seem  to  form  the  anterior  portions  of  the  orbital  margin  and  extend  to 
the  external  nares. 

The  left  and  right  jugals  can  be  seen  directly  in  the  specimen,  although  the  left  jugal  is  the  better  exposed  along 
the  oblique  lateral  fracture  surface  (text-fig.  4a).  The  indefinite  suture  indicated  between  jugal  and  quadratojugal 
represents  a conspicuous  separation  between  two  bony  elements  in  the  specimen,  but  which  cannot  be 
interpreted  confidently  as  either  a suture  or  fracture.  The  dorsal  surfaces  of  these  two  bony  elements  appear  in 
text-fig.  3 as  one  distinct  dark  element.  The  orbital  margin  of  the  jugal  is  visible  in  X-ray  as  well.  This  portion  of 
the  orbital  margin  is  completed  dorsally  by  a large  anterodorsal  process  of  the  jugal. 

The  quadratojugal  is  an  elongate  bone  which  bears  a long  suture  with  the  maxilla  and  a comparatively  brief 
one  with  the  postorbital.  The  posterior  portion  of  the  left  quadratojugal  appears  to  have  been  lost  through 
breakage.  However,  some  indication  of  this  portion  of  the  right  quadratojugal  is  supplied  by  the  lateral  margin 
of  the  subtemporal  fossa  (text-figs.  3 and  5b),  which  is  interpreted  as  quadratojugal.  The  manner  of  attachment 
of  the  quadratojugal  to  the  suspensorium  is  uncertain. 

An  element  appearing  in  the  X-ray  at  the  anterior  extremity  of  the  specimen  to  the  right  of  the  midline  is 
interpreted  as  the  right  premaxilla.  The  element  has  one  process  directed  laterally  toward  the  right  maxilla  and  a 
second  directed  posteriorly  toward  the  right  frontal  bone.  No  teeth  can  be  distinguished,  however. 

Both  left  and  right  maxillae  appear  in  X-ray.  They  are  long  slim  bones  which  extend  well  posterior  to  the  orbits 
and  form  a portion  of  the  narial  margin.  The  right  maxilla  is  essentially  in  place  and  is  cracked  along  its  orbital 
margin.  It  is  expanded  anteriorly  into  a broad  process.  The  indentification  of  this  process  is  uncertain,  for 
depending  upon  the  amount  of  distortion  in  this  region  of  the  skull,  it  may  be  either  a nasal  process  or  palatal 
process  of  the  maxilla.  There  are  11  teeth  in  the  left  maxilla  and  18  in  the  right. 

A limited  portion  of  the  left  squamosal  (text-figs.  1 and  4a)  is  exposed.  It  is  a rounded,  rectangular  bone 
and  is  somewhat  displaced  to  exhibit  a portion  of  its  sutural  contact  with  the  quadrate  ramus  of  the 
pterygoid. 


WELLSTEAD:  LOWER  CARBONIFEROUS  AMPHIBIAN 


199 


C 

2 mm 


text-fig.  4.  Lethiscus  stocki,  MCZ2185.  A,  lateral  view  of  skull.  Composite  of  information  from 
left  and  right  sides  of  skull,  x3.  b,  left  lateral  view  of  vertebra  seven.  Composite,  x 5.  c,  ventral 
view  of  vertebra  fifteen,  x 5.  d,  left  lateral  views  of  vertebrae  57  and  58,  x 5.  e,  notched 
sarcopterygian  scale  and  associated  elongate  element,  x 5. 


Temporal  fenestrae 

The  temporal  fenestrae  (text-figs.  4a  and  6a)  are  bounded  dorsally  by  the  parietal,  supratemporal,  and  tabular 
bones,  anteriorly  by  the  postorbital,  laterally  by  the  quadratojugal,  and  medially  by  the  squamosal  and  by  the 
ascending  flange  of  the  quadrate  ramus  of  the  pterygoid.  The  fenestrae  continue  to  the  posterior  margin  of  the 
skull. 

Palate 

Elements  of  the  palate  (text-fig.  5b)  are  interpreted  entirely  from  the  X-rays. 

Two  elements,  which  appear  in  tomographs  of  ventral  portions  of  the  skull,  may  be  vomers  judging  from  their 
relatively  anterior  position.  Their  relationships  to  each  other  and  to  surrounding  bones  are  not  known.  The  right 
vomer  appears  to  bear  three  teeth  comparable  in  cross-sectional  area  to  maxillary  teeth.  Nothing  of  the  palatine 
bones  can  be  distinguished  in  the  X-rays. 

Portions  of  the  right  pterygoid  and  ectopterygoid  are  seen  in  text-fig.  5.  The  pterygoid  bears  a high  dorsal 
flange  (text-fig.  4a)  which  extends  from  the  quadrate  ramus  toward  the  supratemporal  and  posteriorly  contacts 
the  squamosal.  The  portion  of  the  right  pterygoid  identified  in  text-fig.  1b  is  probably  a fragment  of  this  flange. 
The  anterior  portions  of  the  pterygoids  cannot  be  discerned  in  the  X-ray  and  may  be  missing.  Sutural  contact 
between  the  pterygoid  and  supratemporal  is  obscured  by  matrix  and  bone  loss  along  the  fracture  surface  of  the 
nodule. 


200 


PALAEONTOLOGY,  VOLUME  25 


The  subtemporal  fossa  is  defined  by  the  pterygoid,  ectopterygoid  and  quadratojugal.  The  fossa  is  small  and  is 
constricted  by  a blunt  portion  of  the  quadratojugal.  Its  small  size  and  irregular  lateral  margin  suggest  that  this 
region  has  been  distorted. 

The  parasphenoid  appears  clearly  in  X-ray.  It  is  displaced  to  the  right,  partially  overlying  the  medial  edge  of 
the  right  pterygoid.  The  cultriform  process  is  long  and  has  a narrow  base.  The  basicranial  processes  can  be 
confidently  interpreted  on  either  side  of  the  base  of  the  cultriform  process  of  the  parasphenoid.  The  posterior 
portions  of  the  parasphenoid  cannot  be  distinguished  from  the  basisphenoid. 


text-fig.  5.  Lethiscus  stocki,  MCZ  2185.  a,  mandibles  interpreted  from  X-ray  and  viewed  dorsally, 
x 3.  B,  portions  of  palate  and  braincase  interpreted  from  X-ray  and  viewed  dorsally,  x 3. 


Braincase 

There  is  no  evidence  of  braincase  contact  with  the  dermal  roofing  bones,  suggesting  that  the  dorsal  portions  of 
the  braincase  were  not  ossified. 

The  oblong.  X-ray  opaque  structure  (text-fig.  3)  centrally  placed  in  the  braincase  is  taken  to  be  the  matrix-filled 
brain  cavity.  It  is  distorted  to  the  right  as  is  the  entire  braincase-parasphenoid  unit. 

A suture  distinguishes  the  basisphenoid  region  of  the  braincase  from  the  basioccipital.  The  structures  lateral  to 
the  brain  cavity  are  assumed  to  be  the  otic  capsules,  although  pro-  and  opisthotic  bones  cannot  be  individually 
distinguished  nor  are  the  otic  regions  clearly  delimited  from  the  occipital  elements. 

A tubular  bone  is  exposed  in  the  calcite  filling  of  the  inter-orbital  space  (text-fig.  4a).  It  tilts  posteriorly 
as  it  rises  toward  the  skull  roof  and  is  broken  at  its  ventral  extremity.  From  its  structure  the  bone  appears  to 
be  the  left  epipterygoid,  although  displaced  anteriorly  from  its  normal  position  dorsal  to  the  basicranial 
articulation. 


WELLSTEAD:  LOWER  CARBONIFEROUS  AMPHIBIAN 


201 


text-fig.  6.  Comparison  of  skulls,  a,  reconstruction  of  Lethiscus  stocki,  MCZ  2185,  x3. 
B,  Ophiderpeton  (Thomson  and  Bossy  1970),  drawn  to  length  of  Lethiscus  skull. 


Mandibles 

With  the  exception  of  tooth  alveoli  revealed  in  a fracture  section  of  the  left  mandible,  the  lower  jaws  are  visible 
only  as  outlines  in  X-ray.  The  left  mandible  lies  with  its  medial  surface  turned  dorsally.  The  right  mandible  has  its 
lateral  surface  upward.  In  the  posterior  margin  of  each  mandible,  U-shaped  structures  corresponding  to  the 
positions  of  surangular,  articular,  and  angular  can  be  differentiated,  but  no  sutures  can  be  discerned  between 
these  elements. 

An  elongate,  rectangular  element  lies  across  the  right  mandible  approximately  one-third  of  its  length  from  the 
symphysis.  It  bears  a pointed  process  and  is  tentatively  identified  as  a coronoid,  probably  of  the  right  mandible. 

Fourteen  teeth  are  apparent  in  the  left  mandible.  Nine  can  be  counted  in  the  right.  The  teeth  of  the  mandibles 
seem  to  be  short  and  peg-like  as  are  those  of  the  maxillae,  though  the  nature  of  their  crowns  is  not  certain. 

Hyoid  elements 

Two  elongate  elements  revealed  in  the  X-ray  (text-fig.  3;  also  1b)  are  hyoid  elements,  perhaps  epibranchials. 
Fragments  lateral  to  vertebrae  4 and  5 (text-fig.  7a)  may  be  additional  hyoid  elements  or  possibly  ribs,  but  no 
positive  identification  can  be  made.  There  is  no  evidence  of  gill  rakers,  internal  or  external  gills,  nor  of  the  sickle- 
shaped hyoid  noted  in  other  ai'stopods  (Baird  1964). 

Vertebrae 

Seventy-eight  vertebrae  are  visible  in  sequence  (text-figs.  7,  8,  9).  An  additional  vertebra  can  be  distinguished  in 
X-rays  dorsal  to  vertebra  65.  Vertebrae  1 through  5,  9 through  12,  15  through  28,  and  42  through  46  are  viewed 
ventrally.  Alternating  with  these  series  are  vertebrae  exposed  in  lateral  view  and  vertebrae  49,  50,  and  51  which 
are  seen  end-on. 

The  vertebrae  are  clearly  holospondylous.  Bony  elements  seen  between  centra  of  the  sixth,  seventh,  and  eighth 
vertebrae  are  found  nowhere  else  in  the  column  and  may  be  mineralized  intercentral  cartilages  similar  to  those 
described  in  salamanders  (Wake  1970;  Wake  and  Lawson  1973)  or  merely  displaced  fragments  of  the  adjacent 
centra. 


202 


PALAEONTOLOGY,  VOLUME  25 


The  centra  are  hour-glass-shaped  and  deeply  amphicoelous.  Sections  along  the  column  show  the  notochord  to 
be  severely  constricted  and  possibly  discontinuous  at  midcentrum.  The  lateral  and  ventral  surfaces  of  the  centra 
are  smooth  except  for  slightly  concave,  round  facets  near  the  anterior  rim  which  are  the  articular  surfaces  for  the 
rib  capitulum  (text-fig.  4b).  The  centra  bear  no  pits,  grooves,  or  accessory  processes. 

Vertebral  length  increases  antero-posteriorly  along  most  of  the  column.  The  average  length  for  vertebrae  6, 7, 
and  8 is  4-5  mm.  Vertebrae  34,  35,  and  37  average  5-2  mm  in  length,  while  vertebrae  52  through  78  average 
approximately  6 mm  in  length.  The  untapered  nature  of  these  vertebrae,  as  well  as  the  presence  of  ribs  along  the 
column,  suggests  that  this  portion  of  the  skeleton  represents  the  trunk  of  Lethiscus.  The  isolated  vertebra  is  only 
4 mm  long  and,  by  virtue  of  its  small  size,  is  the  only  indication  of  a tail  in  the  specimen. 

Neural  arches  of  the  vertebrae  are  swollen,  unpaired,  and  are  fused  to  their  centra  with  no  trace  of  suture  The 
pedicel  length  is  approximately  two-thirds  that  of  the  centrum.  The  neural  spines  anterior  to  vertebra  37  are  not 
well  exposed.  However,  a transverse  section  of  vertebra  9 (text-fig.  10a)  shows  the  spine  to  be  relatively  high. 
More  posteriorly,  the  neural  arches  can  be  seen  to  extend  the  length  of  the  neural  arch  and  to  be  tall,  rising 
antero-posteriorly  (text-fig.  9a  and  b).  The  spines  bear  jagged  edges  and  have  faint  grooves  between  the  teeth  of 
the  serrations,  suggesting  a crinkled  appearance.  It  is  not  clear  whether  the  jagged  appearance  is  natural  or  due  to 
poor  ossification  or  breakage.  Transverse  sections  through  the  vertebral  column  show  that  the  neural  spines 
bifurcate  at  their  posterior  extremities  and  bear  a deep  medial  groove  (text-fig.  10  c and  d). 

Neural-arch  processes  bearing  zygapophyses  project  at  approximately  30°  from  the  sagittal  plane,  but  extend 
little  laterally  beyond  the  centrum  (text-fig.  10c  and  d).  Zygapophyses  are  oblique  to  the  sagittal  plane. 

All  post-atlantal  vertebrae  bear  stout  transverse  processes  which,  in  the  more  anterior  vertebrae,  are  located 
on  the  neural  arches  (text-figs.  7a  and  10a),  rather  than  on  the  centra  as  in  other  a'fstopods.  The  processes  project 
laterally,  but  breakage  obscures  the  surface  of  rib  articulation.  In  contrast  to  the  transverse  processes  of  the 
anterior  vertebrae,  the  transverse  processes  of  vertebrae  44,  57,  and  68  through  78  can  be  seen  to  arise,  in  part, 
from  the  centra  (text-figs.  9 and  10b). 


text-fig.  7.  Lethiscus  stocki , MCZ  2185,  postcranial  skeleton,  x 1-4.  a,  vertebrae  l through  15. 
B,  vertebrae  15  through  30. 


4U\\ 


204 


PALAEONTOLOGY,  VOLUME  25 


Where  the  lateral  surfaces  of  the  posterior  vertebrae  are  exposed,  as  in  vertebrae  57,  58,  and  68,  a 
fossa  is  seen  immediately  posterior  to  the  transverse  processes  (text-fig.  9a).  In  the  floor  of  this  fossa 
are  foramina,  presumably  for  exit  of  spinal  nerves.  Such  foramina  cannot  be  identified  in  vertebra  7, 
the  only  anterior  vertebra  suitably  exposed. 

Ribs 

Fragments  of  ribs  are  exposed  along  the  length  of  the  vertebral  column,  but  are  best  revealed  posterior  to 
vertebra  36.  Here,  ribs  are  seen  to  have  a dog-legged  bend  just  posterior  to  the  rib  head.  Distal  to  this  bend  the 
ribs  are  straight.  An  accessory  rib  process  is  suggested  by  the  sharp  angle  of  the  bend,  but  no  substantial  evidence 
of  the  characteristic  K-shaped  ai'stopod  rib  is  exhibited  in  the  specimen.  The  ribs  are  bicipital  anterior  to  vertebra 
44  (text-fig.  1 0 a and  b),  but  the  nature  of  their  vertebral  articulation  is  not  exposed  more  posteriorly.  The  rib  head 
is  set  at  a sharp  angle  to  its  body  as  can  be  seen  in  the  vicinity  of  vertebrae  19,  20,  43,  and  60. 


text-fig.  10.  Selected  sections  of  vertebrae,  x 5.  Roman  numerals  indicate  approximate  position  of  section  in 
reference  vertebra,  a,  vertebra  9.  b,  vertebra  44.  c,  vertebra  15  and  prezygapophyses  of  vertebra  16. 

D,  vertebra  23. 


Ventral  armour 

Numerous,  spindle-shaped  dermal  elements  comprise  the  ventral  armour,  which  would  seem  to  have  been 
continuous  posterior  to  the  level  of  the  sixth  vertebra.  The  pattern  of  the  ventral  armour  has  been  severely 
disrupted.  The  ventral  armour  elements  in  Lethiscus  appear  to  be  stouter  and  to  have  blunter  ends  than  those  in 
Ophiderpeton , but  this  appearance  may  be  due  to  the  irregular  sections  in  which  the  elements  are  exposed. 

Dorsal  osteoderms 

Baird  (1964)  noted  dorsal  osteoderms  in  this  specimen,  but  the  only  elements  which  might  be  interpreted  as 
osteoderms  are  actually  mineralized  gas  bubbles  (text-fig.  8a).  Sections  of  these  elements  show  them  not  to  be 
bone,  but  calcite  cores  coated  with  pyrite.  Gas  bubbles  preserved  in  this  manner  have  been  reported  in  specimens 
from  the  Wardie  Shales  previously  (Wood  1977). 

Enterospira 

In  the  vicinity  of  the  forty-first  vertebra  is  what  appears  to  be  a coprolite.  Its  segmented  appearance  is  likely  due 
to  a spiral  valve  in  the  intestine  of  Lethiscus.  As  the  coprolite  has  not  been  excreted,  the  term  enterospira  may  be 
more  correct  (Williams  1972). 


WELLSTEAD:  LOWER  CARBONIFEROUS  AMPHIBIAN 


205 


Sarcopterygian  scales  and  possible  pectoral  girdle 

Dorsal  to  vertebrae  6,  7,  and  8 are  many  bony  elements.  Most  of  these  are  the  spindle-shaped  armour  seen 
elsewhere  in  the  specimen.  Four  others  are  probably  slim  fragments  of  rib.  Most  interesting  are  the  three  largest 
elongate  bones  and  three  irregularly  round  elements  near  them. 

The  round  elements  bear  faint  concentric  rings  and  compare  favourably  with  scales  of  sarcopterygian  fishes.  It 
should  be  noted,  however,  that  concentric  growth  structures  similar  to  these  rings  occur  in  endochondral  limb 
and  girdle  elements  of  tetrapods  as  well  ( Mesosaurus , de  Ricqles  1974).  The  middle  ‘scale’  has  a notch  in  its  rim 
which  is  similar  to  a glenoid  fossa  (text-fig.  4e).  In  the  notch  is  a tiny  fragment  of  bone,  but  whether  the  fragment 
has  actually  come  to  rest  in  a notch-like  glenoid  or  whether  compressional  forces  merely  created  the  notch  by 
forcing  the  fragment  into  the  rim  is  uncertain. 

The  three  large  elongate  elements  could  be  posteriorly  displaced  hyoid  elements  or  possibly  remnants  of  the 
dermal  pectoral  girdle. 

Although  these  six  bones  occur  where  a pectoral  girdle  would  be  expected,  their  lamentably  poor  exposure 
allows  no  confident  identification.  There  is  no  other  evidence  of  girdles  or  limbs  in  the  specimen. 


COMPARISONS 

Skull 

The  skulls  of  both  Lethiscus  and  Ophiderpeton  (text-fig.  6)  are  relatively  high-sided  and  possess 
temporal  fenestrae  bordered  by  identical  elements  of  the  skull  roof.  The  orbits  are  anterior  in 
position.  Maxillae  are  elongate,  slim,  and  extend  far  posterior  to  the  orbit.  The  frontal  bones  have  a 
similar  proportional  length  relationship  with  the  orbits.  In  Lethiscus,  Ophiderpeton,  and  nectrideans 
(Thomson  and  Bossy  1970)  the  skull  table  is  comprised  primarily  by  the  parietal  and  postparietal 
bones.  Lethiscus,  however,  appears  to  represent  a more  primitive  pattern  in  that  its  parietal- 
postparietal  suture  is  anterior  to  the  supratemporal-tabular  suture  (Panchen  1970).  As  a result, 
Lethiscus  lacks  the  tabular-parietal  contact,  the  presence  of  which  has  been  suggested  as  linking 
a'istopods  and  nectrideans  (Thomson  and  Bossy  1 970).  The  later  establishment  of  the  tabular  parietal 
contact  in  Ophiderpeton  may  then  have  come  as  the  result  of  the  posterior  movement  of  the  parietal- 
postparietal  suture  as  suggested  in  anthracosaurs  (Panchen  1970).  Concomitant  with  the  movement 
of  this  suture  may  have  been  the  proportional  increase  in  length  of  the  parietal  bones  seen  in 
Ophiderpeton. 

In  contrast  to  Ophiderpeton  the  supratemporals  of  Lethiscus  are  large  and  the  skull  table  is  much 
shorter.  There  is  no  indication  of  the  intertemporal  in  Lethiscus.  This  bone  may  have  been 
incorporated  into  the  large  postorbital.  However,  elongate  pustulated  bones  have  been  identified  as 
intertemporals  in  Ophiderpeton  (Thomson  and  Bossy  1970). 

In  contrast  to  Ophiderpeton  the  jugal  and  quadratojugal  in  Lethiscus  contact  one  another.  The 
quadratojugal  in  Lethiscus  also  bears  contacts  with  the  postorbital  and  maxilla,  perhaps 
strengthening  the  connection  of  the  lateral  skull  margin  with  the  skull  table  in  response  to  bone  loss 
resulting  from  fenestration  of  the  skull  roof.  In  Lethiscus  the  maxilla  contacts  the  narial  margin,  but  is 
excluded  from  it  by  the  lacrimal  in  Ophiderpeton  as  reconstructed  by  Thomson  and  Bossy. 

Other  possible  differences,  such  as  the  presence  of  nasals  and  the  extent  of  the  lacrimals  and 
prefrontals  cannot  be  determined  unequivocally.  The  mandibles  in  Lethiscus  and  the  palate  and 
braincase  in  both  Lethiscus  and  Ophiderpeton  are  also  too  poorly  known  to  allow  useful  comparison. 

Vertebrae 

The  vertebrae  of  Lethiscus  differ  from  those  of  other  a’istopods  in  lacking  basipophyseal  accessory 
articulations  and  median  ventral  ridges  (as  seen  in  Ophiderpeton  nanum,  Steen  1931)  and  in  possessing 
high  neural  spines.  Absence  of  incontestable  limb  girdles  makes  distinction  of  trunk  and  caudal 
regions  difficult.  Lethiscus  does,  however,  exhibit  antero-posterior  differentiation  of  the  vertebral 
column  in  the  position  of  transverse  processes  and  by  the  possession  of  spinal-nerve  foramina  only  in 
the  posterior  portion  of  the  column.  Similar  differentiation  has  not  been  found  in  Ophiderpeton, 
although  McGinnis  (1967)  reported  that  the  two  anterior-most  vertebrae  in  Phlegethontia  do  lack 
spinal-nerve  foramina. 


206 


PALAEONTOLOGY,  VOLUME  25 


The  anterior  vertebrae  of  Lethiscus  are  remarkable  in  their  similarity  to  those  of  some  microsaurs, 
as  well  as  to  those  of  early  reptiles.  These  in  each  case  have  smooth,  unpitted  surfaces,  and  are  hour- 
glass-shaped. The  neural-arch  pedicels  also  bear  the  transverse  processes.  In  contrast,  neuro-central 
sutures  are  present  consistently  within  the  microbrachiomorph  microsaurs  and  variably  within  the 
tuditanomorphs,  but  are  absent  in  Lethiscus.  Lethiscus  also  lacks  trunk  intercentra,  which  occur  in 
several  microsaur  genera. 

The  vertebrae  of  adelogyrinids  and  ‘lysorophids’  (including  both  the  Molgophidae  and 
Lysorophidae)  differ  from  those  of  Lethiscus  in  the  consistent  presence  of  neuro-central  sutures,  but 
are  similar  in  lacking  accessory  articulations  and  in  bearing  the  transverse  processes  on  neural  arches. 
The  lysorophids  further  differ  in  the  paired  nature  of  their  neural  arches.  The  paired  status  of  neural 
arches  in  the  adelogyrinids  is  equivocal  (Carrol  1 967;  Brough  and  Brough  1 967;  Watson  1921-1 923). 

The  trunk  vertebrae  of  nectrideans  are  similar  to  those  of  Lethiscus  in  lacking  intercentral  elements 
and  in  ossifying  as  single  units.  However,  the  only  described  nectridean  in  which  intravertebral  spinal 
nerve  foramina  can  actually  be  seen  is  the  urocordylid  Crossotelos,  keraterpetontids  do  not  possess 
such  spinal  nerve  foramina  (Milner,  A.  C.,  pers.  comm.).  Furthermore,  nectridean  neural  spines  are 
specialized  in  their  possession  of  accessory  articulations,  and  rugose  ornamentation  (as  in 
Diploceraspis  and  Diplocaulus),  or  in  being  flat-topped,  fan-shaped  structures  with  crenulated  edges 
(as  in  Sauropleura  and  Keraterpeton , Baird,  1965;  Steen  1938).  Some  faint  suggestion  of  neural-spine 
crenulation  is  present  in  Lethiscus , but,  as  noted,  the  neural  spines  are  otherwise  serrated  and  inclined 
antero-posteriorly  to  the  frontal  plane. 

The  high  number  of  trunk  vertebrae  found  in  Lethiscus  is  seen  elsewhere  only  in  a'fstopods  and  in 
the  tiny-limbed  Lysorophus  (Olson  1971)  among  the  iepospondyl’  amphibians.  Nectrideans,  in 
contrast,  characteristically  have  short  trunks  (and  long  tails). 

Limbs  and  girdles 

Baird  (1964)  wrote  that  there  is  nothing  interpretable  as  pectoral  or  pelvic  girdle  in  any  aistopod. 
Although  there  is  no  contradictory  evidence  in  Lethiscus,  a recent  study  (see  Boyd,  M.  J.  F.,  this 
volume)  has  discovered  an  interclavicle  in  Ophiderpeton  nanum. 

Ventral  armour 

Spindle-shaped  ventral  armour  like  that  in  Lethiscus  is  seen  in  other  ai'stopods  and  nectrideans 
(Fritsch  1879;  Huxley  1867)  and  with  sculptured  surfaces  in  the  adelogyrinid  Adelospondylus  (Carroll 
1967).  Ventral  armour  in  the  microsaurs  is  variable  (Carroll  and  Gaskill  1978),  but  never  consists  of 
spindle-shaped  elements.  Such  dermal  armour  is  unknown  in  the  lysorophids. 


DISCUSSION 

Lethiscus  possesses  nearly  all  the  aistopod  characteristics  compiled  by  Baird  (1964).  Aistopod 
character  states  which  can  not  be  confidently  identified  in  Lethiscus  (e.g.  hypapophyseal  flanges  of 
caudal  vertebrae,  K-shaped  ribs,  and  sickle-shaped  hyoid)  are  those  in  portions  of  the  specimen  not 
preserved  or  which  are  poorly  exposed. 

While  Lethiscus  is  more  closely  comparable  to  Ophiderpeton  than  to  Phlegethontia  in  the  relatively  j 
primitive  nature  of  its  skull,  robust  ribs,  and  heavy  ventral  armour,  it  is  distinct  in  further  details  of  its 
skull  and  post-cranial  anatomy.  Its  short  skull  and  the  absence  of  a tabular-parietal  contact  indicate  i 
a less  derived  state  than  that  of  Ophiderpeton , while  the  absence  of  accessory  vertebral  processes  and 
K-shaped  ribs  and  the  presence  of  tall  neural  spines  indicate  that  Lethiscus  possessed  a differently 
specialized  post-cranial  skeleton  and  trunk  musculature  than  either  the  Ophiderpetontidae  or  the  j 
Phlegethontiidae. 

The  presence  of  spinal-nerve  foramina  in  a portion  of  the  vertebral  column  of  Lethiscus  is  j 
especially  interesting,  for  such  foramina  are  known  to  occur  only  in  urodelan  lissamphibians,  the  j 
A'istopoda,  and  the  nectridean  Crossotelos.  The  presence  of  these  foramina  is  considered  to  be  a 
derived  state  in  salamanders  (Edwards  1 976;  Hecht  and  Edwards  1977)  and  would  appear  to  be  so  in 


WELLSTEAD:  LOWER  CARBONIFEROUS  AMPHIBIAN 


207 


aistopods  and  Crossotelos.  Particularly  significant  is  the  observation  that  the  spinal-nerve  foramina 
of  salamanders  are  expressed  in  patterns  characteristic  of  the  various  families  (Edwards  1976). 
Although  the  pattern  of  spinal-nerve  foramina  is  not  well  known  in  the  vertebral  column  of 
Ophiderpeton,  the  contrast  in  patterns  between  Phlegethontia  and  Lethiscus  suggests  that  the  spinal- 
nerve  foramina  pattern  may  allow  distinction  of  aistopod  families  also. 

The  occurrence  of  an  aistopod  in  mid-Visean  rocks  provides  some  limited  confirmation  of  the 
estimated  several  million  or  tens  of  millions  of  years  required  to  accomplish  limb  loss  in  tetrapods 
(Lande  1977).  Although  Lethiscus  cannot  be  described  with  absolute  certainty  as  lacking  limbs  or 
girdles,  the  rudiments  of  a possible  pectoral  girdle  demonstrate  the  degenerate  nature  of  any  limbs  it 
may  have  possessed.  Assuming  tetrapod  monophyly,  this  limb  reduction  was  achieved  within  a 
period  of  at  least  30  million  to  40  million  years  elapsing  between  Late  Devonian  tetrapod  origins, 
represented  by  Metaxygnathus  (Campbell  and  Bell  1977)  and  Ichthyostega,  and  the  occurrence  of 
Lethiscus  in  the  mid-Visean. 

Lethiscus  is  the  earliest  known  member  of  a group  of  small  Palaeozoic  amphibians  known  as 
‘lepospondyls’,  but  unfortunately  it  reveals  little  about  the  evolution  of  any  of  these  animals  or  of 
tetrapods  in  general  because  of  the  specializations  of  the  skull  and  post-cranial  anatomy  already 
attained  in  the  Lower  Carboniferous.  Lethiscus  seems  to  confuse  the  issue  somewhat,  for,  although 
Thomson  and  Bossy  (1970)  linked  nectrideans  and  aistopods  through  the  shared  possession  of  the 
tabular-parietal  contact,  the  absence  of  such  a contact  in  Lethiscus  suggests  that  it  was  either  achieved 
independently  in  the  two  orders  or  was  lost  in  Lethiscus  subsequent  to  a nectridean-ai'stopod 
dichotomy.  Similarly,  the  extremely  limited  occurrence  of  intravertebral  spinal-nerve  foramina  in 
nectrideans  suggests  that  these  foramina  were  probably  developed  separately  in  nectrideans  and 
aistopods,  arguing  against  consideration  of  the  presence  of  the  foramina  as  a shared  derived- 
character  state. 

Acknowledgements.  I am  pleased  to  acknowledge  Dr.  Robert  L.  Carroll  for  giving  me  the  opportunity  to  study 
this  specimen.  This  report  benefitted  critically  from  comments  received  from  Dr.  Carroll,  Dr.  Donald  Baird,  Dr. 
A.  R.  Milner,  Dr.  A.  C.  Milner,  Mr.  Timothy  Smithson,  and  Mr.  Robert  Holmes,  to  all  of  whom  I am  grateful. 

X-ray  photography  has  been  indispensible  to  the  description  of  portions  of  this  specimen.  I thank  Dr.  Robert 
Hanson  and  Sandra  and  Andre  Hamelin  of  the  Radiology  Unit  of  the  Royal  Victoria  Hospital,  Montreal,  for 
their  kind  and  patient  assistance. 


REFERENCES 

baird,  d.  1955.  Latex  micro-molding  and  latex-plaster  molding  mixture.  Science,  122,  202. 

1964.  The  aistopod  amphibians  surveyed.  Breviora,  No.  206,  1-17. 

1965.  Paleozoic  lepospondyl  amphibians.  Am.  Zool.  5,  287-294. 

brough,  m.  c.  and  brough,  j.  1967.  Studies  of  early  tetrapods.  I.  The  Lower  Carboniferous  microsaurs.  II. 

Microbrachis,  the  type  microsaur.  III.  The  genus  Gephyrostegus.  Phil.  Trans.  R.  Soc.  252,  107-165. 
Campbell,  k.  s.  w.  and  bell,  m.  w.  1977.  A primitive  amphibian  from  the  late  Devonian  of  New  South  Wales. 
Alcheringa,  1,  369-381. 

carroll,  r.  l.  1967.  An  adelogyrinid  lepospondyl  amphibian  from  the  Upper  Carboniferous.  Can.  J.  Zool.  45, 
1-16. 

— 1977.  Patterns  of  amphibian  evolution:  an  extended  example  of  the  incompleteness  of  the  fossil  record.  In 
hallam,  a.  (ed.).  Patterns  of  evolution  as  illustrated  by  the  fossil  record.  Elsevier  Sci.  Publ.  Co.,  Amsterdam. 
591  pp. 

— and  gaskill,  p.  1978.  The  order  Microsauria.  Mem.  Am.  Phil.  Soc.  126,  211  pp. 

edwards,  j.  l.  1976.  Spinal  nerves  and  their  bearing  on  salamander  phylogeny.  J.  Morph.  148,  305-328. 
fitch,  f.  j.,  miller,  j.  a.  and  williams,  s.  c.  1970.  Isotopic  ages  of  British  Carboniferous  rocks.  In  Sixieme 
Congress  International  de  Stratigraphie  et  de  Geologie  du  Carbonifere,  Sheffield,  1967.  2,  771-789. 
fritsch,  A.  1 879.  Fauna  der  Gaskohle  und der  Kalksteine  der  Permformation  Bohmens.  1 . Prague.  1 82  pp.,  48  pis. 
GEORGE,  T.  N.,  JOHNSON,  G.  A.  L.,  MITCHELL,  M.,  PRENTICE,  J.  E.,  RAMSBOTTOM,  W.  H.  C.,  SEVASTOPULO,  G.  D., 

wilson,  r.  b.  1976.  A correlation  of  Dinantion  rocks  in  the  British  Isles.  Spec.  Rept.  7,  Geol.  Soc.  Lond.  87  pp. 


208 


PALAEONTOLOGY,  VOLUME  25 


hecht,  m.  k.  and  Edwards,  J.  l.  1977.  The  methodology  of  phylogenetic  inference  above  the  species  level.  In 
hecht,  m.  k.,  goody,  p.  c.  and  hecht,  B.  M.  (eds.).  Major  patterns  in  vertebrate  evolution.  Plenum,  New  York, 
pp. 3-51. 

huxley,  t.  h.  1 867.  Description  of  vertebrate  remains  from  Jarrow  colliery.  Pt.  1 . Trans.  R.  Ir.  Acad.  24  (Science), 
353-369. 

lande,  r.  1977.  Evolutionary  mechanisms  of  limb  loss  in  tetrapods.  Evolution,  32,  73-92. 
mcginnis,  h.  j.  1967.  The  osteology  of  Phlegethontia,  a Carboniferous  and  Permian  ai'stopod  amphibian.  Univ. 
Calif.  Pubis,  geol.  Sci.  71,  1-46. 

mitchell,  f.  h.  and  mykura,  w.  1962.  The  geology  of  the  neighbourhood  of  Edinburgh.  Mem.  geol.  Surv.  U.K., 
3rd  edn.,  159  pp. 

olson,  e.  c.  1971.  A skeleton  of  Lysorophus  tricarinatus  (Amphibia:  Lepospondyli)  from  the  Hennessey 
Formation  (Permian)  of  Oklahoma.  J.  Paleont.  45,  443-449. 

— 1972.  Diplocaulus  parvus  n.sp.  (Amphibia:  Nectridea)  from  the  Chikasha  formation  (Permian: 
Guadalupian)  of  Oklahoma.  Ibid.  46,  656-659. 

panchen,  a.  l.  1970.  Batrachosauria:  Anthracosauria.  Handb.  Palaoherp.  5/A,  84  pp. 
ricqles,  A.  j.  de  1974.  Recherches  paleohistologiques  sur  les  os  longs  des  tetrapodes.  5.  Cotylosaurs  et 
Mesosaures.  Annls.  Paleont.  Vertebres,  60,  13-48,  7 pis. 
steen,  m.  c.  1931.  The  British  Museum  collection  of  Amphibia  from  the  Middle  Coal  Measures  of  Linton,  Ohio. 
Proc.  cool.  Soc.  Loud.  (1930),  849-891. 

1 938.  On  the  fossil  Amphibia  from  the  Gas  Coal  of  Nyrany  and  other  deposits  in  Czechoslovakia.  Ibid.  B, 
108,  205-283. 

stock,  t.  1882.  Notice  of  some  discoveries  recently  made  in  Carboniferous  vertebrate  paleontology.  Nature,  27, 
22. 

Thomson,  k.  s.  and  bossy,  k.  h.  1970.  Adaptive  trends  and  relationships  in  early  Amphibia.  Forma  et  Functio,  3, 
7-31. 

wake,  d.  b.  1970.  Aspects  of  vertebral  evolution  in  the  modern  Amphibia.  Ibid.  33-60. 

— and  lawson,  r.  1973.  Development  and  adult  morphology  of  the  vertebral  column  in  the  plethodontid 
salamander  Euryces  bislineata,  with  comments  on  vertebral  evolution  in  the  Amphibia.  J.  Morph.  139, 
251-300. 

watson,  D.  m.  s.  1921-23.  The  Carboniferous  Amphibia  of  Scotland.  Palaeont.  hung.  1,  221-251. 

williams,  m.  e.  1972.  The  origin  of  spiral  coprolites.  Paleont.  Contr.  Univ.  Kans.  59,  1-19. 

wood,  s.  p.  1977.  Recent  discoveries  of  Carboniferous  fishes  in  Edinburgh.  Scott.  J.  Geol.  2,  251-258. 

CARL  F.  WELLSTEAD 
Redpath  Museum 
McGill  University 
859  Sherbrook  Street  West 
Montreal,  P.Q.,  Canada  H3A  2K6 


Typescript  received  8 July  1980 

Revised  typescript  received  24  October  1980 


MORPHOLOGY  AND  RELATIONSHIPS  OF  THE 
UPPER  CARBONIFEROUS  AISTOPOD  AMPHIBIAN 
OPHIDERPETON  NANUM 

by  M.  J.  BOYD 


Abstract.  The  holotype  and  only  recorded  specimen  of  the  Carboniferous  ai'stopod  amphibian  Ophiderpeton 
nanum  Hancock  and  Atthey  1868  is  described  in  detail  and  figured  for  the  first  time.  The  vertebrae,  ribs, 
dermal  squamation,  and  premaxilla  are  characteristic  of  Ophiderpeton  and  confirm  that  O.  nanum  is  a member 
of  that  genus.  The  relationship  of  O.  nanum  to  other  described  Ophiderpeton  species  is  obscured  by  the  absence 
of  most  of  the  skull  in  the  holotype  and  by  the  apparently  sub-adult  nature  of  the  specimen.  The  ventral 
osteoderms  are,  however,  unusually  filamentous  for  Ophiderpeton  and  it  is  suggested  that  O.  nanum  be  retained 
as  a distinct  species,  pending  revision  of  the  Ophiderpetontidae.  A small  isolated  bone  in  the  holotype  may 
be  an  interclavicle,  suggesting  the  retention  of  a vestigial  pectoral  girdle  in  ophiderpetontid  a'istopods. 

Ophiderpeton  nanum  Hancock  and  Atthey  1868,  is  a small  ‘lepospondyl’  amphibian  of  the 
order  Aistopoda  from  the  Upper  Carboniferous  of  Great  Britain.  The  holotype  and  only  described 
specimen  of  Ophiderpeton  nanum  was  collected  around  the  middle  of  the  nineteenth  century  and 
is  from  the  black  shale  immediately  overlying  the  Low  Main  coal  seam  at  Newsham,  near  Blyth, 
in  Northumberland.  This  horizon  lies  within  the  Upper  Modiolaris  zone  of  the  Middle  Coal 
Measures  (Land  1974)  and  is  Westphalian  B in  age.  O.  nanum  forms  part  of  a large  and  well-known 
amphibian  assemblage  from  Newsham,  whose  other  members  (listed  by  Land  1974,  p.  61)  include 
eogyrinid  embolomeres,  loxommatid  temnospondyls,  and  keraterpetontid  nectrideans.  The  above 
taxa  appear  to  be  characteristic  of  Upper  Carboniferous  coal-swamp  lake  environments  (Milner 
1978). 

Although  O.  nanum  was  one  of  the  first  a'istopods  to  be  described,  no  detailed  account  of  the 
structure  or  relationships  of  this  species  has  hitherto  been  published  and  the  holotype  specimen 
has  never  been  figured.  Neither  the  original,  necessarily  superficial,  description  of  O.  nanum  by 
Hancock  and  Atthey  (1868)  nor  a subsequent  brief  account  given  by  Steen  (1938),  are  detailed 
enough  to  allow  adequate  comparison  of  this  form  with  other  described  a'istopods.  The  present 
paper  is  intended  to  remedy  this  deficiency. 

The  holotype  of  O.  nanum  consists  of  an  incomplete  skull  and  anterior  postcranial  skeleton 
preserved  in  counterpart  on  two  small  slabs  of  shale.  The  two  halves  of  the  specimen  are  registered 
in  the  collections  of  the  Hancock  Museum,  Newcastle  upon  Tyne,  as  G25.34  and  G25.35.  Because 
of  the  small  size  and  fragile  nature  of  the  holotype,  the  only  preparation  attempted  has  been  the 
removal  of  small  quantities  of  matrix  from  the  immediate  vicinity  of  the  skull  by  means  of  mounted 
needles. 


DESCRIPTION 

The  skull  and  most  of  the  preserved  postcranial  skeleton  are  situated  on  slab  G25.34  (text-fig.  la).  Slab  G25.35, 
henceforth  referred  to  as  the  counterpart,  bears  impressions  of  all  but  the  most  anterior  seven  vertebrae  of 
G25.34,  in  addition  to  a small  number  of  actual  ribs  and  isolated  patches  of  the  ventral  squamation.  At  some 
point  in  the  past,  small  areas  of  the  surface  of  slab  G25.34,  bearing  sections  of  vertebral  column  and  associated 
structures,  have  flaked  away  and  been  lost.  However,  as  the  resultant  gaps  in  the  vertebral  series  are  all 
posterior  to  the  seventh  vertebra,  the  number  of  vertebrae  originally  preserved  may  be  estimated  by  reference 
to  the  unbroken  series  of  impressions  on  the  counterpart.  The  skull  and  skeleton  of  G25.34  are  preserved 

IPalaeontology,  Vol.  25,  Part  1,  1982,  pp.  209-214.] 


text-fig.  1.  Ophiderpeton  nanum  H & A.  a-e,  Hancock  Museum  specimen  G25.34:  a,  semi-diagrammatic 
representation  of  specimen  as  preserved.  Stippling  indicates  extent  of  ventral  squamation;  b,  isolated  (?)skull 
element.  Stippling  represents  area  preserved  as  impression  only;  c , interclavicle  as  preserved;  d,  interclavicle 
restored;  e,  detail  of  ventral  squamation.  /,  Hancock  Museum  G25.35.  Three  isolated  vertebrae  as  preserved. 

All  scale  lines  represent  5 mm.  n.sp.,  neural  spine. 


BOYD:  CARBONIFEROUS  AMPHIBIAN  OPHIDERPETON 


2: 


with  their  ventral  surfaces  uppermost.  A large  number  of  ribs  is  present.  Although  most  of  the  vertebrae  and 
ribs  are  overlain  by  a thin  ‘mat’  of  closely  packed,  filamentous  gastralia,  this  is  often  so  closely  applied  that 
the  contours  of  the  structures  which  it  conceals  may  be  clearly  seen. 

Skull 

The  skull  apparently  became  disarticulated  and  scattered  prior  to  preservation  and  is  largely  absent.  Three 
distinct  masses  of  poorly  preserved  bone  lie  immediately  anterior  to  the  first  preserved  vertebra.  It  is  possible 
that  they  represent  elements  of  the  palate  or  neurocranium.  A slight  depression  in  the  matrix  anterior  to  these 
may  mark  the  site  of  the  isolated  premaxilla  noted  by  Steen  (1938,  p.  223).  No  trace  of  the  bone  itself  remains; 
it  has  presumably  become  detached  from  the  slab  and  lost. 

A relatively  large  isolated  bone,  situated  to  the  right  of  the  most  anterior  four  vertebrae  in  text-figure  1 a, 
may,  doubtfully,  also  belong  to  the  dermal  skull.  Although  much  of  the  bone  itself  has  been  lost,  the  clear 
impression  remaining  in  the  matrix  allows  no  question  as  to  its  original  form.  Its  identity,  however,  is  obscure. 
The  element  (text-fig.  1 b)  does  not  appear  to  correspond  to  any  bone  known  in  the  skulls  of  either  Ophiderpeton 
(A.  C.  Milner,  pers.  comm.)  or  phlegethontiid  a'fstopods  (e.g.  Gregory  1948;  McGinnis  1967).  The  triradiate 
form  of  the  element  makes  it  unlikely  that  it  represents  part  of  the  dermal  pectoral  girdle.  Possibly  the  bone 
in  question  does  not,  in  fact,  pertain  to  O.  nanum. 

A second  isolated  element,  lying  to  the  left  of  the  vertebral  column  at  the  level  of  the  third  and  fourth 
vertebrae  (text-fig.  la),  is  of  some  interest.  Approximately  one-quarter  of  the  bone  is  missing  but  it  is  clear 
that  the  complete  element  was  of  roughly  rhomboidal  shape  (text-fig.  1 c-d).  It  is  suggested  that  this  element 
represents  a displaced  interclavicle.  This  hypothesis  receives  support  from  the  presence  of  a slight  (?parasternal) 
process  at  one  angle  of  the  bone  and  the  fact  that  the  margin  of  the  element  directly  opposite  the  above 
process  is  noticeably  fimbriated.  Interclavicles  with  fimbriated  anterior  margins  have  been  described  in  a 
number  of  both  ‘labyrinthodont’  (e.g.  Milner  1980,  fig.  5)  and  ‘lepospondyf  (e.g.  Carroll  and  Gaskill  1978, 
fig.  120g)  amphibians.  No  clearly  defined  ornament  or  areas  for  clavicular  overlap  are  visible  on  the  suggested 
interclavicle  but  it  is  possibly  preserved  with  its  dorsal  surface  uppermost. 

Vertebrae 

Although  several  vertebrae  have  been  lost  from  the  articulated  series  of  G25.34  (text-fig.  la),  it  is  apparent 
from  the  impressions  on  the  counterpart  that  the  first  forty-three  vertebrae  were  originally  present.  The 
presence  of  a single  median  ventral  ridge  in  all  the  vertebrae  preserved  indicates  that  all  are  precaudal;  the 
centra  of  ai'stopod  caudal  vertebrae  are  characterized  by  a pair  of  hypapophyseal  flanges  demarcating  the 
haemal  canal  (e.g.  Zidek  and  Baird  1978).  The  vertebrae  exhibit  a gradual  increase  in  size  from  anterior  to 
posterior  of  the  series,  the  most  anterior  centra  measuring  2 mm  in  length  and  the  most  posterior  approximately 
3 mm.  Well-developed  parapophyses  are  borne  by  all  but  the  most  anterior  two  vertebrae,  in  which  they  are 
represented  only  by  scarcely  perceptible  lateral  projections  from  the  centrum.  In  vertebrae  3-10  the 
parapophyses,  as  preserved,  have  a posterolateral  orientation.  In  the  remaining  vertebrae  (1 1-43)  they  project 
laterally.  In  the  absence  of  any  indication  of  the  original  position  of  the  pectoral  girdle  it  is  impossible  to 
determine  how  many  of  the  vertebrae  are  cervicals.  However,  Baird  (1964)  has  noted  that  the  parapophyses 
of  ai'stopod  dorsal  vertebrae  change  their  orientation  from  posterolateral  to  lateral  to  anterolateral,  from 
anterior  to  posterior  of  the  column.  It  may  therefore  be  concluded  that,  with  the  exception  of  an  unknown 
number  of  cervicals,  the  vertebrae  of  G25.34  are  all  anterior-  or  mid-dorsal  in  position.  The  vertebral  count 
in  an  entire  specimen  of  O.  nanum  is  unknown.  However,  a figure  of  100+  has  been  given  by  Baird  (1964, 
p.  6)  for  a juvenile  of  O.  granulosum  Fritsch. 

The  structure  of  a typical  mid-dorsal  vertebra  of  O.  nanum  will  be  apparent  from  text-fig.  2b- d.  The  centrum 
is  holospondylous  and  probably  deeply  amphicoelous.  The  latter  is  the  ai'stopod  condition  and  is  suggested 
in  G25.34  by  the  high  degree  of  compression  undergone  by  most  centra.  Ventrally,  the  centrum  bears  a 
median  ridge  running  the  length  of  the  element  and  broadening  at  its  anterior  and  posterior  ends.  The  ventral 
ridge,  which  appears  to  have  been  mistaken  for  a low  neural  spine  by  Hancock  and  Atthey  (1868,  p.  277), 
is  flanked  by  a pair  of  elongate  depressions.  The  parapophyses  are  dorsoventrally  compressed  structures, 
slightly  expanded  distally,  and  have  their  origins  low  on  the  lateral  surface  of  the  centrum  at  approximately 
the  mid-point  of  its  length.  None  of  the  vertebrae  shows  any  evidence  of  the  presence  of  basapophyseal 
processes. 

The  vertebrae  of  G25.34  are  preserved  with  only  their  ventral  surfaces  exposed  and  yield  no  information 
on  the  structure  of  the  neural  arch.  However,  the  counterpart  slab  bears,  in  addition  to  impressions  of  most 
of  the  vertebrae  of  G25.34,  three  isolated  vertebrae  from  a more  posterior  region  of  the  trunk  (text-fig.  1/). 
Each  of  the  three  is  incomplete  but  together  they  allow  a restoration  of  most  of  the  neural  arch.  Two  are 


212 


PALAEONTOLOGY,  VOLUME  25 


almost  entire  vertebrae  with  their  dorsal  surfaces  exposed,  and  the  third  is  a very  clear  impression  of  the 
dorsal  surface  of  the  neural  arch  with  the  neural  spine  itself  remaining  in  position.  The  neural  arch  is  a long 
and  low  structure  extending  the  full  length  of  the  centrum;  whether  the  two  are  fused  or  merely  sutured 
together  cannot  be  determined.  The  neural  spine  is  very  weakly  developed  and  forms  a low  and  narrow  ridge 
running  the  length  of  the,  otherwise  almost  horizontal,  dorsal  surface  of  the  neural  arch.  The  zygapophyses 
are  widely  spaced  and  possess  horizontally  orientated  articular  surfaces.  Unfortunately,  the  nature  of  the 
available  material  makes  it  impossible  to  determine  whether  the  neural  arch  is  pierced  laterally  by  foramina 
for  the  spinal  nerves,  as  is  known  to  be  the  case  in  Phlegethontia  (McGinnis  1967)  and  at  least  one  Ophiderpeton 
species  (Baird  1964). 

Ribs 

Numerous  ribs  are  preserved  along  the  length  of  vertebral  column  of  G25.34  (text-fig.  la).  Almost  all  are 
incomplete  and  most  are  at  least  partially  obscured  by  overlying  areas  of  the  ventral  squamation.  However, 
a composite  restoration  of  rib  structure  in  O.  nanum  may  be  made  with  confidence.  Free  ribs  appear  to  have 
been  borne  by  all  but  the  most  anterior  two  of  the  forty-three  vertebrae  originally  present.  The  first  two 
vertebrae  are  also  known  to  be  without  free  ribs  in  Phlegethontia  (McGinnis  1967).  There  is  no  apparent 
variation  in  rib  form  within  the  preserved  series.  The  ribs  do,  however,  show  a slight  increase  in  size  from 
anterior  to  posterior. 

The  structure  of  a typical  precaudal  rib  of  O.  nanum  is  depicted  in  text-fig.  2a.  The  rib  is  tetraradiate,  in 
the  manner  characteristic  of  the  Ai'stopoda  (Baird  1964).  The  capitulum  is  short  and  stout,  and  appears  to 
possess  a slightly  recessed  head  for  articulation  with  the  parapophysis  of  the  vertebra.  In  all  ribs  in  which 
it  is  preserved,  the  costal  process  is  approximately  twice  the  length  of  the  capitulum  and  is  connected 
to  the  latter  for  most  of  its  length  by  a thin  ‘web’  of  bone.  Whether  the  head  of  the  costal  process  was 
recessed,  as  has  been  stated  to  be  the  case  in  Ophiderpeton  by  Baird  (1964,  p.  7),  cannot  be  determined.  The 
shaft,  which  contributes  about  two-thirds  of  the  over-all  length  of  the  rib,  is  a robust,  stiletto-like  structure 
which  tapers  distally  to  a fine  point.  A well-developed  posteromedial  process  is  present.  In  most  of  the 
preserved  ribs  the  posteromedial  process  exhibits  a slight  lateral  curvature,  towards  the  medial  side  of  the  shaft. 

Scales 

A well-developed  ventral  squamation  is  present  and,  as  preserved,  takes  the  form  of  a ‘mat’  of  closely  packed 
gastralia  extending  from  the  tenth  vertebra  to  the  forty-third  (text-fig.  la).  The  squamation  has  been  displaced 


text-fig.  2.  Restoration  of  mid-dorsal  vertebra  and  rib  of  Ophiderpeton  nanum  H & A.  a,  rib  of  left  side  in 
dorsal  view;  b-d,  vertebra  in  b,  dorsal  view;  c,  left  lateral  view  and  d,  ventral  view.  p-m.  proc.,  posteromedial 

process. 


BOYD:  CARBONIFEROUS  AMPHIBIAN  OPHIDERPETON 


213 


to  one  side  of  the  vertebral  column  in  the  anterior  part  of  the  trunk,  but  more  posteriorly  it  overlies  the 
vertebrae  and  ribs.  The  individual  gastralia  are  extremely  elongate,  almost  filamentous,  structures  (text-fig. 
\e).  In  their  form  they  somewhat  resemble  the  hair-like  gastralia  of  the  phlegethontiid  Aornerpeton  mazonense 
(Gregory)  (Gregory  1948;  Lund  1978),  but  their  tightly  packed  arrangement  and  their  extent  beyond  the 
anterior  thoracic  region  are  characteristic  of  Ophiderpeton  (Baird  1964).  As  preserved,  the  orientation  of  the 
gastralia  is  very  variable  and  Hancock  and  Atthey  (1868,  p.  277)  suggested  that  the  scales  originally  lay  with 
their  long  axes  at  90°  to  the  vertebral  column.  It  would  seem  more  likely,  however,  they  were  originally 
arranged  en  chevron,  as  in  many  other  Palaeozoic  amphibians.  Such  an  arrangement  is  still  preserved  in  some 
areas  of  the  ventral  squamation.  The  absence  of  ventral  osteoderms  in  the  ‘cervical’  region  and  beneath  the 
most  anterior  dorsal  vertebrae  is  of  some  interest  and  may  indicate,  as  Steen  (1938,  pp.  223,  224)  suggested, 
that  the  specimen  is  a sub-adult  individual. 

Although  not  altogether  absent,  as  had  previously  been  thought  (Steen  1938),  the  dorsal  squamation  is 
represented  in  G25.34  only  by  a small  number  of  rounded,  pebble-like  osteoderms  scattered  at  intervals  along 
the  length  of  the  vertebral  column  (text-fig.  la).  The  dorsal  squamation  of  Ophiderpeton  species  usually 
consists  of  numerous  such  scales  covering  the  dorsal  and  lateral  surfaces  of  the  trunk  region  and  tail  (Baird 
1964).  Its  almost  complete  absence  in  the  holotype  of  O.  nanum  may,  like  the  restricted  ventral  squamation, 
indicate  that  the  specimen  is  a juvenile  animal.  Such  an  assumption  also  offers  an  explanation  of  the  ready 
dissociation  of  the  skull  elements  subsequent  to  the  death  of  the  animal. 

DISCUSSION 

Despite  the  incompleteness  of  the  only  available  specimen  of  O.  nanum , it  is  apparent  that  this 
species  is  a member  of  the  family  Ophiderpetontidae,  as  defined  by  Baird  (1964).  Reference  to  the 
Ophiderpetontidae  is  suggested  by  the  following  characters  of  the  holotype: 

1.  The  structure  of  the  ribs,  in  which  the  shaft  is  stout  and  stiletto-like  and  a well-developed 
posteromedial  process  present.  In  the  phlegethontiids  Phlegethontia  and  Aornerpeton  the  shaft  is 
usually  slender  and  flexible  and  the  posteromedial  process  weakly  developed  or  absent  (Baird 
1964;  Lund  1978). 

2.  The  presence  of  a well-developed  ventral  squamation  composed  of  numerous,  closely  packed 
gastralia  forming  a continuous  plastron.  The  phlegethontiid  ventral  squamation  consists  of  a 
series  of  widely  spaced,  filamentous  gastralia  restricted  to  the  anterior  trunk  region  (e.g.  Gregory 
1948,  pi.  1,  fig.  4). 

3.  The  apparent  presence  of  an,  at  least  partial,  dorsal  or  lateral  squamation  of  pebble-shaped 
osteoderms;  no  dorsal  and  lateral  armour  is  known  to  occur  in  the  Phlegethontiidae  (Baird  1964). 

4.  The  form  of  the  premaxilla,  now  lost,  described  by  Steen  (1938,  p.  223).  This  was  stated  exactly 
to  resemble  the  premaxilla  of  O.  amphiuminum , described  by  Steen  in  1931  (fig.  17b-c).  The 
premaxilla  of  O.  amphiuminum  and  other  Ophiderpeton  species  differs  markedly  from  that  of 
Phlegethontia  in  possessing  an  anterior  ascending  process  (McGinnis  1967,  p.  38). 

As  presently  constituted,  the  Ophiderpetontidae  contains  only  the  type  genus,  Ophiderpeton 
Huxley  1867,  and  Coloraderpeton  Vaughn  1969.  The  latter  is  a monotypic  genus,  established  by 
Vaughn  (1969)  for  a number  of  vertebrae,  ventral  osteoderms,  and  an  attributed  rib  from  the 
Upper  Pennsylvanian  of  the  Sangre  de  Cristo  Formation  in  central  Colorado.  The  vertebrae  of 
Coloraderpeton  are  distinguished  from  those  of  both  Ophiderpeton  and  Phlegethontia  principally 
by  the  form  of  the  neural  spine.  This,  although  relatively  somewhat  higher  than  in  the  last  two 
genera,  is  restricted  in  its  anterior-posterior  extent  to  the  middle  one-third  of  the  neural  arch  and 
possesses  a crenulated  dorsal  edge.  The  dorsal  vertebrae  otherwise  appear  typically  a'istopod  and 
possess  foramina  for  exit  of  the  spinal  nerves.  Coloraderpeton  was  placed  in  the  Ophiderpetontidae 
by  Vaughn  on  the  basis  of  the  form  of  the  rib,  which  has  a stout  shaft  and  apparently  also  a 
posteromedial  process,  and  of  the  closely  packed,  fusiform  ventral  osteoderms.  It  is  clear  from  the 
structure  of  the  vertebrae  that  the  holotype  specimen  of  O.  nanum  cannot  be  placed  in  the  genus 
Coloraderpeton.  On  the  other  hand,  the  similarity  of  the  ribs,  vertebrae,  and  ventral  squamation 
to  those  of  other  described  Ophiderpeton  species  appears  to  confirm  the  correctness  of  Hancock 
and  Atthey’s  (1868)  reference  of  the  Newsham  a'istopod  to  this  genus. 


214 


PALAEONTOLOGY,  VOLUME  25 


O.  nanum  was  distinguished  by  Hancock  and  Atthey  (1868)  from  the  only  previously  described 
Ophiderpeton  species,  O.  brownriggii  Huxley  1867,  by  its  relatively  small  size  and  the  almost 
filamentous  nature  of  the  ventral  osteoderms.  The  former  is  clearly  unreliable  as  a taxonomic 
criterion,  especially  in  view  of  the  probability,  noted  above,  that  the  O.  nanum  holotype  represents 
a sub-adult  individual.  The  significance  of  the  second  character  is  uncertain.  None  the  less,  in  view 
of  the  almost  complete  absence  of  a skull  in  the  holotype  specimen  and  the  lack  of  adequately 
detailed  descriptions  of  other  Ophiderpeton  species,  it  would  seem  advisable  to  retain  O.  nanum  as 
a distinct  species,  at  least  pending  thorough  revision  of  the  Ophiderpetontidae. 

The  possible  presence  of  an  interclavicle  in  O.  nanum  is  a feature  of  some  importance.  Steen 
(1931,  p.  876)  identified  two  bones  in  a specimen  of  O.  amphiuminum  from  the  Westphalian  D of 
Linton,  Ohio,  as  clavicle  and  cleithrum,  but  these  were  subsequently  reinterpreted  by  Baird  (1964) 
as  possibly  representing  hyoid  elements.  Most  recent  workers  have  concurred  in  regarding  described 
a'istopod  species  as  lacking  both  limbs  and  limb  girdles.  However,  Goin  and  Goin  (1971,  p.  67) 
have  pointed  out  that  forelimbs  are  not  known  for  most  fossil  members  of  the  urodele  family 
Sirenidae,  despite  their  presence  in  the  three  extant  species  of  this  group.  They  suggested,  therefore, 
that  the  A'istopoda  may  possibly  also  have  possessed  vestigial  forelimbs  which,  together  with  the 
pectoral  girdle,  were  usually  lost  prior  to  fossilization.  If  the  element  tentatively  identified  as  an 
interclavicle  in  the  holotype  of  O.  nanum  is  correctly  so  interpreted,  its  presence  at  least  partially 
confirms  the  above  hypothesis. 

Acknowledgements.  My  thanks  are  due  first  to  Mr.  A.  M.  Tynan,  Curator  of  the  Hancock  Museum,  for 
permission  to  describe  the  specimens  in  his  care.  I also  wish  to  thank  Miss  Susan  Turner  (Hancock  Museum), 
Dr.  A.  L.  Panchen  (University  of  Newcastle  upon  Tyne),  and  Dr.  A.  C.  Milner  (British  Museum  (Natural 
History))  for  helpful  comments  and  discussion. 


REFERENCES 

baird,  d.  1964.  The  a'istopod  amphibians  surveyed.  Breviora,  No.  206,  1-17. 

CARROLL,  r.  l.  and  GASKILL,  p.  1978.  The  Order  Microsauria.  Mem.  Am.  phil.  Soc.  126,  1-211. 
goin,  c.  J.  and  coin,  o.  b.  1971.  Introduction  to  herpetology , 2nd  edn.  W.  H.  Freeman,  San  Francisco.  353  pp. 
Gregory,  J.  t.  1948.  A new  limbless  vertebrate  from  the  Pennsylvanian  of  Mazon  Creek,  Illinois.  Am.  J.  Sci.  246, 
636-663. 

HANCOCK,  A.  and  ATTHEY,  t.  1868.  Notes  on  the  remains  of  some  reptiles  and  fishes  from  the  shales  of  the 
Northumberland  coal  field.  Ann.  Mag.  nat.  Hist.  (4),  1,  266-278,  346-378. 

HUXLEY,  t.  h.  1867.  In  HUXLEY,  t.  h.  and  WRIGHT,  E.  p.  On  a collection  of  fossil  Vertebrata  from  the  Jarrow 
Colliery,  County  of  Kilkenny,  Ireland.  Trans.  R.  Ir.  Acad.  24,  351-368. 
land,  d.  h.  1974.  Geology  of  the  Tynemouth  district.  Mem.  geol.  Surv.  Gt  Br.  15,  61-62. 
lund,  R.  1978.  Anatomy  and  relationships  of  the  Family  Phlegethontiidae  (Amphibia,  A'istopoda).  Ann. 
Carnegie  Mus.  47  (4),  53-79. 

mcginnis,  h.  j.  1967.  The  osteology  of  Phlegethontia,  a Carboniferous  and  Permian  a'istopod  amphibian.  Univ. 
Calif.  Pubis  geol.  Sci.  71,  1-46. 

MILNER,  a.  r.  1978.  A reappraisal  of  the  early  Permian  amphibians  Memonomenos  dyscriton  and  Cricotillus 
brachydens.  Palaeontology , 21,  667-686. 

— 1980.  The  temnospondyl  amphibian  Dendrerpeton  from  the  Upper  Carboniferous  of  Ireland.  Ibid.  23, 
i25- I41 . r . 

steen,  m.  c.  1931.  The  British  Museum  collection  of  Amphibia  from  the  Middle  Coal  Measures  of  Linton,  Ohio. 
Proc.  zool.  Soc.  Lond.  1930,  849-892.  . 

1938.  On  the  fossil  Amphibia  from  the  Gas  Coal  of  Nyrany  and  other  deposits  in  Czechoslovakia.  Ibid.  (B) 

108,  205-284. 

vaughn,  p.  p.  1969.  Upper  Pennsylvanian  vertebrates  from  the  Sangre  de  Cristo  Formation  of  central  Colorado. 
Contr.  Sci.  Los  Angeles  Mus.  164,  1-28. 

zidek,  j.  and  baird,  d.  1978.  Cercariomorphus  Cope,  1885,  identified  as  the  a’istopod  amphibian  Ophiderpeton. 
J.  Paleont.  52,  561-564. 

MICHAEL  J.  BOYD 
Department  of  Natural  History 
Kingston  upon  Hull  Museum 

Typescript  received  3 August  1980  Queen  victoria  Square 

Revised  typescript  received  15  October  1980  Kingston  upon  Hull,  North  Humberside 


A NEW  SPECIES  OF  THE  LYCOPSID  PLEUROMEIA 
FROM  THE  EARLY  TRIASSIC  OF  SHANXI, 
CHINA,  AND  ITS  ECOLOGY 

by  wang  ziqiang  and  WANG  lixin 


Abstract.  A new  species,  Pleuromeia  jiaochengensis,  is  recorded  from  the  early  Triassic  of  Jiaocheng  district 
in  Shanxi  (Shansi)  Province,  China.  Its  small  size,  morphological  features  of  the  strobilus  and  sporophylls, 
abortive  leaves,  and  the  undeveloping  rhizophore  separate  this  from  all  other  species.  The  succulent  sporophylls 
may  be  a major  area  of  photosynthesis.  Based  on  lithology  and  distance  from  known  marine  strata,  an  inland 
desert  environment  is  suggested  for  this  new  species.  Its  stratigraphic  significance  is  discussed. 

It  is  well  known  that  Palaeozoic  floras  flourished  during  Upper  Carboniferous  and  Permian  time, 
but  suffered  from  a worldwide  arid  climate  at  the  end  of  the  Permian.  Only  a few  relics  survived 
into  the  early  Triassic  where  they  faced  severe  climatic  conditions.  One  such  genus  is  Pleuromeia 
which  is  regarded  as  an  important  early  Triassic  index  fossil,  and  thought  to  be  a link  between 
the  Palaeozoic  Sigillaria  and  modern  Isoetes. 

Pleuromeia  was  first  recorded  from  the  Bunter  Sandstone  in  Germany  nearly  a century  ago. 
Prior  to  1960,  apart  from  a few  records  from  France,  Spain,  and  the  eastern  U.S.S.R.  all  the 
records  were  from  Germany.  In  particular,  Magdefrau  (1930,  1931)  provided  much  information 
about  the  genus  from  a large  number  of  German  specimens. 

Since  the  1960s  new  material  has  been  recorded  from  the  early  Triassic  in  the  Soviet  Union, 
Japan,  and  China.  Pleuromeia  rossica  was  described  by  Neuberg  (1960u)  from  the  Russian  Platform, 
whilst  Krassilov  and  Zakharov  (1975)  added  further  information  on  the  genus  from  material  in 
the  far  eastern  part  of  the  Soviet  Union.  Kon’no  (1973)  described  P.  hatai  from  the  Scythian  in 
north-eastern  Japan,  and  in  China,  Wang,  Xie,  and  Wang  (1978)  described  well-preserved  specimens 
of  P.  sternbergi  (Munster)  Corda  and  P.  rossica  Neuberg  from  the  early  Triassic  of  the  Qinshui 
Basin  in  Shanxi.  The  genus  Pleuromeia  therefore  has  a widespread  distribution.  It  is  worth  noting 
that  from  the  early  Triassic  of  eastern  Australia,  Cylostrobus  a lycopsid-like  strobilus  which  shows 
some  similarity  to  Pleuromeia  of  the  Northern  Hemisphere,  has  recently  been  regarded  by  Retallack 
(1975)  as  Pleuromeia  longicaulis  (Burgess). 

After  the  discovery  of  P.  sternbergi  and  P.  rossica  from  the  early  Triassic  Heshankou  Formation 
in  Qinshui  Basin,  Shanxi  Province,  the  present  writers  found  many  well-preserved  specimens 
belonging  to  a new  species  of  Pleuromeia  in  the  Luijiakou  Formation  beneath  the  Heshankou 
Formation  in  Jiaocheng  district  (text-fig.  1),  not  far  from  the  famous  Xuan-zhong  Temple  of  Shanxi 
Province.  This  paper  describes  these  specimens. 


FOSSIL-BEARING  STRATA 

The  Luijiakou  Formation  in  Jiaocheng  district  is  the  western  extension  of  the  sandstone  beds  of  the 
‘Shischienfeng  Series’  at  West  Hill  in  Taiyuan,  and  consists  mainly  of  reddish-purple,  fine-grained  sandstones. 
According  to  data  supplied  by  the  Regional  Geological  Survey  of  Shanxi  Province,  its  thickness  in  Jiaocheng 
district  exceeds  460  m,  the  Formation  being  subdivided  into  three  parts.  The  upper  part,  consists  of  108  m 
of  grey  and  reddish-purple  fine-grained  feldspathic  sandstones  intercalated  with  reddish-purple  siltstones  and 
sandy  shales.  The  middle  part  is  composed  of  215  m of  red,  grey  to  greyish-purple  fine-grained  feldspathic 
sandstones  interbedded  with  reddish-purple  silty  shales  and  shales.  The  shales  contain  abundant  ripple  marks. 


IPalaeontology,  Vol.  25,  Part  1,  1982,  pp.  215-225,  pis.  23-24.| 


216 


PALAEONTOLOGY,  VOLUME  25 


sun-cracks,  and  cross-bedded  structures  sometimes  intercalated  with  a few  lenses  of  greyish-white  sandstones 
containing  fossil  plants  and  greyish-green  shales  with  fossil  estherians.  The  lower  part,  consists  of  138  m of 
reddish-grey  and  purplish-grey  fine-grained,  cross-bedded  feldspathic  sandstones  containing  abundant 
laminations  of  magnetite,  intercalated  with  lenses  of  greyish-white  feldspathic  sandstones.  The  fossil  plants, 
which  include  Pleuromeia,  are  within  the  middle  part  of  the  Formation  approximately  130  m above  its  base 
at  four  localities  (numbered  Z01-Z04)  near  Jaoertou  village,  in  the  north-western  mountainous  area  of 


text-fig.  1.  Map  showing  region  of  the  fossiliferous  localities.  Inset  shows  eastern  half  of  China. 


EXPLANATION  OF  PLATE  23 

All  figures  x 1 unless  otherwise  stated.  All  specimens  from  the  Luijiakou  Formation,  Shanxi  Province. 

Figs.  1-13.  Pleuromeia  jiaochengensis  sp.  nov.  1-2;  6-11.  Various  strobili;  1,  Z01-021;  2,  showing  the  awl-like 
leaves  on  the  part  of  stem  beneath  the  strobilus,  Z01-01;  6,  Z01-024;  7,  Z01-027;  8,  Z01-019;  9,  Z01-022;  10, 
Z01-019;  11,  syntype,  a strobilus  containing  megaspores  (at  a),  Z01-061.  3-5,  complete  young  plants;  3, 

syntype,  Z0 1-020;  4,  the  same  x4,  5,  Z0 1-061.  12.  Shows  a decorated  stem,  with  short  surface  ridges, 
Z0 1-061.  13.  Syntype,  stem  surface  showing  sparse,  faint  leaf-scars,  Z0 1-206,  x2. 


PLATE  23 


WANG  and  WANG,  Pleuromeia 


218 


PALAEONTOLOGY,  VOLUME  25 


Jiaocheng  district.  At  locality  Z01,  there  are  many  well-preserved  specimens  of  strobili,  stems,  rhizophores, 
and  even  some  complete  plants  in  a lens  of  greyish-white  sandstone,  0-2  m thick  and  2 m across.  At  Z04, 
there  are  many  sporophylls,  stems,  and  rhizophores  occurring  in  larger  lenses.  In  addition,  at  localities  Z02 
and  Z03,  there  are  specimens  tentatively  included  within  the  genera  Crematopteris  sp.  Phyllotheca  sp. 
Taeniopteris  sp.  Neocalamites  sp.  by  the  present  authors.  All  specimens  illustrated  in  this  paper,  are  reposited 
in  the  Tianjin  Institute  of  Geology  and  Mineral  Resources,  Ministry  of  Geology. 

SYSTEMATIC  PALAEOBOTANY 

Family  pleuromeiaceae 
Genus  pleuromeia  Corda,  1852 
Pleuromeia  jiaochengensis  sp.  nov. 

Plates  23,  24,  text-figs.  2-3 

Syntypes.  Z0 1-020  (small  complete  plant);  Z0 1-061  (showing  strobilus  and  megaspores;  Z0 1-206  (showing 
stem  surface);  Z04-159  (showing  rhizophore). 

Diagnosis.  A lycopsid  of  small  shrub  size,  probably  dioecious,  generally  20-30  cm  high  but  may 
reach  almost  50  cm  (see  text-fig.  2,  the  reconstruction  of  a whole  plant).  Stem  erect,  unbranched, 
maximum  diameter  1-5  cm,  terminating  in  a large  strobilus  measuring  one-quarter  to  one-third  of 
the  plant  height  (PI.  23,  figs.  3-4).  Surface  of  stem  covered  with  sparse,  faint,  at  times  barely  visible 
leaf-scars,  usually  smooth,  with  small  obscure  pits  where  the  leaves  were  originally  attached. 
Decorticated  stems  with  short  ridges  (PI.  23,  fig.  12).  Leaf-scar  lens-shaped,  obscure  in  outline, 
with  a central  pit  (PI.  23,  fig.  13).  Leaves  awl  or  spine-like,  2-3  mm  in  length,  about  1 cm  apart 
on  the  upper  portion  of  the  stem  (PI.  24,  fig.  6),  having  dehisced  from  the  lower  part.  Rhizophores 
tuberous  when  mature,  covered  with  oblong  to  oval,  well-separated  appendage-scars;  the  younger 
rhizophores  only  slightly  swollen  at  the  base  of  the  stem  and  covered  with  rather  thick  appendages, 
maximum  length  3 cm,  usually  on  the  lower  part.  Strobilus  probably  unisexual,  narrowly 
spike-shaped,  maximum  length  20  cm.  Sporophylls  spirally  arranged,  imbricate,  4-5  in  each  spiral 
(PI.  23,  fig.  6),  the  spiral  from  60°  to  quite  an  acute  angle  to  the  rachilla,  giving  a total  of  more 
than  100  sporophylls  in  a mature  strobilus.  Sporophylls  spatulate  with  a sagittate  apex,  longer 
than  wide,  ovate  to  oblong  in  outline,  adaxially  concave  where  the  sporangium  lies.  The  border 
of  the  sporophyll  is  wider  anteriorly  than  laterally,  and  in  mature  strobili  the  apex  of  the  anterior 
border  of  larger  sporophylls  is  slightly  reflexed  upwards  (PI.  24,  fig.  3;  text-fig.  3d  left),  and  is 
probably  visible  on  the  outside  of  the  strobilus.  Where  the  upward  reflexed  part  of  a sporophyll  is 
truncated,  it  clearly  shows  a retuse,  small  anterior  border  (PI.  24,  fig.  5).  Sporangium  discoid,  oval  or 
orbicular  in  outline,  attached  to  the  rachilla  only  at  its  proximal  end.  Surface  of  sporangium  with 
many  more  or  less  parallel  lines  (PI.  24,  fig.  4)  which  converge  slightly  at  both  ends.  Thickness  of 


EXPLANATION  OF  PLATE  24 

All  figures  x 1 unless  otherwise  stated.  All  specimens  from  the  Luijiakou  Formation,  Shanxi  Province. 

Figs.  1-15.  Pleuromeia  jiaochengensis  sp.  nov.  1 -2.  Part  of  strobili  containing  megaspores  (at  a).  1 , Z0 1 —0 1 , 

x 2.  2,  an  enlargement  of  syntype,  PI.  21,  fig.  11,  showing  megaspores  associated  with  nearly  all  the 

sporophylls.  3-5.  Sporophylls  and  sporangia.  3,  two  sporophylls,  the  right  one  with  upward  reflexed  apex, 
Z01-08  x 2;  4,  sporangia  near  base  of  a strobilus,  showing  the  parallel  lines  on  their  surface,  Z0 1-042;  5,  a 
larger  sporophyll  having  lost  its  reflexed  apex,  Z0 1-061.  6.  Part  of  a stem,  showing  spine-like  leaves, 

Z0i-00.  7.  Enlargement  of  a sporangium  containing  megaspores  (a),  from  the  strobilus  of  PI.  21,  fig.  11, 

Z01-019,  x 10.  8.  A megaspore,  Z01-022,  x 50.  9-15.  Various  rhizophores;  9,  10,  showing  appendage 

scars,  Z01-060,  Z01-135;  1 1,  on  the  right,  a young  rhizophore  showing  a few  appendages  on  the  lower  part, 
Z0 1-054;  12,  the  same  x 2;  13-15,  showing  appendages  on  the  lower  part,  and  appendage  scars  on  the  upper 
part;  13,  Z01-027;  14,  Z04-151;  15,  syntype,  Z04-159. 


PLATE  24 


WANG  and  WANG,  Pleuromeia 


220 


PALAEONTOLOGY,  VOLUME  25 


TEXT-FI 
sis  sp.  I 


..■C  A 


r:;---w.v*’  iS  v v 


i®S# 

->y: 

'••  '■<  5 " .>■' 


d e c ortic  ate  d 
stem 


,^>:c  & 

11 

Fife 

#15  fefc 


G.  2.  Pleuromeia  jiaochengen- 
iov.  Reconstruction  of  a whole 
plant,  x 1. 


sporangium  decreases  gradually  from  proximal  to  distal  end, 
resulting  in  an  aggregation  of  spores  at  the  proximal  end  near 
the  rachilla.  Megaspores  tetrad,  spherical  or  triangular,  laevi- 
gate  or  granulate,  300-500  p.m  in  diameter.  Trilete  mark 
distinct,  shorter  than  spore  radius.  Microspores  unknown. 

Discussion  and  comparison.  The  single,  erect,  terminal  strobilus 
terminating  on  unbranched  woody  stem,  are  characters  asso- 
ciated with  the  genus  Pleuromeia  rather  than  Selaginella  or 
Isoetes.  Though  the  sporophylls  of  this  new  species  have  an 
elongate  anterior  border  similar  to  those  in  Selaginella , they 
differ  from  it  in  lacking  elongate  acuminate  distal  tips 
(Retallack  1975).  Also,  the  thick,  spatulate  shape  is  quite  unlike 
the  sporophyll  of  Selaginella.  In  addition,  Selaginella  is 
herbaceous  with  a dichotomous  stem  and  dimorphic  leaves 
often  in  four  rows  (Chaloner  1967;  Smith  1955),  quite  unlike 
those  of  P.  jiaochengensis. 

Isoetes,  which  is  usually  considered  a close  relative  of 
Pleuromeia,  lacks  a definite  stem  from  the  lobed  rhizophore, 
and  also  lacks  a distinct  strobilus.  Unlike  Pleuromeia,  the 
sporophylls  of  Isoetes  are  placed  within  the  centre  of  a cluster 
of  leaves. 

In  the  past,  when  identifying  species  of  Pleuromeia,  authors 
have  placed  considerable  emphasis  on  features  of  the  leaf-scars 
on  the  stem  surface.  During  the  past  two  decades,  more  recent 
finds  have  greatly  increased  knowledge  of  the  reproductive 
organs;  for  example,  the  strobilus,  sporophylls,  sporangia, 
megaspores,  and  microspores.  These  are  now  used  much  more 
in  identification.  Also,  Dobruskina  (1974)  pointed  out  that  the 
leaf-scars  on  specimens  assigned  to  Pleuromeia  and  Pleuro- 
mopsis  by  Brick  (1936)  and  Sixtel  (1962)  respectively,  from  the 
late  Triassic  of  Central  Asia,  are  not  from  these  genera,  and 
that  the  leaf-scars  of  Pleuromeia  vary  in  shape  depending 
on  the  degree  of  decortication  prior  to  burial.  Some  specimens 
from  the  late  Permian  of  the  Petchora  Basin  in  the  Soviet 
Union  included  by  Neuberg  (19606,  pi.  5,  fig.  2;  pi.  14, 
right)  in  the  genus  Viatscheslavia  have  similar  leaf-scars 
to  those  in  Pleuromeia.  Hence  leaf-scar  features  are  not 
now  considered  as  important  as  in  the  past  in  identifying 
Pleuromeia. 

Of  the  six  previously  described  species  of  Pleuromeia  from 
the  Northern  Hemisphere,  only  three,  namely  P.  sternbergi, 
P.  rossica,  and  P.  hatai,  have  been  found  in  sufficient  numbers 
and  are  well  preserved  enough  to  give  knowledge  of  the 
complete  plant.  Of  the  rest,  there  is  not  enough  real  evidence 
for  their  assignation  as  distinct  species.  P.  oculina  (Blancken- 
horn,  1886)  lacks  reproductive  organs,  P.  olenekensis 
(Krassilov  and  Zakhalov,  1975)  is  too  incomplete  to  compare 
with  other  species,  and  the  designation  of  P.  obrutschewi  Elias  is 
in  dispute  (Krassilov  and  Zakhalov  1975). 

P.  jiaochengensis  differs  from  all  previously  described  species, 
by  its  small  size,  relatively  large  strobilus  with  elongate 


WANG  AND  WANG:  PLEUROME1A  JIAOCHENGENSIS 


221 


sporophylls,  and  its  undeveloping  rhizophore  with  appendage-scars  which  are  less  in  number,  and 
larger.  P.  jiaochengensis  is  usually  20-30  cm  high,  with  a maximum  height  of  under  50 
cm.  This  compares  with  2 m in  P.  sternbergi  and  1 m in  P.  rossica.  The  smallest  specimens  of  P. 
hatai  may  be  of  similar  size. 

The  strobilus  of  P.  jiaochengensis  is  larger  in  proportion  to  the  whole  plant  than  in  other  species, 
except  P.  hatai.  The  strobilus  of  P.  hatai,  however,  differs  from  P . jiaochengensis  in  being  cylindrical 
rather  than  spike-shaped.  Also,  in  P.  hatai  there  are  more  sporophylls  in  a strobilus,  and  they 
are  smaller  in  size. 

The  sagittate  shape  of  the  sporophyll  of  P.  jaiochengensis  is  slightly  different  from  that  in  all 


text-fig.  3.  Pleuromeia  jiaochengensis  sp.  nov.  a,  complete  young  plant,  x 2;  b,  a strobilus  with 
megaspores,  x 2;  c,  a strobilus,  x 1;  d,  two  sporophylls,  the  left  one  with  an  upward  reflexed  apex, 
x2;  e,  a sporangium  containing  megaspores,  x 10;  /,  a megaspore,  x 50;  g,  a rhizophore  covered 
with  appendage-scars,  x 1 ; h,  a stem  surface  with  a few  leaf-scars,  x 2. 


222 


PALAEONTOLOGY,  VOLUME  25 


other  species.  Plate  24,  fig.  5,  right,  shows  a larger  sporophyll  of  P . jiaochengensis,  its  upward  reflexed 
tip  having  been  truncated,  exposing  a retuse  anterior  margin  which  differs  from  P.  rossica,  in  having  a 
notched  margin  (PI.  24,  fig.  5).  In  P.  sternbergi  the  mature  sporophylls  also  have  an  elongate  outline, 
but  unlike  P . jaiochengensis  the  lateral  and  anterior  margins  are  of  approximately  equal  width. 

Although  the  nature  of  the  sporangium  of  P.  jaiochengensis  is  not  as  yet  clear,  the  uniform 
megaspores  are  different  from  those  in  P.  rossica,  but  similar  to  those  in  P.  sternbergi  and  P.  hatai. 
However,  unlike  these  two  species,  the  sporangium  has  a markedly  thickened  proximal  end  (PI. 
24,  fig.  7;  text-fig.  3e),  which  contains  a mass  of  megaspores. 

The  awl  or  spine-like  leaves  of  P.  jaiochengensis  is  one  of  the  most  important  characters  in 
comparing  this  with  other  species.  No  clear  leaf-scars  are  present  after  the  leaves  have  dehisced.  The 
leaves  of  P.  sternbergi  and  P.  hatai  are  linear  in  outline  and,  after  dehiscence,  leave  obvious  leaf-scars 
on  the  outer  stem  surface.  The  leaves  of  P.  rossica  are  probably  scale-like  (Dobruskina  1974),  and 
rather  similar  to  those  in  P.  jaiochengensis,  but  when  dehisced  they  leave  densely  covered  and 
distinctive  leaf-scars. 

An  undeveloping  rhizophore  is  also  regarded  by  the  present  writers  as  an  important  character 
for  distinguishing  the  new  species.  Its  appendage  scars  are  larger  and  less  in  number  than  other 
species.  In  addition,  unlike  the  other  species,  the  rhizophore  has  neither  horns  nor  suture  lines 
(kliifte)  on  the  lower  surface. 

When  describing  Pleuromeia  from  the  Qinshui  basin  of  Shanxi,  Wang  et  al.  (1978)  noted  that 
the  difference  between  the  strobili  of  P.  sternbergi  and  P.  rossica  was  so  great  that,  as  pointed  out 
by  Kon’no  (1974),  they  might  represent  different  genera.  With  regard  to  this,  P.  jaiochengensis 
throws  more  light  on  P.  sternbergi  based  on  its  unisexual  strobilus  without  a pedicel,  and  on  the 
character  of  the  sporophylls. 

Retallack  (1975)  transferred  the  lycopsid-like  strobilus  Cylostrobus  (originally  designated  by 
Helby  and  Martin  1965)  which  had  previously  been  considered  more  like  the  palaeozoic  lycopsids,  to 
Pleuromeia  longicaulis  (Burgess).  The  specimens  are  from  the  early  Triassic  of  eastern  Australia,  and 
is  the  first  record  of  this  genus  in  the  Southern  Hemisphere.  This  type  of  strobilus  shows  some 
important  features,  for  example,  it  is  bisexual  (monoecious),  the  sporophylls  have  a remarkable 
ribbed  keel-like  apex  on  their  dorsal  surface,  and  the  microspores  are  monolete  and  covered  with 
dense  spines.  All  these  features  are  strikingly  different  from  Northern  Hemisphere  Pleuromeia 
species,  including  P.  jaiochengensis.  The  basis  for  the  designation  of  the  Australian  species  given  by 
Retallack,  mainly  results  from  an  analysis  of  their  environment  and  a few  associated  fossil  stems.  In 
fact,  there  are  two  types  of  stems  associated  with  Cylostrobus  sydneyensis  from  the  upper  Narrabeen 
Group;  one  bearing  a stigmaria-like  stem-base,  and  the  other  with  branchlets  (Helby  and  Martin 
1965,  p.  399,  pi.  1 , fig.  6)  more  like  a palaeozoic  lycopsid  than  Pleuromeia.  The  present  authors  are  not 
as  yet  convinced  that  this  Australian  species  should  be  included  within  the  genus  Pleuromeia. 
Recently,  Ash  (1979)  has  described  Skilliostrobus,  a new  type  of  lycopsid  strobilus  also  from  the  early 
Triassic  of  Australia,  and  this  is  similar  to  Cylostrobus  in  having  bisexual,  monoecious  features. 


ECOLOGICAL  PROBLEMS  CONCERNING  P.  JAIOCHENGENSIS 

Pleuromeia  was  originally  regarded  as  a desert  xerophyte  similar  to  the  present-day  Cactaceae. 
More  recently,  linking  the  plant-bearing  beds  with  associated  sediments  containing  dolomite  and 
halite,  together  with  marine  fossil  animals,  a marginal  marine  habitat  was  suggested.  Magdefrau 
(1931)  was  first  to  suggest  an  halophytic  habitat  for  Pleuromeia.  Hirmer  (1933),  Neuberg  (19606), 
and  Kon’no  (1973)  have  supported  this  view,  and  more  recently  Krassilov  and  Zakharov  (1975) 
and  Retallack  (1975)  have  suggested  a mangrove-type  environment  along  the  sea-shore,  or 
marginally  on  deltas.  This  type  of  habitat  is  not  envisaged  for  P.  jiaochengensis.  Unlike  the  records 
from  western  Europe,  Siberia,  and  Japan,  no  marine  animal  fossils  have  been  recorded  from  the 
strata  containing  P.  jiaochengensis  in  Jiaocheng  district.  Also,  there  is  no  evidence  for  lagoonal 
sediments,  with  the  absence  of  dolomite  and  halite,  and  only  a few  doubtful  reports  of  gypsum. 


WANG  AND  WANG:  PLEUROMEIA  JIAOCHENGENSIS 


223 


According  to  the  geological  exploratory  reports  of  the  Shanxi  Regional  Geological  Survey  Team 
the  Formation  in  this  area  contains  a lot  of  unweathered  minerals  such  as  feldspar  and  magnetite. 
Cross-bedding,  ripple  marks,  and  sun-cracks  are  relatively  common,  and  indicate  an  arid  climate. 
The  ‘Shischienfeng  Series’  with  marine  intercalated  beds  are  known  only  from  Linyuao,  and 
Tongchuan  districts  of  north  Shanxi  (Shengsi),  and  are  at  least  300  km  from  the  Jiaocheng  district 
(see  text-fig.  1 and  Yin  and  Lin  1979).  We  therefore  consider  P.  jiaochengsis  to  be  a small  plant 
growing  near  desert  oases.  As  all  specimens  are  found  in  greyish-green  sandstones,  the  new  species 
might  grow  very  near  or  even  partly  in  the  water  bodies.  Its  small  size  also  suggests  arid  conditions; 
the  undeveloped  and  abortive  leaves  probably  resulted  from  extreme  transpiration  in  an  arid  climate. 
The  appendages  are  stout  and  strong,  probably  to  withstand  desert  conditions.  Further  evidence 
for  an  arid  environment  comes  from  the  Estheridae  which  are  recorded  from  the  shales  intercalated 
with  the  plant-bearing  sandstones.  This  type  of  small  fauna  is  often  found  in  temporary  inland 
bodies  of  water. 

Plate  23,  figs.  3-4,  show  a small  but  complete  plant  which  is  only  3 cm  high.  The  stem  is  almost 
smooth,  the  large  strobilus  is  well  developed  at  the  apex,  and  has  a few  distinctive  appendages  at  its 
base.  The  appendages  are  plano-concave  and  fleshy,  and  were  probably  succulent.  They  are  more 
developed  than  the  sterile  leaves,  and  may  well  have  carried  out  much  of  the  photosynthesis. 


TABLE  1.  The  stratigraphic  sequence  of  the  ‘Shischienfeng  Series’  in  Shanxi. 


Early  Triassic 

Induan  Olenekian 

‘Shischienfeng  Series’ 

Heshankou  Formation 

Pleuromeia  sternbergi 
P.  rossica 

Luijiakou  Formation 

P.  jiaochengensis 

Later  Permian 

Tartarian 

Sunjiakou  Formation 

Shihtienfenia  permica 

224 


PALAEONTOLOGY,  VOLUME  25 


STRATIGRAPHIC  OCCURRENCE  OF  THE  GENUS  PLEUROMEIA, 
AND  THE  AGE  OF  P.  JIAOCHENGENSIS 

In  Germany  the  Pleuromeia- bearing  beds  have  a rather  short  vertical  range,  extending  from  Middle  to  Upper 
Bunter  (i.e.  from  Bausandstein  to  Chirotheriumsandstein).  The  age  according  to  Lozovsky,  Movschovich, 
and  Mimich  (1973)  is  considered  to  be  equivalent  to  the  Tirolites-Columbites  ammonite  zone  of  Middle  and 
Upper  Olenekian  age,  despite  the  lack  of  the  ammonite  evidence.  At  Russian  Island  near  Vladivostock  and 
north-east  Japan,  Pleuromeia  occurs  in  rocks  of  early  Triassic  age  dated  on  the  ammonites.  On  the  Russian 
Platform,  P.  rossica  occurs  in  the  Ribinsk  Formation  together  with  the  amphibian  Benthosuchus,  and  is  dated 
as  early  Triassic.  According  to  Lozovsky  et  al.  (1973),  this  Formation  is  equivalent  to  the  Owenites  ammonite 
zone,  and  may  be  referred  to  the  early  Olenekian. 

At  Mangeschlack  near  the  Caspian  sea,  P.  sternbergi  occurs  with  marine  lamellibranchs  which  also  occur 
in  the  Upper  Bunter  Sandstone  in  Germany  (Dobruskina  1974)  and  so  are  dated  as  late  Olenekian.  The 
Pleuromeia  localities  in  the  Vosges,  at  Mangeschlack  in  Central  Asia,  and  on  Russian  Island  in  the  Far  East, 
are  suggested  by  some  authors  as  being  as  late  as  Middle  Triassic.  However,  in  the  present  writers’  opinion, 
the  exact  age  is  in  doubt,  through  lack  of  detailed  information. 

The  Luijiakou  Formation  in  central  Shanxi  is  thought  to  be  equivalent  to  the  middle  part  of  the  so-called 
‘Shischienfeng  Series’.  For  many  years,  no  fossils  were  found,  and  the  age  was  in  doubt,  but  thought  to  be  late 
Permian  to  early  Triassic,  depending  on  the  exact  age  of  the  ?late  Permian  reptile  Shihtienfenia  permica  Young, 
found  in  the  underlying  Sunjiakou  Formation  (see  Table  1). 

Recently  Wang,  Xie,  and  Wang  (1978)  described  P.  sternbergi  and  P.  rossica  from  the  overlying  Heshankou 
Formation  and  suggested  an  early-middle  Olenekian  age  for  that  Formation.  The  discovery  of  P.jiaochengensis 
in  the  Luijiakou  Formation  lends  further  support  to  the  above  age  determination.  In  addition,  another  of 
the  fossil  plants  associated  with  P.  jiaochengensis,  is  Crematopteris  sp.  This  species  is  confined  to  the  Bunter 
Sandstone  in  western  Europe.  If  the  age  of  the  Heshankou  Formation  containing  P.  rossica  is  referred  to  the 
Olenekian  stage,  or  is  equivalent  to  the  Ribinsk  Formation  on  the  Russian  Platform,  then  the  age  of  the 
Luijiakou  Formation  can  accurately  be  assigned  to  the  Induan  of  early  Triassic  age. 


Acknowledgements.  Sincere  thanks  are  due  to  Professor  X.  X.  Lee  of  Nanjing  Institute  of  Geology  and 
Palaeontology,  Academia  Sinica,  for  his  encouragement. 


REFERENCES 

ash,  s.  R.  1979.  Skilliostrobus  gen.  nov.  a new  Lycopsid  cone  from  the  Early  Triassic  of  Australia.  Alcheringa,  3, 
73-79. 

blanckenhorn,  m.  1886.  Die  fossilen  Flora  des  Buntsandsteins  und  des  Muschelkalke  der  Umgegend  von 
Commern.  Palaeontographica,  32,  117-154. 

brick,  m.  M.  1936.  The  first  discovery  of  a Lower  Triassic  flora  in  Central  Asia.  Trans.  Geol.  Inst.  Acad.  Sci.  5, 
461-474.  [In  Russian.] 

chaloner,  w.  g.  1967.  In  Traite  de paleobotanique.  Tome  II.  Bryophyta,  Psilophyta,  Lycophyta.  Ed.  boureau,  e. 
Masson  et  Cie,  Paris.  Pp.  1-802. 

dobruskina,  I.  a.  1974.  Triassic  lycopods.  Palaeont.  Z.  111-124.  [In  Russian.] 

helby,  r.  and  martin,  a.  r.  h.  1965.  Cylostrobus  gen.  nov.  cones  of  Lycopsidean  plants  from  the  Narrabeen 
Group  (Triassic)  of  New  South  Wales.  Aust.  J.  Bot.  13,  384-404. 
hirmer,  m.  1933.  Rekonstruktion  von  Pleuromeia  sternbergi  (Muster)  Corda,  Nebst  Bermerkungen  zur 
morphologie  der  Lycopodiales.  Palaeontographica , B 78,  47-56. 
kon’no,  e.  1973.  New  species  of  Pleuromeia  and  Neocalamites  from  the  upper  Scythian  bed  in  the  Kitakami 
Massif.  Sci.  Rep.  Res.  Insts  To  hoku,  43,  2nd  ser.  (Geol.),  97-115. 
krassilov,  v.  a.  and  zakharov,  d.  1975.  Pleuromeia  from  the  Lower  Triassic  of  the  Far  East  of  the  USSR.  Rev. 
Palaeobot.  Palynol.  19,  221-232. 

lozovsky,  v.  r.,  mavschovich,  e.  v.  and  mimich,  m.  t.  1973.  On  the  stratigraphy  of  Early  Triassic  sediments  on 
the  Russian  Platform.  Proc.  Acad.  Sci.  USSR,  ser.  Geol.  111-117.  [In  Russian.] 
magdefrau,  K.  1930.  Beitrage  zur  Kenntnis  der  Thuringischen  Buntsandsteinzeit.  Geol.  Thtir.  2,  285-289. 

— 1931.  Zur  Morphologie  und  Phylogenetische  Bedeutung  der  fossilen  Pflanzengattung  Pleuromeia.  Beih. 
Bot.  61.  48,  119-140. 


WANG  AND  WANG:  P LEU  RO  M El  A JIAOCHENGENSIS 


225 


neuberg,  M.  f.  1960a.  Permian  Flora  of  Petchora  Basin,  Part  I.  Trans.  Geol.  Inst.  Acad.  Nauk.  SSSR,  43,  1 64. 
[In  Russian.] 

— 19606.  Pleuromeia  Corda  from  the  Lower  Triassic  deposits  of  the  Russian  platform.  Ibid.  65-90. 
retallack,  G.  J.  1975.  The  life  and  times  of  a Triassic  Lycopod.  Alcheringa,  1,  3-29. 

sixtel,  t.  a.  1962.  Later  Permian  and  Early  Triassic  floras  in  South  Fergana.  Stratigr.  Palaeont.  Uzbekstana  and 
its  neighbouring  Region.  [In  Russian.] 

smith,  G.  M.  1955.  Cryptogamic  Botany,  Vol.  II:  Bryophytes  and  Pteridophytes.  Mcgraw-Hill  Co.  1959.  Chinese 
translation.  Pp.  1-287. 

WANG  LIXIN,  XIE  zhimin,  and  WANG  ziqiang,  1978.  On  the  occurrence  of  Pleuromeia  from  Qinshui  Basin  in 
Shanxi  Province.  Acta  Palaeont.  Sin.  17,  195-212.  [In  Chinese,  English  summary.] 
yin  hong-fu  and  lin  he-mao,  1979.  Triassic  marine  sediments  from  northern  Shanxi  with  note  on  the  age  of  the 
Shischienfeng  Group.  Acta  Stratigr.  Sin.  3,  233-241.  [In  Chinese.] 


WANG  ZIQIANG 

Tianjin  Institute  of  Geology  and  Mineral  Resources 
Ministry  of  Geology 

No.  26,  Jintang  Highway,  Tianjin  (Tientsin) 
People’s  Republic  of  China 


Typescript  received  13  June  1980 

Revised  typescript  received  19  December  1980 


WANG  LIXIN 

The  Regional  Geological  Survey  Team  of  Shanxi 
Yuci,  Shanxi  Province 
People’s  Republic  of  China 


THE  PALAEONTOLOGICAL  ASSOCIATION 


The  Association  was  founded  in  1957  to  further  the  study  of  palaeontology.  It  holds  meetings  and  demonstrations  as  well 
as  publishing  Palaeontology  and  Special  Papers  in  Palaeontology.  Membership  is  open  to  individuals  and  to  institutions  on 
payment  of  the  appropriate  annual  subscription.  Rates  for  1981  are: 

Institutional  membership.  . £32  (U.S.  S71) 

Ordinary  membership.  £15  (U.S.  $33) 

Student  membership.  . £9-50  (U.S.  $21) 

There  is  no  admission  fee.  Correspondence  concerned  with  Institutional  Membership  should  be  addressed  to  Dr.  C.  H.  C. 
Brunton,  Department  of  Palaeontology,  British  Museum  (Natural  History),  Cromwell  Road,  London  SW7  5BD,  England. 
Student  members  are  persons  receiving  full-time  instruction  at  educational  institutions  recognized  by  the  Council.  On  first 
applying  for  membership,  an  application  form  should  be  obtained  from  the  Membership  Treasurer.  Subscriptions  cover  one 
calendar  year  and  are  due  each  January;  they  should  be  sent  to  the  Membership  Treasurer  (for  name  and  address  see 
Council  list  below). 


PALAEONTOLOGY 

All  members  who  join  for  1982  will  receive  Volume  25,  Parts  1-4.  All  back  numbers  are  still  in  print  and  may  be  ordered  from 
Marston  Book  Services,  P.O.  Box  87,  Oxford  0X4  1LB,  England,  at  £15  (U.S.  $33)  per  part  (post  free). 

SPECIAL  PAPERS  IN  PALAEONTOLOGY 

Members  may  subscribe  to  this  by  writing  to  the  Membership  Treasurer  (for  name  and  address  see  Council  list  below): 
the  subscription  rate  for  1982  is  £25  (U.S.  $55)  for  Institutional  Members,  and  £12-50  (U.S.  $27-50)  for  Ordinary  and  Student 
Members.  A single  copy  of  each  Special  Paper  is  available  to  Ordinary  and  Student  Members  only,  for  their  personal  use, 
at  a discount  of  25%  below  the  prices  shown  in  the  list  inside  the  front  cover.  Non-members  may  obtain  copies,  at  the 
listed  prices,  from  Marston  Book  Services,  P.O.  Box  87,  Oxford  0X4  1LB,  England. 


COUNCIL  1981-1982 

President:  Dr.  W.  H.  C.  Ramsbottom,  Institute  of  Geological  Sciences,  Leeds  LS15  8TQ 
Vice-Presidents:  Professor  C.  B.  Cox,  Department  of  Zoology,  King’s  College,  Strand,  London  WC2R  2LS 
Professor  J.  W.  Murray,  Department  of  Geology,  The  University,  Exeter  EX4  4QE 
Treasurer:  Mr.  R.  P.  Tripp,  High  Wood,  Botsom  Lane,  West  Kings  Down,  Sevenoaks,  Kent  TNI  5 6BN 
Membership  Treasurer  : Dr.  J.  C.  W,  Cope,  Department  of  Geology,  University  College,  Swansea  SA2  8PP 
Secretary:  Dr.  R.  Riding,  Department  of  Geology,  University  College,  Cardiff  CF1  1XL 
Marketing  Manager:  Dr.  R.  J.  Aldridge,  Department  of  Geology,  The  University,  Nottingham,  NG7  2RD 


Editors 

Dr.  M.  G.  Bassett,  Department  of  Geology,  National  Museum  of  Wales,  Cardiff  CF1  3NP 
Dr.  K.  C.  Allen,  Department  of  Botany,  Bristol  University,  Bristol  BS8  1UG 
Dr.  R.  A.  Fortey,  Department  of  Palaeontology,  British  Museum  (Natural  History), 

Cromwell  Road,  London  SW7  5BD 

Dr.  A.  L.  Panchen,  Department  of  Zoology,  The  University,  Newcastle  upon  Tyne  NE1  7RU 
Dr  L.  R.  M.  Cocks,  Department  of  Palaeontology.  British  Museum  (Natural  History).  Cromwell  Road.  London  SW7  5BD 

Other  Members  of  Council 


Dr.  D.  E.  G.  Briggs,  London 
Dr.  C.  H.  C.  Brunton,  London 
Dr.  R.  Harland,  Leeds 
Dr.  A.  R.  Lord,  London 
Dr.  J.  Miller,  Edinburgh 
Dr.  T.  J.  Palmer,  Oxford 
Dr.  M.  Romano,  Sheffield 


Dr.  D.  J.  Siveter,  Leicester 

Dr.  P.  W.  Skelton,  Open  University 

Dr.  P.  D.  Taylor,  London 

Dr.  A.  T.  Thomas,  Birmingham 

Dr.  H.  S.  Torrens,  Keele 

Dr.  N.  H.  Trewin,  Aberdeen 

Dr.  J.  Watson,  Manchester 


Overseas  Representatives 

Australia  : Professor  B.  D.  Webby,  Department  of  Geology,  Sydney  University,  Sydney,  N.S.W.,  2006 
Canada:  Dr.  B.  S.  Norford,  Institute  of  Sedimentary  and  Petroleum  Geology,  3303-33rd  Street  NW„  Calgary,  Alberta 
New  Zealand:  Dr.  G.  R.  Stevens,  New  Zealand  Geological  Survey,  P.O.  Box  30368,  Lower  Hutt 
IVest  Indies  and  Central  America:  Mr.  John  B.  Saunders,  Geological  Laboratory,  Texaco  Trinidad,  Inc.,  Pointe-a-Pierre, 
Trinidad,  West  Indies 

U.S. A:  Dr  R.  Cuffey,  Department  of  Geology,  Pennsylvania  State  University, Pennsylvania,  U.S. A. 

South  America:  Dr.  O.  A.  Reig,  Departmento  de  Ecologia,  Universidad  Simon  Bolivar,  Caracas  108,  Venezuela 


Palaeontology 

VOLUME  25  ■ PART  1 


CONTENTS 

Rapid  evolution  in  echinoids 

P.  M.  KIER  1 

The  palaeobiology  of  the  Cretaceous  irregular  echinoids 
Infulaster  and  Hagenowia 

A.  S.  GALE  and  A.  B.  SMITH  1 1 

A revision  of  late  Ordovician  bivalves  from  Pomeroy, 

Co.  Tyrone,  Ireland. 

S.  P.  TUNNICLIFF  43 

Juvenile  specimens  of  the  ornithischian  dinosaur  Psittaco- 
saurus 

W.  P.  COOMBS,  JR.  89 

A description  of  the  generating  curve  of  bivalves  with  straight 

hinges 

M.  J.  ROGERS 

Reworked  acritarchs  from  the  type  section  of  the  Ordovician 
Caradoc  Series,  Shropshire 

R.  E.  TURNER  1 

Lower  Carboniferous  conodont  faunas  from  Ravenstonedale, 
Cumbria 

A.  C.  HIGGINS  and  W.  J.  VARKER  145 

Somasteroidea,  Asteroidea,  and  the  affinities  of  Luidia 
{Platasterias)  latiradiata 

D.  B.  BLAKE  167 

A Lower  Carboniferous  aistopod  amphibian  from  Scotland 

C.  F.  WELLSTEAD  193 

Morphology  and  relationships  of  the  Upper  Carboniferous 
aistopod  amphibian  Ophiderpeton  nanum 

M.  J.  BOYD  209 

A new  species  of  the  lycopsid  Pleuromeia  from  the  early  Tri- 
assic  of  Shanxi,  China,  and  its  ecology 

WANG  ZIQIANG  and  WANG  LIXIN  2 1 5 


Printed  in  Great  Britain  at  the  University  Press,  Oxford 
by  Eric  Buckley,  Printer  to  the  University 


issn  0031-0239 


Palaeontology 


VOLUME  25  • PART  2 APRIL  1982 


Published  by 


The  Palaeontological  Association  London 


Price  £1 7 


THE  PALAEONTOLOGICAL  ASSOCIATION 


The  Association  publishes  Palaeontology  and  Special  Papers  in  Palaeontology.  Details  of  membership  and  subscription  rate; 
may  be  found  inside  the  back  cover. 

The  journal  Palaeontology  is  devoted  to  the  publication  of  papers  on  all  aspects  of  palaeontology.  Review  articles  arc 
particularly  welcome,  and  short  papers  can  often  be  published  rapidly.  A high  standard  of  illustration  is  a feature  of  th<( 
journal.  Four  parts  are  published  each  year  and  are  sent  free  to  all  members  of  the  Association.  Typescripts  on  all.aspects  o 
palaeontology  and  stratigraphical  palaeontology  are  invited.  They  should  conform  in  style  to  those  already  published  in  thi;l 
journal,  and  should  be  sent  to  Dr.  R.  A.  Fortey,  Palaeontological  Association,  Department  of  Palaeontology,  British  Museun  i 
(Natural  History),  Cromwell  Road,  London  SW7  5BD,  England,  who  will  supply  detailed  instructions  for  authors  on  reques 
(these  were  published  in  Palaeontology  1977,  20,  pp.  921-929). 

Special  Papers  in  Palaeontology  is  a series  of  substantial  separate  works;  the  following  are  available  (post  free).- 

1.  (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  plates.  Price  £8  (U.S.  $17-50). 

2.  (for  1968):  Evolution  of  the  Shell  Structure  of  Articulate  Brachiopods,  by  A.  williams.  55  pp.,  27  text-figs.,  24  plateA 
Price  £5  (U.S.  $11). 

3.  (for  1968):  Upper  Maestrichtian  Radiolaria  of  California,  by  Helen  p.  foreman.  82  pp.,  8 plates.  Price  £3  (U.S.  $6-50 

4.  (for  1969):  Lower  Turonian  Ammonites  from  Israel,  by  R.  freund  and  m.  raab.  83  pp.,  15  text-figs.,  10  plates.  Price  £j 
(U.S.  $6-50). 

5.  (for  1969):  Chitinozoa  from  the  OrdoVician  Viola  and  Fernvale  Limestones  of  the  Arbuckle  Mountains,  Oklahoma 
by  w.  a.  M.  jenkins.  44  pp.,  10  text-figs.,  9 plates.  Price  £2  (U.S.  $4-50). 

6.  (for  1969):  Ammonoidea  from  the  Mata  Series  (Santonian-Maastrichtian)  of  New  Zealand,  by  r.  a.  Henderson.  82  pp 
13  text-figs.,  15  plates.  Price  £3  (U.S.  $6-50). 

7.  (for  1970):  Shell  Structure  of  the  Craniacea  and  other  Calcareous  Inarticulate  Brachiopoda,  by  A.  williams  an\ 
a.  d.  Wright.  51  pp.,  17  text-figs.,  15  plates.  Price  £1-50  (U.S.  $3-50). 

8.  (for  1970):  Cenomanian  Ammonites  from  Southern  England,  by  w.  J.  Kennedy.  272  pp.,  5 tables,  64  plates.  Price  £ 
(U.S.  $17-50). 

9.  (for  1971):  Fish  from  the  Freshwater  Lower  Cretaceous  of  Victoria,  Australia,  with  Comments  on  the  Palaeoj 
environment,  by  M.  waldman.  130  pp.,  37  text-figs.,  18  plates.  Price  £5  (U.S.  $11). 

10.  (for  1971):  Upper  Cretaceous  Ostracoda  from  the  Carnarvon  Basin,  Western  Australia,  by  r.  h.  bate.  148  pp.,  43  tex  \ 
figs.,  27  plates.  Price  £5  (U.S.  $11). 

11.  (for  1972):  Stromatolites  and  the  Biostratigraphy  of  the  Australian  Precambrian  and  Cambrian,  by  M.  R.  waltei 
268  pp.,  55  text-figs.,  34  plates.  Price  £10  (U.S.  $22). 

12.  (for  1973):  Organisms  and  Continents  through  Time.  A Symposium  of  23  papers  edited  by  N.  F.  hughes.  340  pp 
132  text-figs.  Price  £10  (U.S.  $22)  (published  with  the  Systematics  Association). 

13.  (for  1974):  Graptolite  studies  in  honour  of  O.  M.  B.  Bulman.  Edited  by  R.  B.  rickards,  d.  e.  jackson,  and c.  p.  hughe] 
261  pp.,  26  plates.  Price  £10  (U.S.  $22). 

14.  (for  1974):  Palaeogene  Foraminiferida  and  Palaeoecology,  Hampshire  and  Paris  Basins  and  the  English  Channel,  h 
j.  w.  Murray  and c.  a.  wright.  171  pp.,  45  text-figs.,  20  plates.  Price  £8  (U.S.  $17-50). 

15.  (for  1975):  Lower  and  Middle  Devonian  Conodonts  from  the  Broken  River  Embayment,  North  Queensland,  Australia 
by  p.  G.  telford.  100  pp.,  9 text-figs.,  16  plates.  Price  £5-50  (U.S.  $12). 

16.  (for  1975):  The  Ostracod  Fauna  from  the  Santonian  Chalk  (Upper  Cretaceous)  of  Gingin,  Western  Australia,  /| 
j.  w.  neale.  131  pp.,  40  text-figs.,  22  plates.  Price  £6-50  (U.S.  $14-50). 

17.  (for  1976):  Aspects  of  Ammonite  Biology,  Biogeography,  and  Biostratigraphy,  by  w.  J.  Kennedy  and  w.  a.  cobba 
94  pp.,  24  text-figs.,  11  plates.  Price  £6  (U.S.  $13). 

18.  (for  1976):  Ostracoderm  Faunas  of  the  Delorme  and  Associated  Siluro-Devonian  Formations,  North  West  Terri  torie 
Canada,  by  d.  l.  dineley  and  e.  j.  loeffler.  218  pp.,  78  text-figs.,  33  plates.  Price  £20  (U.S.  $44). 

19.  (for  1977):  The  Palynology  of  Early  Tertiary  Sediments,  Ninetyeast  Ridge,  Indian  Ocean,  byE.  m.  kemp  and  w.  k.  harr 
74  pp.,  2 text-figs.,  8 plates.  Price  £7  (U.S.  $15-50). 

20.  (for  1977):  Fossil  Priapulid  Worms,  by  s.  c.  morris.  159  pp.,  99  text-figs.,  30  plates.  Price  £16  (U.S.  $35). 

21.  (for  1978):  Devonian  Ammonoids  from  the  Appalachians  and  their  bearing  on  International  Zonation  and  Correlt 
tion,  by  M.  r.  house.  70  pp.,  12  text-figs.,  10  plates.  Price  £12  (U.S.  $26-50). 

22.  (for  1978,  published  1979):  Curation  of  palaeontological  collections.  A joint  colloquium  of  The  Palaeontological  Associ 
tion  and  Geological  Curators’  Group.  Edited  by  m.  g.  bassett.  280  pp.,  53  text-figs.  Price  £25  (U.S.  $55). 

23.  (for  1979):  The  Devonian  System.  A Palaeontological  Association  International  Symposium.  Edited  by  M.  R.  house,  cJ 
scrutton,  and  M.  G.  bassett.  353  pp.,  102  text-figs.,  1 plate.  Price  £30  (U.S.  $66). 

24.  (for  1980):  Dinofiagellate  Cysts  and  Acritarchs  from  the  Eocene  of  southern  England,  by  J.  b.  bujak,  c.  downie,  G. 
eaton,  and  G.  l.  williams.  104  pp.,  24  text-figs.,  22  plates.  Price  £15  (U.S.  $33). 

25.  (for  1980):  Stereom  Microstructure  of  the  Echinoid  Test,  by  a.  b.  smith.  81  pp.,  20  text-figs .,  24  plates.  Price  £15  ($31; 

26.  (for  1981):  The  fine  structure  of  Graptolite  periderm,  by  p.  r.  crowther.  1 19  pp.,  37  text-figs.,  20  plates.  Price  £25  ($51 

27.  (for  1981):  Late  Devonian  Acritarchs  from  the  Carnarvon  Basin, Western  Australia,  by  G.  playford  andR.  s.  dring.  78  p 
10  text-figs.,  19  plates.  Price  £15  ($33). 

© The  Palaeontological  Association , 1982 


Cover:  An  acritarch  of  the  genus  Multiplicisphaeridium  from  the  Tournaisian  Bedford  Shale  of  Ohio,  U.S. A.  Institute; 
Geological  Sciences  specimen  number  MPK  3569.  S.E.M.  x 1000.  Within  the  last  twenty-five  years  the  number  of  descrit 
acritarch  taxa  has  increased  dramatically,  and  these  organisms  are  now  used  extensively  to  date  rocks  of  late  Prccambr 

to  Devonian  age. 


ECOLOGY  AND  POPULATION  STRUCTURE  OF 
THE  RECENT  BRACHIOPOD  TEREBRATU LIN  A 
FROM  SCOTLAND 

by  GORDON  B.  CURRY 


Abstract.  The  ecology  and  population  structure  of  the  Recent  articulate  brachiopod  Terebratulina  retusa 
(Linnaeus)  are  described.  The  population  studied  occurs  around  the  margins  of  a depression  of  more  than  220  m 
in  the  Firth  of  Lome,  Scotland,  and  is  predominantly  attached  to  the  horse-mussel  Modiolus  modiolus 
(Linnaeus).  Spawning  occurs  regularly  in  late  spring  and  late  autumn,  and  is  initiated  at  temperatures  of 
10-1 1°C.  The  highly  synchronized  reproductive  cycle,  from  spawning  to  spatfall,  occurs  within  3 weeks  in 
nature.  Length-frequency  histograms  prepared  from  large  representative  samples  collected  at  regular  inter- 
vals during  1977-1979  are  unimodal  and  right-skewed  due  to  the  predominance  of  juveniles.  Regularly  spaced 
subsidiary  peaks  in  the  histograms  correspond  to  biannual  settlement  cohorts;  in  later  life  successive  peaks 
merge  to  form  a single  annual  peak.  This  pattern  is  identical  to  that  predicted  by  computer-based  simulations. 
Recently  settled  specimens  grow  rapidly  to  an  average  length  of  2-75  mm  within  3 months  during  both  spring  and 
autumn;  thereafter  the  animals  grow  (initially  by  4 mm  per  year)  throughout  life,  although  at  a progressively 
reducing  rate  from  the  third  year  of  life  onwards.  Growth  slows  or  ceases  in  winter  in  all  but  recently  settled 
specimens.  The  maximum  life  span  is  7 years.  The  mortality  rate  remains  constant,  although  the  causes  of  death 
are  not  apparent.  The  growth-lines  form  biannually,  at  times  of  pronounced  environmental  and  physiological 
disturbance. 


It  is  not  generally  realized  that  the  Recent  brachiopod  fauna  of  the  British  Isles  (21  species  of  17 
genera,  Brunton  and  Curry  1979)  is  significantly  more  diverse  than  that  of  New  Zealand  (12  species 
of  9 genera).  Considering  the  relative  diversity  of  the  two  faunas,  it  may  appear  strange  that  New 
Zealand  has  become  the  classic  area  for  Recent  brachiopod  research.  However,  this  apparent 
anomaly  is  readily  explained  since  the  New  Zealand  species  are  far  more  accessible,  with  25%  of  the 
species  abundant  intertidally  and  a further  30%  common  in  nearshore  shallow  subtidal  habitats.  By 
contrast,  British  brachiopods  are  very  rarely  found  intertidally  and  the  majority  of  species  have  only 
been  collected  in  small  numbers  from  widely  dispersed  and  often  inaccessible  localities. 

Intermittent  research  during  the  last  100  years  has  demonstrated  that  several  of  the  British  species 
are  locally  abundant,  especially  off  the  west  coast  of  Scotland  and  in  the  Western  Approaches  (Atkins 
1959a,  b,  1960 a,  b,  c,  1961;  Davidson  1886-1888;  Chumley  1918;  see  summary  maps  in  Brunton  and 
Curry  1979),  although  dredging  is  the  only  practicable  method  of  sampling  these  predominantly 
deeper-water  populations.  The  recent  upsurge  of  interest  in  palaeoecology  has  emphasized  the  need 
for  precise  data  on  the  life-habits  and  ecology  of  Recent  representatives  of  fossil  phyla,  and  more  data 
are  needed  to  augment  the  patchy  and  often  contradictory  information  available  to  the  brachiopod 
palaeoecologist.  The  present  study  was  initiated  in  the  light  of  such  inadequacies,  and  when  it  became 
clear  that  the  comprehensive  dredging  facilities  available  at  Dunstaffnage  Marine  Research 
Laboratory,  nr.  Oban,  Scotland,  could  ensure  access  to  the  abundant  brachiopod  populations  off  the 
west  coast  of  Scotland. 

Although  dealing  entirely  with  living  animals,  this  study  was  conducted  from  a palaeontological 
viewpoint.  As  such,  the  main  interest  was  the  interpretation  of  features  on  the  shell,  in  particular  the 
analysis  of  population  structure  and  dynamics  based  on  size-frequency  histograms  prepared  from 
large  samples  of  brachiopods  collected  at  regular  intervals  throughout  1977-1979.  The  main 
advantage  of  working  with  living,  as  opposed  to  fossil,  populations  is  that  it  is  possible  to  seek 


IPalaeontology,  Vol.  25,  Part  2, 1982,  pp.  227-246.| 


228 


PALAEONTOLOGY,  VOLUME  25 


confirmatory  evidence  for  conclusions  reached  by  direct  observations  on  the  living  population.  This 
paper  first  describes  aspects  of  the  biology  and  ecology  of  the  animal,  and  goes  on  to  apply  the  results 
to  the  interpretation  of  size-frequency  histograms. 

MATERIAL 

The  species  under  investigation  was  Terebratulina  retusa  (Linnaeus),  which  is  the  most  abundant  of  the  Recent 
British  brachiopods.  Initially  the  intention  was  to  collect  regular  bulk  samples  from  the  large  populations  of  T. 
retusa  which  were  known  to  occur  in  the  Sound  of  Mull  (text-fig.  1).  However,  during  an  exploratory  cruise  in 
March  1977  an  extremely  abundant  population,  ideally  situated  for  regular  sampling,  was  discovered  near  a 
depression  in  the  nearby  Firth  of  Lome  (text-fig.  1).  This  depression,  which  reaches  a maximum  depth  in  excess 
of  220  m,  is  mid-way  between  the  Island  of  Kerrera  and  the  south-west  coast  of  the  Island  of  Mull  (Grid.  ref.  NM 
745  265;  Stn.  1 in  text-fig.  1),  and  is  thought  to  lie  along  the  Firth  of  Lome  Fault  (a  subsidiary  splay  of  the  Great 
Glen  Fault,  Barber  et  al.  1979).  At  this  locality  the  brachiopods  are  predominantly  attached  to  the  horse-mussel 
Modiolus  modiolus  (Linnaeus)  which  occurs  in  dense  beds  around  the  smoothly  sloping  margins  of  the 
depression. 

The  samples  were  collected,  by  the  R/V  Calanus  or  Seol  Mara,  using  a conventional  ‘clam-dredge’  (1-2  m 
wide  x 2 m long),  the  body  of  which  consists  of  an  outer  framework  of  interlocking  iron  chain  and  an  inner 
nylon  meshwork  with  a maximum  aperture  of  15  mm.  This  dredge  is  designed  to  collect  material  resting  on,  or 
partially  buried  within,  the  substrate,  and  proved  to  be  extremely  efficient  at  sampling  the  Firth  of  Lome  mussel 
beds.  For  each  sample  the  dredge  was  trawled  slowly  along  the  substrate  for  approximately  5-10  minutes,  in  a 


text-figure  1.  Map  showing  locations  of  R/V  Calanus 
sample  stations  off  the  west  coast  of  Scotland;  station 
numbers  refer  to  Table  1 (see  opposite). 


CURRY:  B R ACHIOPOD  TEREB  RATE  LIN  A FROM  SCOTLAND 


229 


north-easterly  or  south-westerly  direction.  Once  landed  on  deck,  the  sample  was  washed  with  sea-water  to 
remove  any  adherent  sediment,  and  then  placed  in  plastic  baths  supplied  with  flowing  fresh  sea-water. 

The  disturbance  caused  by  the  dredging  was  minimal,  and  the  majority  of  the  sample  had  become 
‘acclimatized’  to  the  plastic  baths  and  commenced  feeding  within  half  an  hour  of  arriving  on  deck.  Mechanical 
damage  during  dredging  was  not  significant,  and  only  seven  specimens  out  of  the  818  brachiopods  collected  in 
March  1977  (i.e.  0-85%)  proved  unmeasurable  because  of  shell  damage.  Due  to  the  efficiency  of  the  ‘clam- 
dredge’,  and  the  density  and  abundance  of  both  mussels  and  brachiopods,  all  attempts  to  sample  the  Firth  of 
Lome  populations  were  successful.  In  addition,  the  Firth  of  Lome  depression  is  a prominent  submarine  feature, 
easily  recognizable  using  the  ship’s  depth-sounding  equipment,  and  it  was  therefore  possible  to  ensure  that  all 
samples  were  collected  from  the  same  population. 


table  1.  Results  of  the  dredging  operations  of  the  R/V  Calanus,  west  coast  of  Scotland,  1977-1979 
(a  = abundant,  c = common,  r = rare,  x = absent). 


STATION  No 

LOCATION. 

DEPTH. 

SEDIMENT . 

abund 

T.retusa 
. j a t ta  : tuna  n f. . 

abund 

anomala 
.]  attachment . 

ASSOCIATED  FAUNA. 

1 

Firth  of  Lome 

146-I83m. 

fine  mud 

a 

1 

1 Modiolus 

a 

1 

1 Modiolus 

see  text. 

2 

Rabbit  Island 

13m. 

-- 

c 

1 rocks 

r 

, ri  :ks 

Pecten 

Garbh  Reisa 

20m. 

— 

X 

! -- 

A 

1 -- 

Pecten 

4 

Sound  of  Jura 
(north-east ) 

?3-128m. 

-- 

X 

1 

X 

1 

1 

Modi  plus  and 
hydrozoans 

5 

Sound  of  Jura 

110m. 

- 

r 

1 

1 vesicular 
, basalt 

X 

large  no.  of  dis- 
articulated Modi  plus 

6 

Sound  of  Jura 
(south) 

37m. 

-- 

1 

x 

_J 

Pecten.  Balanus. 

? 

E.  of  Colonsay 

37m. 

sand 

X 

' -- 

X 

seaweed,  crabs. 

8 

North-east  of 
Colonsay 

55m. 

— 

0 

1 rocks  and 
1 clinker 

c 

1 rocks  and 
1 clinker 

-- 

9 

Torran  Rooks 

91-llPm. 

— 

X 

1 

| 

A 

1 .. 

— 

10 

North-west  of 
Iona 

?3-110m 

-- 

r 

1 rocks 

1 

X 

— 1 

1 

— 

11 

North  of  Iona 

73-123m. 

sand 

1 

1 

X 

1 — 

— 

12 

Treshnish  Is. 

29-55m. 

- 

x 

1 

X 

“ 

1 

Balanus . serpulids 
on  rocks 

13 

Lunga 

37m. 

— 

X 

! - 

X 

1 -- 

Balanus  on  rocks 

14 

North-west  of 
Mull 

37-46m. 

X 

1 

Pecten.  eohinoids 

15 

East  of  Coll 

146-183m. 

— 

X 

1 .. 

0 

1 rooks 

-- 

16 

Ardmore  Point, 
Mull 

73-213m. 

mud 

X 

1 

X 

1 

1? 

Mingary  Bay, 
Ardnamurchan 

37m. 

mud 

X 

1 

= 

rocks 

1 

13 

Tobermory  Bay, 
Mull 

73m. 

mud 

c 

1 clinker 

1 

a 

clinker  and 
1 rocks 

Modiolus 

19 

Sound  of  Mull 
(north) 

110m. 

mud 

c 

' rocks 

a 

* rocks 

l 

20 

Sound  of  Mull 
(north) 

Qlm. 

mud 

a 

1 modiolus 

1 

c 

1 Modiolus 

1 

21 

Sound  of  Mull 
(north ) 

91m. 

mud 

x 

1 

1 

1 

c 

1 rocks 

-- 

22 

Sound  of  Mull 
(north ) 

91m. 

mud 

a 

1 Modiolus 

1 

a 

. Modiolus 

_J 

23 

Sound  of  Mull 
(north ) 

91-l46m. 

mud 

r 

1 

r 

1 Modiolus 

| 

?4 

Sound  of  Mull 
(north) 

18-  37m. 

- 

0 

1 shell 
1 fragments 

X 

| 

25 

Sound  of  Mull 
( south ) 

110- 128m. 

mud 

a 

Modiolus 

l 

a 

Modiolus 

26 

Sound  of  Mull 
( south ) 

123-146m. 

mud 

c 

1 clinker 

1 

r 

1 clinker 

1 

2? 

Sound  of  Mull 
(south) 

l8-37m. 

mud 

c 

1 clinker 

1 

r 

1 clinker 

_l 

2 3 

Li smore  Island 

37m. 

- 

r 

1 vesicular 
1 basalt 

r 

vesicular 
1 basalt 

29 

Lismore  Island 

15-2 6m. 

- 

r 

1 rocks 

1 - 

230 


PALAEONTOLOGY,  VOLUME  25 


The  samples  were  transported  back  to  the  laboratory  and  placed  in  an  outside  aquarium  through  which  fresh 
sea-water  was  continually  being  circulated.  Because  of  the  difficulty  in  seeing  very  small  brachiopods  (recently 
settled  post-larvae  are  less  than  0-5  mm  in  length  and  their  shells  are  transparent),  the  surfaces  of  all  mussels  and 
other  potential  brachiopod  substrates  were  examined  using  a binocular  microscope.  Once  detected,  each 
brachiopod  was  removed  from  its  substrate  by  severing  its  pedicle  with  a sharp  scalpel,  measured  to  the  nearest 
01  mm  using  ‘MITUTOYO’  dial  calipers,  and  then  preserved  in  either  10%  formalin  or  70%  alcohol.  The 
samples  have  been  deposited  in  the  Department  of  Palaeontology,  British  Museum  (Natural  History),  London, 
and  the  registration  numbers  quoted  in  the  text  (with  the  prefix  ZB)  refer  to  the  Recent  brachiopod  collections  in 
that  museum. 


ECOLOGY  OF  TEREBRATULINA  RETUSA 

Distribution.  The  precise  geographic  limits  of  the  distribution  of  T.  retusa  are  unknown,  as  there  are 
two  morphologically  similar  and  often  confused  species  of  Terebratulina  in  the  North  Atlantic. 
Positively  identified  T.  retusa  have  been  collected  from  as  far  north  as  Norway  (the  type  area)  and  as 
far  south  as  Spain  and  the  Mediterranean.  The  species  would  appear  to  be  confined  to  the  north- 
eastern North  Atlantic,  although  it  has  been  recorded  from  the  east  coast  of  Greenland  (Wesenberg- 
Lund  1940).  The  possibility  that  the  two  named  species  are  members  of  a Terebratulina  cline  has  been 
considered  by  several  authors  (e.g.  Wesenberg-Lund  1941),  and  certainly  the  presence  of  individuals 
with  intermediate  morphological  characteristics  in  the  mid  North  Atlantic  could  explain  the 
confusion  over  the  geographic  distribution  of  the  two  species,  which  are  quite  distinct. 

T.  retusa  is  relatively  common  off  the  west  coast  of  Scotland,  and  has  been  recorded  from  many  of 
the  sea-lochs  (Davidson  1886-1888;  Chumley  1918).  On  a more  local  scale  the  comprehensive 
dredging  equipment  available  on  the  R/V  Calanus  provided  an  opportunity  to  investigate  the  pattern 
of  distribution  of  T.  retusa  around  the  Islands  of  Mull  and  Jura  (text-fig.  1).  The  results  of  these 
dredging  operations  provide  strong  evidence  of  an  essentially  patchy  distribution,  which  is  illustrated 
by  the  results  of  a comprehensive  sampling  programme  in  the  Sound  of  Mull  (Stns.  18-27  in  text-fig.  1 
and  Table  1).  In  most  of  these  samples  the  horse-mussel  M.  modiolus  formed  the  bulk  of  the  sample, 
whilst  T.  retusa  varied  from  being  the  numerically  dominant  constituent  (e.g.  Stn.  22),  to  being  a 
minor  constituent  (e.g.  Stn.  23),  or  being  totally  absent  (e.g.  Stn.  21).  Suitable  substrate  for 
brachiopod  attachment  was  present  at  the  latter  station,  and  the  fact  that  the  habitat  occupied  by  T. 
retusa  is  not  occupied  by  other  sessile  epibenthos  suggests  that  competition  for  space  is  not  an 
important  factor  in  these  deeper  water  localities.  It  is  likely,  therefore,  that  the  patchy  distribution  is 
an  inherent  feature  of  the  T.  retusa  population,  reflecting  the  short  duration,  and  hence  limited 
dispersal  range,  of  the  pelagic  larval  stage  (see  below).  Many  other  Recent  brachiopod  populations 
appear  to  be  patchily  distributed  (Rudwick  1970,  p.  156). 

It  is  appropriate  to  sound  a cautionary  note  on  the  use  of  benthic  dredging  to  determine  the 
distribution  pattern  of  sessile  epibenthos  such  as  T.  retusa.  Benthic  dredges  will  only  collect  free-lying 
or  loosely  attached  material  which  falls  within  a restricted  size  range,  and  it  is  clear  that  the  substrates 
used  by  brachiopods  are  very  variable  both  in  grain-size  and  in  the  extent  to  which  they  are  anchored 
to  the  sea  bed.  Therefore  the  absence  of  brachiopods  in  a particular  sample  may  simply  be  due  to  the 
inability  of  the  dredge  to  sample  substrates  such  as  large  boulders  or  rock-faces.  This  feature  was 
evident  during  the  present  study,  as  dredged  samples  from  the  Sound  of  Jura  yielded  no  brachiopods 
(Stn.  4),  and  yet  divers  have  reported  quantities  of  T.  retusa  attached  to  large  boulders  in  this  region 
(R.  Harvey,  pers.  comm.  1978).  Benthic  dredging  is  only  a satisfactory  method  of  sampling 
populations  in  which  the  predominant  substrate  is  of  a size  and  nature  that  can  readily  be  collected,  as 
is  the  case  with  the  mussel  beds  in  the  Firth  of  Lome. 

Depth.  Within  the  study  area  the  most  abundant  populations  of  T.  retusa  occur  at  depths  of  100-200 
m,  although  a few  specimens  have  been  collected  from  an  estimated  depth  of  3 m in  the  Sound  of 
Raasay  (M.  K.  Howarth,  pers.  comm.  1978),  and  a reasonably  abundant  population  is  known  to 
occur  at  a depth  of  13-20  m off  the  coast  of  Rabbit  Island  (Stn.  2 in  text-fig.  1 and  Table  1).  Despite 
these  occurrences,  T.  retusa  is  not  a common  constituent  of  the  shallow  subtidal  ecosystem  along  the 
Scottish  west  coast,  and  there  is  no  record  of  any  dead  shells  being  washed  up  on  beaches  in  this 


CURRY:  B RACHIOPOD  TEREB  RATE  LIN  A FROM  SCOTLAND 


23 


region.  The  known  depth  range  of  T.  retusa  is  from  3-1478  m,  although  it  is  most  commonly  found 
between  100  and  500  m.  Environmental  conditions  vary  considerably  with  depth,  and  it  is  the 
tolerance  of  an  organism  to  such  variations  which  determines  its  depth  range;  depth  in  itself  is  not 
considered  to  be  a controlling  factor  (Moore  1958).  Presumably,  therefore,  T.  retusa  is  prevented 
from  colonizing  intertidal  or  shallow  tidal  habitats  by  the  combined  effects  of  biological  and  physio- 
chemical  factors  characteristic  of  such  habitats,  such  as  competition  with  other  organisms,  rapid 
daily  temperature  fluctuations,  the  possibility  of  increased  predation  intensity,  periodic  desiccation, 
salinity  fluctuations,  wave  turbulence,  and  intensity  of  incident  light.  Competition  for  space  is 
certainly  an  inhibiting  factor,  as  was  illustrated  during  the  present  study  when  brachiopods  collected 
from  the  Sound  of  Mull  were  reintroduced  into  the  shallow  marine  environment  of  Dunstaflfnage  Bay 
in  recoverable  cages.  These  brachiopods  were  killed  by  a dense  settlement  of  fast-growing  barnacles 
which  engulfed  and  smothered  them  within  3 months.  Prolific  swarms  of  barnacles  are  common  in 
shallow-water  habitats  along  the  Scottish  coast,  but  in  the  deep-water  localities  where  T.  retusa  is 
abundant  they  are  rare  and  solitary  (in  fact  a different  species).  Clearly  the  relatively  slow-growing 
T.  retusa  is  at  a considerable  disadvantage  in  areas  where  suitable  substrates  are  in  short  supply. 

Temperature.  Temperature  is  one  of  the  most  important  environmental  parameters  for  cold-blooded 
marine  invertebrates  (Moore  1958),  limiting  their  geographic  range,  and  affecting  all  aspects  of  life  by 
virtue  of  its  controlling  effect  on  the  rates  of  metabolic  processes  (i.e.  Van’t  Hoff’s  Law— see  Prosser 
1973).  The  general  pattern  of  temperature  fluctuation  experienced  by  T.  retusa  in  the  Firth  of  Lome 
depression  was  determined  (sea-water  samples  were  collected  from  the  vicinity  of  the  brachiopod 
population  using  insulated  sampling  bottles),  and  has  been  compared  with  surface  water 
temperatures  in  text-fig.  2.  The  temperature  of  bottom  waters  in  the  Firth  of  Lome  range  from  a 
February  minimum  of  6-5  °C  to  a maximum  of  13°C  in  August,  an  annual  range  of  6-5  °C  which  is 
smaller  than  would  be  expected  at  this  latitude  because  of  the  warming  effect  of  the  North  Atlantic 
Drift  Current.  The  over-all  pattern  and  range  is  similar  in  surface  waters,  (minimum  in  January  of 
6°C  and  a maximum  of  14°C  in  August— annual  range  of  8°C),  although  it  is  apparent  from  text-fig. 
2 that  the  bottom  and  surface  waters  are  virtually  never  isothermal  and  can  differ  by  as  much  as 
2-5  °C.  This  vertical  stratification  of  the  water  column,  the  thermocline,  is  a common  feature  of 
Scottish  sea-lochs  and,  although  complex,  can  be  considered  as  reflecting  the  greater  susceptibility  of 


text-figure  2.  Annual  temperature  curves  for  the  surface  and 
bottom  (approx.  150  m)  waters  in  the  Firth  of  Lome,  Scotland. 


232 


PALAEONTOLOGY,  VOLUME  25 


surface  waters  to  prevailing  air  temperatures  (e.g.  being  warmer  in  summer  and  cooler  in  winter). 
Other  factors  which  enhance  this  vertical  stratification  are  density  and  salinity  gradients  produced  by 
the  high  levels  of  surface  run-off  of  fresh  water  from  the  surrounding  land.  The  thermocline  breaks 
down  during  two  short  periods  in  spring  and  autumn  when  the  water  column  becomes  thoroughly 
mixed,  isothermal,  and  uniformly  saturated  with  oxygen  (i.e.  the  points  at  which  the  curves  in  text-fig. 
2 cross). 

Substrate.  The  predominant  utilization  of  M.  modiolus  as  substrate  by  the  brachiopods  in  the  Firth  of 
Lome  depression  is  well  illustrated  by  the  analysis  of  the  substrate  of  attachment  of  the  786  specimens 
of  T.  re  tusa  collected  on  5 May  1977  (Table  2).  The  sample  contained  204  living  mussels,  166  of  which 


No.  of 

No.  of 

°/.  °f 

table  2.  Substrate  of  attachment  of  the  786 

Substrate 

Occurrences 

Brachiopods 

Total 

specimens  of  T.  retusa  collected  on  5 May  1977 
from  the  Firth  of  Lome. 

Living  mussel 

166 

591 

75 

Dead  mussel 

27 

53 

7 

Mussel  frags. 

33 

63 

9 

Hydrozoan 

10 

13 

2 

Dead  gastropod 

5 

14 

2 

Dead  brachiopod 

4 

10 

1 

Living  & dead 

4 

8 

! 

Sponge/ascidian 

4 

16 

2 

Unattached 

4 

8 

1 

Rock 

2 

9 

1 

Tube-worm 

1 

1 

0.1 

(i.e.  81%)  had  been  utilized  as  substrate  by  T.  retusa.  A total  of  707  brachiopods  (91%  of  the  sample) 
were  attached  to  living  or  dead  mussels  or  to  fragments  of  mussel  shell.  Mussels  are  ideal  substrate  for 
brachiopods,  being  sessile  and  of  much  longer  life  span  (at  least  20  years  (Comley  1978)  as  compared 
with  a maximum  of  7 years  for  T.  retusa).  The  relatively  large  surface  area  of  the  mussels  is  readily 
bored  by  the  anchoring  pedicle  rootlets  of  T.  retusa,  and  is  free  of  obstructions  which  would  hinder 
the  rotation  of  brachiopods  around  their  pedicles,  a procedure  which  enables  them  to  move  away 
from  localized  disturbances  and  to  take  up  preferred  feeding  orientations.  In  addition,  the  majority  of 
brachiopods  are  attached  anteriorly,  close  to  the  inhalant  and  exhalant  feeding  currents  of  the 
mussels,  and  are  likely  to  benefit  from  the  enhanced  flow  of  nutrient-rich  sea-water  in  such  regions. 
The  selective  sampling  bias  of  the  benthic  dredge  makes  it  impossible  to  determine  to  what  extent  the 
data  in  Table  2 can  be  considered  representative.  However,  the  thin  covering  of  fine-grained 
glauconitic  mud  in  this  area  is  likely  to  preclude  dense  settlements  directly  on  to  rock  surfaces, 
because  of  the  high  risk  of  being  smothered  or  choked  by  suspensions  of  sediment  stirred  up  by 
bottom  currents.  Under  such  circumstances  the  elevated  position  resulting  from  attachment  to 
mussels  would  be  all  the  more  advantageous  for  brachiopods,  especially  as  the  mussels’  ability 
for  limited  reorientation  will  ensure  that  their  anterior  region  remains  above  the  substrate  in  the  event 
of  sediment  accumulations.  Nevertheless,  direct  attachment  to  substrates  such  as  boulders  or  rock- 
faces  is  to  be  expected  in  the  Firth  of  Lome  depression  when  suitable  surfaces  are  free  of  sediment  due 
to  inclination  or  current  scour. 

Current.  Little  can  be  said  about  the  currents  in  the  vicinity  of  the  Firth  of  Lome  population,  as  direct 
measurements  of  velocity  was  beyond  the  scope  of  this  study.  However,  indirect  evidence  of  strong 
bottom  currents  comes  from  the  fact  that  the  great  majority  of  dead  shells  are  moved  away  from 
the  living  population  and  have  presumably  accumulated  in  the  deepest  regions  of  the  depression.  This 
phenomenon  is  being  investigated  with  the  aim  of  assessing  the  fossilization  potential  of  the  Firth  of 
Lome  brachiopod  population.  A strong  and  constantly  flowing  current  is  a necessity  for  sessile 
benthic  invertebrates  such  as  T.  retusa,  which  feeds  on  material  carried  in  suspension  in  the 
surrounding  sea-water. 


CURRY:  BRACHIOPOD  TEREBRATULINA  FROM  SCOTLAND 


233 


Associated  fauna.  Apart  from  T.  retusa  the  surfaces  of  the  Firth  of  Lome  mussels  are  used  as  substrate 
by  a wide  variety  of  sessile  organisms  and  plants,  the  most  numerous  of  which  being  the  cemented 
inarticulate  brachiopod  Crania  anomala  (Muller),  hydroids,  sponges,  chitons,  spirorbid  worms, 
foraminifera,  the  bivalve  Anomia,  and  gastropods  (e.g.  Cappula).  The  dredged  samples  also  contained 
abundant  ophiuroids  ( Ophiothrix  and  Ophiura),  and  asteroids  ( Crossaster , Solaster,  and  Asterias ) are 
often  present  but  in  much  smaller  numbers.  Occasionally  echinoids  ( Echinus ) and  decapods  ( Munida 
and  Galathea)  have  been  recovered.  As  compared  with  shallow-water  localities  in  the  Firth  of  Lome, 
this  fauna  is  impoverished,  especially  in  sessile  epibenthic  organisms  which  would  be  in  direct 
competition  with  T.  retusa  for  space  and  nutrients.  Burrowing  animals,  such  as  bivalves  and 
polychaetes,  are  rare,  presumably  because  of  the  thin  sediment  cover.  The  external  shell  surfaces  of  T. 
retusa  are  themselves  used  as  substrate  by  a wide  variety  of  animals  and  plants,  in  particular  sponges, 
hydroids,  bryozoans,  and  spirorbid  worms.  Shallow  borings,  almost  entirely  confined  to  the  primary 
shell  layer  of  T.  retusa,  are  thought  to  be  the  work  of  phoronids  and  fungi  (B.  Akpan,  pers.  comm. 
1980).  These  borings  are  not  predatory,  and  none  of  the  major  predators  in  this  fauna  appears  to  feed 
on  T.  retusa,  although  it  is  impossible  to  assess  the  effect  of  all  potential  predators  (e.g.  fish).  Some 
small  worms  have  been  found  in  the  brachial  cavity  of  a few  specimens  of  T.  retusa,  but  it  is  not  clear  if 
these  are  parasitic  or  have  merely  been  drawn  in  by  the  feeding  currents. 


REPRODUCTION 

It  was  of  great  importance  to  accurately  determine  the  timing  and  frequency  of  spawning  in  T.  retusa, 
as  such  information  is  necessary  to  establish  the  population  structure  and  dynamics.  In  T.  retusa  the 
sexes  are  separate  and  can  be  distinguished  on  the  basis  of  colour  (males  are  whitish,  whilst  the  tissues 
of  females  are  orange-red),  and  by  the  examination  of  the  gonads.  Males  and  females  are 
approximately  equally  represented  (342  adult  specimens  could  be  sexed  in  the  sample  collected  on  26 
October  1977,  181  of  which  were  male,  and  161  female).  One  pair  of  gonads  is  present  in  each  valve, 
developing  within  the  mantle  canals  between  the  inner  and  outer  mantle  epithelium  and  extending 
posteriorly  into  the  body  cavity.  Individual  gametes  develop  along  narrow  interconnecting  genital 
canals,  which  are  anchored  to  the  outer  epithelium  by  linear  membranes  (Hancock  1859).  Columns  of 
tissue,  situated  in  the  interstices  between  the  genital  canals,  prevent  the  gonads  from  being  damaged 
or  crushed  between  the  two  layers  of  tissue.  It  seems  likely  that  the  gonadal  pits  which  have  been 
recognized  on  the  internal  surfaces  of  some  fossil  brachiopods  represent  the  points  of  attachment  of 
the  supportive  columns  rather  than,  as  Rudwick  (1970)  suggests,  the  points  at  which  the  gonad  itself 
was  attached.  Rigid  calcareous  spicules  are  present  in  many  regions  of  the  body  tissues  of 
T.  retusa,  but  are  particularly  densely  developed  above  the  gonads  thereby  providing  further 
protection. 

The  periodicity  of  larval  recruitment  was  determined  by  monitoring  the  state  of  development  of 
gonadal  tissues  throughout  the  year  (by  comparing  at  least  twenty  adult  specimens  from  each 
sample),  and  by  microscopically  examining  the  surfaces  of  all  potential  brachiopod  substrate  for 
recently  settled  post-larvae.  It  became  clear  that  the  gonads  of  T.  retusa  pass  through  two  full 
development  cycles  per  year,  culminating  in  the  release  of  gametes  in  late  spring  and  late  autumn. 
Recently  settled  post-larvae  have  only  been  observed  at  these  times,  and  the  spawning  event  must  be 
highly  synchronized,  of  short  duration,  and  affecting  the  vast  majority  of  sexually  mature 
individuals.  For  example,  the  entire  reproductive  cycle,  from  spawning  through  to  spatfall,  was 
completed  within  the  3-week  period  between  samples  collected  in  early  and  late  May  1977.  The 
precise  timing  of  the  natural  spawning  event  must  be  related  to  temperature,  as  T.  retusa  spawns  at  a 
temperature  of  10-1 1 °C  in  both  autumn  and  spring.  A similar  temperature  control  on  the  timing  of 
spawning  activity  has  been  recognized  in  both  articulate  (Rickwood  1968)  and  inarticulate 
brachiopods  (Paine  1963;  Yatsu  1902). 

To  investigate  the  spawning  behaviour  in  greater  detail,  ripe  individuals  collected  immediately 
prior  to  one  of  the  autumn  spawning  periods  were  induced  to  spawn  in  the  laboratory  by  rapid 
controlled  fluctuations  of  temperature  and  the  addition  of  mobile  mature  sperm  teased  from  males. 


234 


PALAEONTOLOGY,  VOLUME  25 


Temperature  variation  is  a common  method  of  inducing  spawning  in  marine  invertebrates 
(P.  Redfern,  pers.  comm.  1978)  because  the  resulting  stress  stimulates  the  release  of  gametes.  This 
method  had  previously  been  used  by  the  author  and  Dr.  J.  Richardson  in  a successful  attempt  to 
induce  spawning  in  the  Recent  New  Zealand  brachiopod  Terebratella  inconspicua  (Sowerby).  It 
seems  that  synchronization  of  spawning  in  neighbouring  specimens  depends  upon  the  presence  of 
male  sperm,  or  perhaps  hormones  secreted  along  with  it,  in  the  surrounding  sea-water;  it  was 
noticeable,  both  with  T.  inconspicua  and  Terebratulina  retusa,  that  spawning  commenced  shortly 
after  suspensions  of  sperm  had  been  drawn  into  the  brachial  cavity  by  feeding  current.  Once 
spawning  is  initiated  the  ova  and  sperm  are  moved  from  the  gonads  to  the  brachial  cavity  via  the 
metanephridia,  and  then  extruded  from  the  valves  by  a series  of  snapping  movements  similar  to  those 
used  during  the  ejection  of  faecal  or  pseudofaecal  pellets  (Rudwick  1970). 

Fertilization  occurs  both  in  the  brachial  cavity  of  females  and  on  the  surrounding  substrate,  and  a 
high  degree  of  synchronization  is  obviously  vital  as  the  sperm  remain  in  suspension  and  is  rapidly 
carried  away  from  the  breeding  population  by  bottom  currents.  The  ova,  by  contrast,  settle  in  dense 
clusters  around  the  margin  of  the  parent.  Only  a small  proportion  of  ova  ejected  during  the 
laboratory  experiment  were  fertilized,  although  in  nature  the  proportion  may  be  greater.  Both 
fertilized  and  non-fertilized  ova  appeared  to  be  particularly  susceptible  to  bacteriological  attack, 
although  once  again  this  may  have  been  due  to  atypical  laboratory  conditions,  in  particular 
unavoidable  increases  in  sea-water  temperature  due  to  the  heat  from  the  microscope  lamps.  The  free- 
swimming  larvae  which  developed  from  ova  fertilized  in  the  laboratory  appeared  to  be  very  similar  in 
size,  shape,  and  activity  to  those  of  T.  septentrionalis  (Couthony)  from  the  coast  of  Maine  (Morse 
1873)  although  no  attempt  was  made  to  study  them  in  detail.  The  laboratory-reared  larvae  of  T. 
retusa  had  reached  an  advanced  stage  of  development  (with  a third  peduncular  segment  indicating 
settlement  was  imminent)  within  5 days  of  spawning,  which  is  in  keeping  with  the  observed  maximum 
duration  of  3 weeks  in  nature. 

During  one  of  the  natural  spawning  periods,  attempts  were  made  to  collect  pelagic  larvae  from  the 
Firth  of  Lome  using  a plankton  net  trawled  at  the  surface  and  at  an  estimated  depth  of  100  m.  No 
brachiopod  larvae  were  recovered,  indicating  perhaps  that  the  larvae  remain  close  to  the  sea-floor 
during  their  pelagic  stage.  However,  the  evidence  for  this  is  not  conclusive,  especially  as  the  pelagic 
stage  appears  to  be  of  at  most  a few  days’  or  perhaps  hours’  (see  below)  duration,  and  it  was 
impossible  to  conduct  a comprehensive  plankton  study  for  the  expected  duration  of  the  spawning 
period.  Some  pelagic  larvae  are  known  to  actively  swim  towards  light  sources  during  their  early 
development  stages  (e.g.  Paine  1963),  and  certainly  the  larvae  of  the  abyssal  inarticulate  Pelagodiscus 
atlanticus  (King)  has  been  recovered  at  the  surface  (Ashworth  1915).  An  alternative  explanation  for 
the  lack  of  brachiopod  larvae  in  the  plankton  samples  is  that  the  larvae  are  brooded  to  an  advanced 
stage  of  development  within  the  brachial  cavity  of  adult  females,  thereby  further  curtailing  the 
duration  of  the  free-swimming  stage.  Such  a phenomenon  has  been  observed  in  T.  septentrionalis 
from  the  Bay  of  Fundy,  Canada  (Webb  et  al.  1976),  and  is  of  particular  significance  as  no  special 
brood  pouches  are  developed.  Instead  the  brooded  larvae  are  simply  held  between  the  filaments  of  the 
lophophore  and  the  body  wall.  In  the  Firth  of  Lome  the  disturbance  caused  by  the  dredging  process 
would  almost  certainly  cause  the  release  of  any  brooded  larvae  within  T.  retusa  and  hence  it  will  be 
difficult  to  determine  if  such  a phenomenon  is  widespread.  However,  the  space  available  for  brooded 
larvae  within  the  brachial  cavity  of  T.  retusa,  assuming  a mechanism  similar  to  that  of  T. 
septentrionalis,  would  be  insufficient  to  hold  all  ova  shed  by  one  individual  during  a spawning  period, 
and  therefore  a proportion  must  still  be  extruded  from  the  brachial  cavity. 

The  density  of  larval  settlement  on  the  Firth  of  Lome  mussel  beds  following  each  spawning  event  is 
remarkable.  The  number  and  size  of  brachiopods  attached  to  each  mussel  collected  on  24  May  1977 
was  recorded,  and  the  percentage  of  mussels  with  attached  brachiopods  from  each  year-class  was 
plotted  (text-fig.  3).  Representatives  of  the  ‘0’  year-class  (which  in  this  sample  had  settled  within  the  3 
weeks  prior  to  collection)  and  the  1st  year-class  (the  two  1976  cohorts)  were  both  present  on  more 
than  50%  of  the  mussels  in  the  sample,  which  represents  a considerably  greater  proportion  of  the 
available  substrate,  as  the  mussels  live  in  dense  clusters  with  the  lowermost  specimens  being  less 


CURRY:  BRACHIOPOD  TEREBRATU  LIN  A FROM  SCOTLAND 


235 


text-figure  3.  Percentage  of  the  total  number  of 
specimens  of  Modiolus  modiolus  (Linnaeus) 
collected  on  24  May  1 977  which  had  brachiopods 
from  each  of  the  seven  year  classes  of  brachipods 
attached.  The  sample  yielded  127  mussels  and  554 
brachiopods;  a total  of  21  mussels  ( = 16-5%)  had 
no  attached  brachiopods. 


YEAR-CLASS 


accessible  for  brachiopod  larvae.  The  fact  that  fewer  mussels  had  ‘0’  year-class  than  1st  year-class 
brachiopods  attached  is  probably  due  to  the  difficulty  in  picking  out  the  former,  and  the  trend  of  the 
graph  would  suggest  that  more  than  50%  of  the  mussels  are  covered  at  a spawning  event.  As  many  as 
thirty  brachiopods  of  different  ages  have  been  found  on  a single  mussel,  and  on  average  each  mussel 
in  a sample  has  3-4  brachiopods  attached.  The  density  of  settlement  provides  further  indirect 
evidence  of  a short  pelagic  larval  stage,  as  the  degree  of  dispersion  increases  with  the  length  of  the 
free-swimming  stage.  Both  autumn  and  spring  spawnings  appeared  to  be  equally  successful,  although 
there  was  no  practicable  method  of  comparing  the  numbers  of  larvae  produced. 

POPULATION  STRUCTURE  AND  DYNAMICS 

Problems  such  as  biased  sampling,  selective  preservation,  and  post-mortem  transportation  have 
bedevilled  attempts  to  interpret  fossil  population  structure  and  dynamics  by  means  of  size-frequency 
distributions.  Size-frequency  distribution  does,  however,  have  considerable  potential  as  an  analytical 
tool  in  palaeontology,  provided  the  fundamental  problem  of  obtaining  a representative  sample  can 
be  overcome.  The  simplest  and  most  effective  method  of  checking  for  ‘representativeness’  is  to  collect 
a series  of  large  samples  from  the  population  in  question  and  to  compare  the  resulting  size-frequency 
distributions.  Providing  adequate  care  has  been  taken  to  eradicate  recurring  sampling  inadequacies, 
similar  results  from  a series  of  samples  can  be  considered  as  good  evidence  that  such  results  are  an 
accurate  reflection  of  the  situation  in  the  parent  population.  The  ability  to  rationalize  the  recurring 
pattern  of  size-frequency  distribution  within  the  limits  of  theoretical  population  biology  can  be 
regarded  as  a further  useful  check  on  the  validity  of  such  results.  In  this  study  both  approaches 
provided  strong  evidence  that  the  regular  samples  from  the  Firth  of  Lome  population  were 
representative,  and  certainly  the  favourable  situation  of  the  T.  retusa  population  makes  sampling 
bias  unlikely,  and  obviously  the  distortion  resulting  from  preservational  bias  is  not  pertinent  to 
studies  of  living  animals.  The  number  of  specimens  in  a sample  is  certainly  a critical  factor,  however, 
and  the  necessity  of  having  a large  sample  has  been  stressed  by  several  authors  (see  Hallam,  in 
discussion  at  end  of  Craig  and  Oertel  1966).  The  density  and  abundance  of  brachiopods  at  the  Firth 
of  Lome  locality,  combined  with  their  favourable  situation  for  sampling,  facilitated  the  collection  of 
acceptably  large  samples. 

Further  problems  have  arisen  because  of  confusion  over  the  descriptive  terminology  applied  to 
size-frequency  diagrams,  which  warrants  mention  for  the  sake  of  clarity.  In  conventional  statistical 


236 


PALAEONTOLOGY,  VOLUME  25 


usage  the  term  ‘skewed’  refers  to  the  gradually  sloping  ‘tail’  of  an  assymetrical  distribution  (Simpson 
et  al.  1960),  and  therefore  a unimodal  skewed  distribution  which  slopes  away  gradually  on  the  right- 
hand  side  (e.g.  text-figs.  4, 6)  would  be  described  as  right-skewed,  or  positively  skewed.  A left-skewed, 
or  negatively  skewed,  distribution  slopes  gradually  from  the  left-hand  side  to  a prominent  mode  on 
the  right-hand  side  of  the  diagram.  This  convention  has  been  reversed  by  some  authors,  who  relate 
‘skewed’  to  the  mode  rather  than  the  ‘tail’  (e.g.  Raup  and  Stanley  1978;  Thayer  1978).  Throughout 
this  study  the  more  conventional  usage  has  been  adopted. 

The  over-all  shape  of  a size-frequency  distribution  reflects  the  relative  proportion  of  animals  of 
different  age-groups,  and  is  an  important  diagnostic  feature.  All  samples  analysed  during  this  study 
yielded  a unimodal  right-skewed  length-frequency  histogram  (e.g.  text-fig.  4)  with  sexually  immature 
juveniles  forming  considerably  more  than  50%  of  the  sample  at  all  seasons.  For  example,  71%  of  the 
81 1 specimens  collected  in  March  1977  were  less  than  9-6  mm  in  length  (i.e.  not  more  than  2 years 
old — see  below).  This  predominance  of  juveniles  is  not  unexpected,  as  the  long-term  success  of  any 
brachiopod  population  depends  upon  the  regular  influx  of  large  numbers  of  larvae.  The  population 
structure  of  T.  septentrionalis  from  the  Bay  of  Fundy,  Canada,  is  also  unimodal  and  right-skewed  at 
all  seasons  (Noble  et  al.  1976),  although,  just  as  in  T.  retusa,  the  position  of  the  main  mode  varied 
slightly  depending  on  the  average  size  of  the  most  recently  settled  cohort. 

Significantly  there  is  no  over-all  bimodality  discernible  in  any  of  the  length-frequency  histograms 
of  T.  retusa.  A bimodal  size-frequency  distribution  is  characteristic  of  some  living  brachiopod 
populations  (e.g.  Rudwick  1962;  Doherty  1976,  1979),  and  it  is  now  generally  accepted  that  the 
secondary  (right-hand)  mode  forms  as  a result  of  the  merging  of  older  age-groups  because  of  the 
gerontic  slowing  or  cessation  of  growth.  Such  a growth  strategy  (known  as  determinate  growth 
(Doherty  1976)  because  adults  continue  to  survive  without  growing  after  attaining  their  maximum, 
or  determinate,  size)  is  very  different  to  that  of  T.  retusa , which  continues  to  grow  throughout  life. 

When  examined  in  greater  detail,  it  is  apparent  that  the  length-frequency  histograms  of  T.  retusa 
are  characterized  by  regularly  spaced  subsidiary  peaks  (e.g.  text-fig.  4).  Once  the  biannual  spawning 
season  and  its  timing  had  been  determined  by  direct  observation,  it  was  then  possible  to  interpret 
these  peaks  in  terms  of  settlement  cohorts.  The  underlying  principle  on  which  this  analysis  is  based  is 
that  each  of  the  peaks  represents  a cohort  of  brachiopods  which  settled  on  the  mussel  beds  after  one 


table  3.  Analysis  of  the  March  1977  length- 
frequency  histogram  (Fig.  4).  All  measure- 
ments are  in  mm. 

Annual  Biannual  Date  of  Year- 

Increment  Peak  Increment  Settlement  Class 

Autumn  1976 la 

Spring  19 76  1 b 

Autumn  1975 2 a 

Spring  1975  2 b 

Autumn  1974  - 3a 

Spring  1974  3 b 

...1973  4 

---1972  5 

---1971  6 

---1970  7 


^^2 .75- 
4<<%4  25' 
4<%>.75< 
5<%J8  25' 
;.5<%10.25- 
3 %^.75, 

*<2L570  ' 


text-figure  4.  Length-frequency  histogram  of  T.  retusa, 
March  1977  sample  (ZB3727-ZB3736).  la,  b,  etc.,  refer  to 
settlement  cohorts  (see  text). 


CURRY:  BRACHIOPOD  TEREBRATULINA  FROM  SCOTLAND 


237 


text-figure  5.  Width-frequency  (a)  and  height-frequency  (b)  histo- 
gram of  T.  retusa,  March  1977  sample  (ZB3727-ZB3736).  1 a,  b,  etc., 
refer  to  settlement  cohorts  (see  text). 


of  the  biannual  reproductive  events.  Such  a technique  has  been  widely  used,  and  is  a logical 
assumption  provided  that  the  animals  have  a sharply  defined  breeding  season,  and  that  the  average 
growth-rate  of  individual  cohorts  is  such  that  successive  peaks  do  not  merge  (at  least  in  the  early 
stages  of  life).  The  majority  of  these  subsidiary  peaks  can  be  recognized  in  all  histograms,  but  the 
March  1977  length-frequency  histogram  (text-fig.  4)  has  been  selected  as  a standard  and  will  be 
analysed  in  detail.  It  can  be  demonstrated  that  the  width-frequency  histogram  (text-fig.  5a),  and  to  a 
lesser  extent  the  height-frequency  histogram  (text-fig.  5b),  yield  essentially  the  same  results  as  the 
length-frequency  histogram,  but  in  T.  retusa  the  maximum  incremental  increase  in  shell  dimensions 
occurs  anteriorly  (i.e.  in  length)  throughout  ontogeny,  and  hence  the  measured  shell  width  and  height 
dimensions  are  of  smaller  absolute  value  with  the  result  that  the  resolution  between  peaks  in  the 
frequency  distribution  is  reduced.  An  alternative  method  of  analysing  length-frequency  data,  in 
which  the  data  is  plotted  on  probability  paper,  is  described  in  the  Appendix. 

The  March  1977  sample  was  collected  approximately  2 months  prior  to  the  spring  1977  spawning 
period,  and  therefore  the  first  two  peaks  on  the  left-hand  side  of  text-fig.  4 (at  2-75  mm  and  4-25  mm) 
correspond  to  the  autumn  and  spring  cohorts  in  1976,  and  are  therefore  labelled  1(a)  and  \(b) 
respectively  in  text-fig.  4 and  Table  3.  Similarly  the  paired  peaks  at  6-75  mm  (2a)  and  8-25  mm  (2b) 
represent  the  two  cohorts  which  settled  in  autumn  and  spring  1975  respectively;  the  autumn  and 
spring  cohorts  in  1974  are  likewise  clearly  represented  by  twin  peaks  at  10-25  mm  (3a)  and  1 1 -75  mm 
(3b).  Each  of  these  pairs  of  peaks  have  identical  spacing  (i.e.  1 -5  mm— ‘Biannual’  Increment  in  Table 
3),  and  the  interval  between  cohorts  which  settled  1 year  apart  remains  remarkably  constant  (i.e. 
approx.  4 mm— Annual  Increment  in  Table  3).  Clearly  the  rate  of  growth  of  T.  retusa  remains 
constant  throughout  the  first  3 years  of  life,  at  least  within  the  limits  of  resolution  of  this  method  of 
mensuration.  Another  striking  feature  of  this  analysis  is  that  each  of  the  biannual  spawning  periods 
in  the  previous  3 years  has  been  highly  successful,  indicating  that  this  relatively  deep-water 
population  is  not  subject  to  environmental  disturbances  known  to  cause  aperiodic  disruption  to  the 


238 


PALAEONTOLOGY,  VOLUME  25 


text-figure  6.  Length-frequency  histograms  of  T.  retusa  a,  5 May  1977  sample 
(ZB3737-ZB3746).  B,  24  May  1977  sample  (ZB3747-ZB3756).  c,  January  1979  sample 
(ZB3717-ZB3726).  D,  July  1977  sample  (ZB3757-ZB3766).  E,  August  1977 
sample  (ZB3767-ZB3776). 


CURRY:  BRACHIOPOD  TEREBRA  TULIN  A FROM  SCOTLAND 


239 


reproductive  cycle  of  shallow-water  marine  invertebrates.  As  mentioned  above,  the  spawning  events 
observed  subsequently  during  1977-1979  appeared  to  be  equally  successful. 

The  position  and  spacing  of  the  remaining  peaks  in  text-fig.  4,  indicates  that  the  growth-rate 
progressively  decreases  from  the  third  year  of  life  onwards,  resulting  in  the  merging  of  the  paired 
peaks  to  form  a single  broad  peak  representing  the  biannual  cohorts  in  1973  (14-73  mm— labelled  4 in 
text-fig.  4 and  Table  3)  and  in  1972  (17-25  mm— labelled  5).  There  is  no  peak  corresponding  to  the 
1971  cohorts  in  text-fig.  4,  but  such  a peak  is  present  at  19-75  mm  in  the  5 May  1977  sample  (text-fig. 
6a)  and  such  a value  is  included  in  Table  3 for  the  sake  of  completeness  (as  described  below  there  is 
very  little  movement  of  peaks  between  March  and  May).  Ambiguity  over  the  position  of  the  peaks 
corresponding  to  the  older  age-groups  is  not  surprising  considering  the  rarity  of  specimens— the 
grouping  tentatively  identified  as  representing  the  1970  cohorts  in  text-fig.  4 (i.e.  21-75  mm— labelled 
7)  includes  a mere  two  specimens  out  of  a total  sample  of  81 1 . However,  similar  groupings  of  slightly 
more  specimens  do  occur  in  other  samples,  and  its  interpretation  as  a 7th  year-class  can  also  be 
justified  on  the  basis  of  the  spacing  between  previous  peaks  in  text-fig.  4 (see  Table  3). 


table  4.  T.  retusa  age-groups  in  the  March  1977 
sample;  size  range  refers  to  maximum  shell  length 
measured  in  mm. 


Age  Year  of  Size 

Group  Settlement  Range 

1 1976  up  to  5.5 

2 1975  5.6  - 9.5 

3 1974  9.6-  13-5 

4 1973  13.6  - 16.0 

5 1972  16.1-  18.5 

6 1971  18.6  - 20.0 

7 1970  over  20.0 


No.  of  °/0  of  °/ o Rate  of 


Specimens  Total  Mortality 


365  45 

213  26  42 

121  15  43 

63  8 45 

40  5 40 

7 09  83 

2 0.2  71 


Following  on  from  this  analysis  it  was  possible  to  divide  the  March  1977  sample  into  age-groups 
(=  year-classes)  on  the  basis  of  a range  of  lengths  (Table  4).  The  boundaries  of  the  age-groups  are 
somewhat  arbitrary,  and  the  higher  and  lower  values  in  each  grouping  include  specimens  which 
rightly  belong  in  the  age-groups  on  either  side.  Nevertheless,  the  exercise  is  useful  as  it  quantifies  the 
relative  proportions  of  age-groups  and  provides  an  estimation  of  mortality-rates. 

Comparison  with  theoretical  model.  One  of  the  most  significant  contributions  to  the  study  of 
population  structure  and  dynamics  in  recent  years  has  been  the  comprehensive  theoretical 
simulations  of  Craig  and  Oertel  (1966).  Having  identified  five  main  factors  which  influence  the  over- 
all shape  of  a size-frequency  distribution,  Craig  and  Oertel  programmed  a computer  to  produce  an 
exhaustive  series  of  histograms  using  various  combinations  of  these  five  factors.  Having  established 
the  characteristic  population  structure  of  T.  retusa  in  nature,  it  was  obviously  of  interest  to  identify 
which  of  Craig  and  Oertel’s  experiments  was  conducted  using  a combination  of  factors  similar  to 
those  prevailing  in  the  Firth  of  Lome  population,  and  to  compare  the  theoretical  simulation  with  the 
actual  population  structure. 

Craig  and  Oertel’s  five  factors  are  (1)  recruitment  strategy,  (2)  growth-rate,  (3)  coefficient  of 
variation  of  growth-rate,  (4)  mortality-rate,  (5)  cessation  of  growth.  T.  retusa  clearly  falls  within  the 
‘boreal’  recruitment  strategy  of  Craig  and  Oertel,  which  was  defined  as  two  short  spawning  periods  in 
autumn  and  spring.  The  growth-curve  of  T.  retusa  is  intermediatory  between  the  ‘linear’  and  the 
‘high-to-low’  curves  used  by  Craig  and  Oertel,  but  is  closer  to  the  former  than  the  latter.  The  third 
factor,  the  coefficient  of  variation  of  growth-rate,  was  an  attempt  to  allow  for  the  effects  of  the 
varying  growth-rates  of  individuals  in  a single  cohort.  There  is  no  indication  of  the  extent  of  such 
variability  in  T.  retusa,  but  Craig  and  Oertel  mostly  used  a coefficient  of  2,  which  they  believed  to  be 
an  acceptable  average  value.  The  mortality-rate  of  T.  retusa  remains  constant,  at  least  during  the  first 
5 years  of  life  (see  below).  The  final  factor  related  to  the  timing  and  duration  of  any  cessation  of 
growth  during  the  year,  and  there  is  good  evidence  that  T.  retusa  falls  within  Craig  and  Oertel’s 
category  (a),  namely  a winter  stoppage  of  3 months’  duration  (see  below). 

This  permutation  of  factors  was  combined  in  Craig  and  Oertel’s  experiment  number  34,  and  the 
resulting  size-frequency  histogram  (Craig  and  Oertel,  1966,  fig.  17,  p.  346)  is  strikingly  similar  to  the 
length-frequency  histograms  of  T.  retusa.  The  over-all  shape  of  all  these  distributions  is  identical. 


240 


PALAEONTOLOGY,  VOLUME  25 


being  unimodal  and  right-skewed.  The  computer  simulation  had  evenly  spaced  peaks  corresponding 
to  the  biannual  cohorts,  which  merge  into  a single  annual  peak  in  later  life,  exactly  as  in  the  natural 
population.  Craig  and  Oertel  described  this  experiment  as  follows: 

Boreal  recruitment,  three  winter  months  cessation  of  growth  with  doubling  of  mortality,  coefficient  of 
variation  2 . . . growth-rate  linear  . . . mortality  constant. 

Boreal  recruitment  consists  of  two  equal  waves  in  late  spring  and  early  autumn,  separated  by  a short  summer 
interval  and  a long  winter  interval.  This  forms  twin  peaks  in  the  living  population ...  the  groups  of  twin  peaks  are 
equidistant,  and  the  twins  have  identical  intervals.  This  peak  spacing  is  diagnostic  for  linear  growth. 

Obviously  there  are  differences  in  detail  between  the  theoretical  and  the  actual,  especially  as  the 
critical  factors  in  nature  tend  to  vary  slightly  with  increasing  age  rather  than  remain  constant. 
Nevertheless,  the  results  of  this  comparison  are  very  encouraging,  and  demonstrate  the  potential  of  a 
combined  theoretical  and  empirical  approach  to  the  study  of  population  structure  and  dynamics. 


Annual  Biannual  Year- 

Increment  Peak  Increment  Class 


table  5.  Analysis  of  the  peak  spacing  in  the  width- 
frequency  histogram  of  Rostricellula  rostrata, 
Ulrich  and  Cooper  (data,  in  mm,  are  from  Walker 
and  Parker  1976). 


Relevance  for  fossil  populations.  The  analysis  of  the  population  structure  and  dynamics  of  fossil 
brachiopods  is  best  conducted  using  both  size-frequency  histograms  and  growth-line  counts,  as  the 
inherent  deficiences  of  one  method  are  compensated  for  by  the  strengths  of  the  other.  However,  under 
ideal  conditions  of  preservation,  a frequency  histogram  prepared  from  a sample  of  fossil  brachiopods 
can,  on  its  own,  yield  precise  data  on  the  autecology  of  that  species.  In  preparing  a width-frequency 
histogram  from  a sample  of  the  Middle  Ordovician  species  Rostricellula  rostrata,  Ulrich  and  Cooper 
from  Tennessee,  Walker  and  Parker  (1976)  used  the  same  grouping  adopted  for  the  T.  retusa 
histograms  (i.e.  0-5  mm),  and  hence  the  prominent  peaks  in  the  R.  rostrata  histogram  can  be  analysed 
in  the  usual  tabular  form  (i.e.  Table  5)  allowing  direct  comparison  with  the  living  population  (i.e. 
Table  3). 

This  method  of  analysis  emphasizes  the  remarkable  similarity  in  the  population  structure  between 
the  living  and  fossil  population,  especially  in  the  pattern  of  peak  spacing.  Both  histograms  have 
prominent  regularly  spaced  twin  peaks,  with  the  separation  between  each  peak  in  a twin  (i.e.  the 
‘biannual  increment’— see  Tables  5 and  3)  being  identical  in  both  populations  (i.e.  1 -5  mm).  Just  as  in 
the  T.  retusa  histogram,  these  twin  peaks  are  not  discernible  on  the  right-hand  side  of  the  diagram, 
because  the  slowing  of  the  growth-rate  in  older  specimens  has  resulted  in  the  merging  of  the  twins  to 
form  a single  peak  representing  the  total  annual  recruitment  (e.g.  at  14-25  mm  in  Table  5).  There  are 
differences  in  the  absolute  values  of  comparable  increments  (e.g.  annual  increments— compare 
column  1 in  Tables  5 and  3)  and  the  fossil  species  had  a shorter  life  span  and  a more  rapidly  decreasing 
growth-rate,  but  such  differences  are  of  minor  significance  compared  to  the  similarity  in  the  over-all 
pattern  of  peak  spacing.  Both  theoretical  and  empirical  data  indicate  that  this  pattern  of  regularly 
spaced  twin  peaks  is  characteristic  of  animals  inhabiting  temperate  latitudes,  and  the  fossil 
population  can  reasonably  be  assumed  to  have  experienced  broadly  similar  seasonal  cycles  of 
temperature,  food  supply,  etc.,  to  those  prevalent  today  in  the  Firth  of  Lome  or  other  temperate 
habitats.  Apart  from  comparison  with  equivalent  living  populations,  the  reconstruction  of 
population  structure  and  dynamics  in  fossil  brachiopods  would,  under  ideal  circumstances,  be  based 
on  the  analysis  of  several  large  samples  rather  than  just  one,  and  would  also  take  into  consideration 
pertinent  localized  environmental  and  biological  factors;  as  mentioned  above,  the  interpretation  of 
growth-lines  can  further  refine  the  resulting  data.  Nevertheless,  the  success  of  this  preliminary 
comparison  between  the  population  structure  of  living  and  fossil  brachiopods  augers  well  for  this 
technique. 


CURRY:  B R ACHIOPOD  TEREBRATU  LIN  A FROM  SCOTLAND 


241 


Seasonal  growth  pattern.  To  obtain  a more  precise  picture  of  the  pattern  of  seasonal  growth  in  T. 
retusa,  the  position  of  peaks  corresponding  to  individual  cohorts  was  recorded  in  samples  collected 
throughout  the  year.  The  results  (Table  6)  suggest  that  the  growth-rate  of  adults  decreases 
significantly  in  winter  months,  as  the  peaks  appear  to  remain  stationary  between  January  and  May. 


table  6.  Movements  of  peaks  corresponding  to  the  4th 
and  5th  year-classes  of  T.  retusa  (measurements,  in  mm, 
refer  to  maximum  shell  length). 


Age  Expected 

Group  Mar  May  Aug  Oct  Jan  in  Mar 

4 14.75  14.75  16.25  17.25  1725  17.25 

5 17.25  17.25  18.25  18.75  19.75  19.75 


The  evidence  for  this  is  based  on  the  assumption  that  the  mid-point  of  a particular  peak  is  an 
acceptable  measure  of  the  average  length  of  specimens  in  a cohort;  although  the  growth-history  of 
individuals  within  a cohort  will  vary  depending  on  localized  environmental  conditions.  The  slowing 
of  growth  in  winter  is  an  expected  and  readily  explained  phenomenon,  bearing  in  mind  the  direct 
relationship  between  decreasing  temperature  and  a reduction  of  the  rate  of  metabolic  activity  in  cold- 
blooded invertebrates.  It  seems  likely  that  the  rate  of  growth  of  T.  retusa  slows  progressively  during 
the  autumn  and  winter,  and  reaches  a minimum,  or  perhaps  ceases  altogether,  in  mid-winter  when 
temperatures  are  lowest  and  food  supplies  minimal.  There  is  some  evidence  that  sexually-mature 
females  have  a slightly  slower  rate  of  growth  than  mature  males  (text-fig.  7a),  perhaps  indicating  that 
the  production  of  ova  is  more  demanding  in  terms  of  available  nutrients.  However,  the  pattern  of 
modes  in  text-fig.  7a  is  rather  confused  and  far  from  unequivocal,  and  this  facet  of  sexual  dimorphism 
does  not  appear  to  be  noticeable  in  the  combined  length-frequency  distributions.  The  increments 
involved,  like  those  produced  by  variable  growth-rates  in  a single  cohort,  are  so  small  as  to  be  outside 
the  limits  of  resolution  of  the  length-frequency  histogram. 

In  contrast,  the  growth  of  recently  settled  post-larvae  appears  to  proceed  rapidly  during  the  first  3 
months  of  settled  life,  and  to  be  independent  of  prevailing  environmental  conditions.  For  example, 
the  cohort  which  settled  in  late  autumn  1976  had  attained  an  average  length  of  2-75  mm  by  the 
following  March  (i.e.  \{a)  in  text-fig.  4),  a rate  of  growth  greater  than  at  any  other  stage  of  life. 
Similarly,  the  cohort  which  settled  in  May  1977  had  grown  to  an  average  length  of  2-75  m by  the 
following  August  (compare  text-figs.  6b,  6e).  The  growth-history  of  T.  retusa,  therefore,  has  three 
distinct  phases:  (1)  approximately  3 months  of  rapid  growth  immediately  following  settlement  in 
both  autumn  and  spring;  (2)  the  remainder  of  the  first  3 years  of  life,  when  the  annual  growth-rate 
remains  constant,  but  varies  seasonally  depending  upon  ambient  sea-water  temperatures;  (3)  the 
remainder  of  life,  during  which  annual  growth-rates  decrease  progressively. 

The  initial  period  of  rapid  post-settlement  growth  is  a prudent  feature,  as  the  early  development  of 
a rigid  protective  exo-skeleton  will  reduce  mortality  rates  during  this  most  vulnerable  stage  of  settled 
life.  The  ability  of  the  autumn  cohorts  to  grow  in  winter  months  indicates  perhaps  that  their  modest 
nutrient  requirements  are  adequately  satisfied  by  even  the  reduced  winter  food  supplies.  It  may  also 
be  of  significance  that  the  volume  of  the  feeding  ‘chambers’  in  juveniles  is  considerably  greater  than 
that  enclosed  by  the  valves  of  the  shell,  as  the  filaments  of  the  lophophore  extend  a considerable 
distance  beyond  the  shell  margin  when  feeding.  In  adults  the  filaments  are  almost  entirely  contained 
within  the  brachial  cavity  when  feeding;  in  juveniles  the  increased  risk  of  predation  resulting  from  the 
exposed  filaments  and  widely  gaping  shell  (as  much  as  45°  as  compared  to  « 15°  in  adults)  may  be 
offset  by  the  advantages  gained  by  the  relatively  rapid  shell  growth  due  to  the  proportionately  high 
levels  of  nutrient  intake.  The  reduction  in  the  rate  of  annual  growth-rate  at  the  end  of  3 years 
coincides  with  the  onset  of  sexual  maturity,  and  probably  reflects  a fundamental  change,  with  the 
nutritional  requirements  of  the  developing  gonads  receiving  precedence  over  other  metabolic 
processes. 

Growth-line  analysis.  The  analysis  of  the  pattern  of  growth-line  formation  in  brachiopods  is  best 
conducted  on  a cumulative  basis  using  a large  number  of  individuals  from  a single  population.  The 
main  difficulty  of  attempting  a growth-line  analysis  of  an  individual  is  that  the  pattern  is  often 
incomplete,  and  hence  difficult  to  interpret.  Nevertheless,  a sample  analysis  of  a single  adult  specimen 


242 


PALAEONTOLOGY,  VOLUME  25 


text-figure  7.  A,  sexual  dimorphism  in  shell  growth  as  determined 
from  342  sexually  mature  specimens  of  T.  retusa  collected  26  October 
1977  from  the  Firth  of  Lome  population,  b,  c,  age  pyramids  for  T. 
retusa  population  in  the  Firth  of  Lome  before  (b,  from  March  1977 
sample)  and  after  (c,  from  24  May  1977  sample)  spatfall.  D,  Growth 
curve  for  T.  retusa  as  determined  from  the  analysis  of  the  March  1977 
sample  (i.e.  Table  3). 


of  T.  retusa  with  a complete  record  of  growth-line  formation  is  included  at  this  stage  as  it  contributes 
further  to  our  knowledge  of  the  growth  history  of  the  population.  The  specimen  (ZB  3717)  was 
collected  during  January  1979  at  an  approximate  depth  of  165  m from  the  Firth  of  Lome,  and  was 
15-5  mm  in  length.  Having  measured  the  distance  to  each  growth-line  from  the  posterior  margin  of 
the  shell  along  the  medial  axis  of  the  pedicle  valve  (Table  7),  and  determined  the  spacing  between 
individual  growth-lines  (column  2 in  Table  7),  it  was  then  apparent  that  growth-lines  form  biannually. 
The  age-group  analysis  in  Table  4 indicates  that  a specimen  of  length  15-5  mm  should  be  4 years  old;  a 
conclusion  which  is  confirmed  by  growth-line  analysis  which  indicates  that  the  1st  growth-line  on  this 
specimen  was  formed  in  autumn  1975  (Table  7)  and  therefore  it  must  have  settled  in  spring  1975. 

The  spacing  between  growth-lines  (column  2 in  Table  7)  on  this  specimen  indicates  that 
approximately  two-thirds  of  the  annual  growth  occurs  during  the  ‘summer’  period,  whilst  the 


CURRY:  BRACHIOPOD  TEREBRATULINA  FROM  SCOTLAND 


243 


remaining  one-third  occurs  during  ‘winter’.  The  exact  timing  of  growth-line  formation  has  not  been 
determined,  but  is  assumed  to  occur  in  autumn  and  spring,  and  at  times  of  pronounced 
environmental  and  physiological  stress.  If  so,  the  ‘summer’  and  ‘winter’  periods  would  therefore  be  of 
approximately  6 months’  duration,  and  would  probably  be  more  accurately  designated  as  ‘summer/ 
autumn’  and  ‘winter/spring’  respectively. 


table  7.  Analysis  of  growth  lines  on  single  specimen  of  T. 
retusa  (ZB3717)  collected  from  the  Firth  of  Lome  on  21 
January  1979  (data  in  mm). 


Increment  Growth-line  Formation 


-Autumn  1975 
-Spring  1976 
-Autumn  1976 
-Spring  1977 

- Autumn  1977 
-Spring  1978 

- Autumn  1978 
-Spring  1979 


Mortality.  The  data  in  Table  4 indicates  that  roughly  40%  of  the  brachiopods  in  each  age-group  die 
per  year.  This  estimation  applies  only  to  specimens  more  than  1 year  old,  as  there  is  no  viable  method 
of  determining  the  mortality  rate  during  the  periods  prior  to  and  immediately  following  settlement. 
Significantly  higher  mortality-rates  are  likely  to  occur  during  these  early  stages  of  life,  as  has  been 
determined  in  other  living  brachiopod  populations  (e.g.  Doherty  1976,  1979).  As  the  number  of 
specimens  in  the  older  age-groups  is  so  small,  the  significant  increase  in  mortality-rates  amongst  the 
6th  and  7th  age-group  (Table  4)  may  not  necessarily  be  representative,  although  older  specimens  may 
indeed  be  more  susceptible  to  disease,  stress,  etc. 

The  causes  of  mortality  in  the  T.  retusa  population  are  not  apparent,  although  it  was  not  possible  to 
examine  large  numbers  of  dead  shells  for  signs  of  predation.  There  is,  however,  very  little  evidence  of 
repaired  shell  damage  in  living  specimens,  which  suggests  that  the  level  of  predation  is  low.  Probably 
a large  proportion  of  deaths  occur  in  winter  because  of  the  stress  caused  by  less  favourable 
environmental  conditions.  It  may  be  significant  that  Craig  and  Oertel  (1966)  incorporated  a doubling 
of  the  mortality  rate  in  winter  in  the  experiment  which  yielded  a size-frequency  distribution  similar  to 
that  of  T.  retusa.  The  fact  that  a few  adult  specimens  did  not  develop  gonads  during  spawning  periods 
may  be  symptomatic  of  diseases  or  infections  which  may  account  for  a small  proportion  of  the  annual 
mortality.  However,  as  gonad  development  is  discernible  during  winter  months,  it  would  appear  that 
starvation  is  not  a major  cause  of  death. 

The  small  tissue  content  of  brachiopods  probably  partially  explains  the  apparent  lack  of  predators 
on  T.  retusa , although  predatory  gastropod  borings  have  been  found  in  other  species  (e.g.  Logan 
1979).  Because  of  their  low  nutritional  value,  it  may  indeed  be  more  reasonable  to  look  for  potential 
predators  at  the  ‘micro’  rather  than  the  ‘macro’  level,  and  certainly  carnivorous  micro-organisms 
could  readily  gain  access  to  the  brachial,  and  perhaps  body,  cavity  via  the  feeding  currents.  Some 
nematoid  worms  do  appear  to  feed  on  brachiopod  lophophoral  filaments  (McCammon  1971) 
although  there  has  been  no  detailed  study  of  the  extent  and  effect  of  such  predation/parasitism. 


CONCLUSIONS 

The  large  population  of  T.  retusa  in  the  Firth  of  Lome  is  well  established,  and  the  species  is  relatively 
abundant  in  deeper  waters  off  the  west  coast  of  Scotland.  The  main  reasons  for  this  success  are  the 
recurring  efficiency  of  the  reproductive  activity  and  the  absence  of  competing  organisms  or  readily 
apparent  predators.  Representatives  of  the  genus  Terebratulina  have  a long  history,  and  have  been 
present  in  the  North  Atlantic  since  its  inception.  It  survives  as  one  of  the  most  cosmopolitan  and 
abundant  of  living  brachiopod  genera,  occurring  in  all  oceans.  The  longevity  and  success  of  this 
genus  are  probably  due  to  an  adaptability  of  pedicle  morphology  which  greatly  increases  its  range  of 
substrates,  combined  with  an  ability  to  colonize  habitats  (such  as  the  Firth  of  Lome  depression) 
inimical  to  other  more  ‘successful’  epibenthonic  organisms.  Critical  factors  in  this  latter  capability  may 


244 


PALAEONTOLOGY,  VOLUME  25 


be  a generally  low  level  of  nutrient  requirements,  and  the  ability  to  survive  prolonged  periods  of 
adverse  conditions  by  a virtual  cessation  of  metabolic  processes,  as  demonstrated  by  the  specimens 
which  survive  in  the  outside  aquarium  at  Dunstaffnage  despite  being  subjected  to  rapid  daily 
temperature  fluctuations.  Clearly  another  important  factor  in  the  longevity  and  success  of  the  genus 
is  its  remarkable  morphological  conservatism,  with  Cretaceous  and  Recent  species  being  almost 
indistinguishable.  T.  retusa  is  unlikely,  however,  to  reassume  the  role  of  its  ancestors  as  a common 
constituent  of  shallow  marine  ecosystems  because  of  its  inability  to  compete  for  available  space  in 
such  environments. 

Acknowledgements.  I am  indebted  to  my  joint  supervisors  Dr.  C.  H.  C.  Brunton  and  Dr.  P.  Wallace  for  their 
advice,  guidance,  and  encouragement  throughout  the  course  of  this  study.  Other  staff  of  the  British  Museum 
(Natural  History)  also  contributed  greatly;  in  particular  I would  like  to  acknowledge  Dr.  L.  R.  M.  Cocks, 
Mr.  E.  F.  Owen,  Mr.  A.  Rissone,  Dr.  M.  K.  Howarth,  Mrs.  H.  Brunton,  Miss  L.  Cody,  and  Mrs.  P.  P. 
Hamilton-Waters.  Dr.  A.  Ansell  kindly  supervised  my  work  at  the  Oban  Laboratory,  and  I would  also  like  to 
thank  Mr.  C.  Comley,  Mrs.  L.  Robb,  Mr.  and  Mrs.  R.  Harvey,  Mr.  S.  Knight,  and  the  Captains  and  crews  of  the 
R/V  Calanus  and  Seol  Mara  for  their  willing  co-operation.  I am  also  grateful  to  Dr.  A.  Williams  for  his  interest 
and  support.  The  work  was  carried  out  during  the  tenure  of  a Department  of  Education  (Northern  Ireland) 
Postgraduate  Research  Studentship,  which  is  gratefully  acknowledged. 

REFERENCES 

ashworth,  j.  h.  1915.  On  the  larvae  of  Lingula  and  Pelagodiscus  (Discinisca).  Trans.  R.  Soc.  Edin.  51,  45-69. 
atkins,  d.  1959a.  The  growth  stages  of  the  lophophore  of  the  brachiopods  Platidia  davidsoni  (Eudes 
Deslongchamps)  and  P.  anomioides  (Phillippi)  with  notes  on  the  feeding  mechanism.  J.  mar.  biol.  Ass.  U.K.  38, 
103-132. 

19596.  The  early  growth  stages  and  adult  structure  of  the  lophophore  of  Macandrevia  cranium  (Muller), 

(Brachiopoda,  Dallinidae).  Ibid.  335-350. 

— 1960a.  A new  species  of  Brachiopoda  from  the  Western  Approaches,  and  the  growth  stages  of  the 
lophophore.  Ibid.  39,  71-89. 

— 19606.  A note  on  Dallina  septigera  (Loven),  (Brachiopoda,  Dallinidae).  Ibid.  91-99. 

— 1 960c.  The  ciliary  feeding  mechanism  of  the  Megathyridae  (Brachiopoda)  and  the  growth  stages  of  the 
lophophore.  Ibid.  459-479. 

— 1961.  The  growth  stages  and  adult  structure  of  the  lophophore  of  the  brachiopods  Megerlia  truncata  (L.) 
and  M.  echinata  (Fischer  and  Oehlert).  Ibid.  41,  95-111. 

barber,  p.  l.,  dobson,  m.  r.  and  Whittington,  R.  J.  1979.  The  geology  of  the  Firth  of  Lome,  as  determined  by 
seismic  and  live  sampling  methods.  Scott.  J.  Geol.  15,  217-230. 
brunton,  c.  h.  c.  and  curry,  G.  b.  1979.  British  brachiopods  Synopses  of  the  British  Fauna  (New  Series),  No.  17, 
64  pp.  Linnean  Society  and  Academic  Press,  London. 

cerrato,  m.  1980.  In  rhoads,  D.  c.  and  lutz,  R.  a.  (eds.).  Skeletal  growth  of  aquatic  organisms,  pp.  417-465, 
Plenum  Press. 

chumley,  j.  1918.  The  fauna  of  the  Clyde  sea  area.  Glasgow  Univ.  Press.  200  pp. 

comley,  c.  A.  1978.  Modiolus  modiolus  (L.)  from  the  Scottish  west  coast.  I.  Biology.  Ophelia , 17,  167-193. 
craig,  G.  Y.  and  oertel,  G.  1966.  Deterministic  models  of  living  and  fossil  populations  of  animals.  Q.  Jl  geol. 
Soc.  Lond.  122,  315-355. 

davidson,  t.  1886-1888.  A monograph  of  Recent  Brachiopoda.  Trans.  Linn.  Soc.  Ser.  2,  4,  pts.  I— III. 
doherty,  p.  j.  1976.  Aspects  of  the  feeding  ecology  of  the  brachiopod  Terebratella  inconspicua  (Sowerby). 
Unpublished  M.Sc.  thesis,  Univ.  of  Auckland. 

— 1979.  A demographic  study  of  a subtidal  population  of  the  New  Zealand  articulate  brachiopod 
Terebratella  inconspicua.  Mar.  Biol.  52,  331-342. 

Hancock,  a.  1859.  On  the  organisation  of  the  Brachiopoda.  Phil.  Trans.  R.  Soc.  (for  1858),  148,  791-869. 
Harding,  J.  p.  1949.  The  use  of  probability  paper  for  the  graphical  analysis  of  polymodal  frequency 
distributions.  J.  mar.  Biol.  Ass.  U.K.  28,  141-153. 

logan,  a.  1979.  The  Recent  Brachiopoda  of  the  Mediterranean  Sea.  Bull,  de  Tlnstitut  oceanograph.,  Monaco,  72, 
no.  1434.  112  pp. 

mccammon,  h.  m.  1971 . Behaviour  in  the  brachiopod  Terebratulina  septentrionalis  (Couthouy).  J.  Exp.  mar.  biol. 
Ecol.  6,  35-45. 


CURRY:  BRACHIOPOD  TEREBRATULINA  FROM  SCOTLAND 


245 


moore,  h.  b.  1958.  Marine  ecology.  John  Wiley  & Sons,  New  York. 

morse,  E.  s.  1873.  Embryology  of  Terebratulina,  Mem.  Bos.  Soc.  nat.  Hist.  II,  249-264. 

noble,  j.  p.  A.,  logan,  A.  and  webb,  G.  r.  1976.  The  Recent  Terebratulina  community  in  the  rocky  subtidal  zone 
of  the  Bay  of  Fundy.  Lethaia,  9,  1-17. 

paine,  R.  t.  1963.  Ecology  of  the  brachiopod  Glottidia  pyramidata.  Ecol.  Monog.  33,  255-280. 
prosser,  c.  l.  1973.  Comparative  animal  physiology  (3rd  edn.).  W.  B.  Saunders,  Philadelphia. 
raup,  d.  m.  and  Stanley,  s.  m.  1978.  Principles  of  paleontology  (2nd  edn.).  Freeman  & Company,  San  Francisco. 
481  pp. 

rickwood,  a.  e.  1968.  A contribution  to  the  life-history  and  biology  of  the  brachiopod  Pumilus  antiquatus , 
Atkins.  Trans.  R.  Soc.  N.Z.  (Zool.),  10,  163-182. 
rudwick,  m.  j.  s.  1962.  Notes  on  the  ecology  of  brachiopods  in  New  Zealand.  Ibid.  1,  327-335. 

1970.  Living  and  fossil  brachiopods.  Hutchinson,  London.  199  pp. 

simpson,  G.  G.,  roe,  A.  and  lewontin,  r.  c.  1960.  Quantitative  zoology  (revised  edn.).  Harcourt,  Brace  & Co., 
New  York.  440  pp. 

thayer,  c.  w.  1975.  Size-frequency  and  population  structure  of  brachiopods.  Palaeogeog.,  Palaeoclim ., 
Palaeoecol.  17,  139-148. 

walker,  K.  R.,  and  Parker,  w.  c.  1976.  Population  structure  of  a pioneer  and  a later  stage  species  in  an 
Ordovician  ecological  succession.  Paleobiol.  2,  191-201. 
webb,  G.  R.,  logan,  a.  and  noble,  j.  p.  a.  1976.  Occurrence  and  significance  of  brooded  larvae  in  a Recent 
brachiopod.  Bay  of  Fundy,  Canada.  J.  Paleont.  50,  869-870. 
wesenberg-lund,  e.  1940.  Brachiopoda  in  the  zoology  of  east  Greenland.  Meddr  Gron.  121  (5),  12  pp. 

1941.  Brachiopoda.  Danish  Ingolf  expedition,  4,  (12),  17  pp. 

yatsu,  n.  1902.  On  the  habits  of  the  Japanese  Lingula.  Annot.  Zool.  Japan,  4,  61-67. 

G.  B.  CURRY 

Department  of  Geology 

Original  typescript  received  14  October  1980  ^he  University 

Glasgow 

Final  typescript  received  10  January  1981  G12  8QQ 


APPENDIX.  The  use  of  probability  graph  paper  for  the  analysis  of  polymodal  length-frequency  histograms 

By  way  of  an  additional  check  on  the  interpretation  of  the  population  structure  of  T.  retusa,  the  length-frequency 
data  used  to  construct  text-fig.  4 was  also  plotted  on  probability  graph  paper  (text-fig.  8).  By  this  means  it  is 


0.01  0.1  1 10  50  90  99  99  9 99  99 

CUMULATIVE  PERCENTAGE  — > 


text-fig.  8.  Length-frequency  data  collected  22  March  1977  («  = 811) 
plotted  on  probability  graph  paper. 


246 


PALAEONTOLOGY,  VOLUME  25 


possible  to  check  that  text-fig.  4 is  indeed  composed  of  a series  of  more  or  less  overlapping  normal  distributions 
corresponding  to  individual  settlement  cohorts.  Harding  (1949)  was  the  first  to  realize  that  the  ability  of 
probability  paper  to  pick  out  individual  normal  distributions  in  a polymodal  distribution  would  be  of  great  use 
to  population  biologists,  but  few  workers  have  made  use  of  this  method  (Cerrato  1980).  The  methodology  of  this 
method  of  analysis  is  straightforward.  The  data  is  plotted  as  a cumulative  percentage  (i.e.  % of  total  sample  less 
than  x mm)  on  a non-linear  horizontal  scale,  which  is  so  arranged  that  a plot  of  points  corresponding  to  a normal 
distribution  yields  a straight  line.  As  Harding  (1949)  demonstrated,  a polymodal  distribution  plots  out  as  a series 
of  straight  lines  corresponding  to  each  of  the  constituent  normal  distributions.  Although  there  is  no  information 
as  to  the  size  distribution  of  animals  in  each  settlement  cohort  in  the  Firth  of  Lome,  it  is  a reasonable  and  widely 
accepted  assumption  that  each  cohort  would  plot  out  as  an  essentially  normal  distribution.  By  this  method  of 
analysis,  therefore,  it  is  possible  to  check  that  each  peak  in  text-fig.  4 does  represent  a settlement  cohort  (or 
amalgamation  of  two  cohorts  in  the  adult  specimens)  as  suggested  in  the  main  text. 

When  the  March  1977  data  is  plotted  in  this  manner  the  results  (text-fig.  8)  confirm  the  interpretation  of 
population  structure  outlined  in  the  text.  Short  discrete  lines  are  clearly  distinguishable  amongst  juvenile 
specimens,  and  correspond  to  biannual  cohorts.  The  degree  of  differentiation  is  less  amongst  3-  and  4-year-old 
specimens,  which  is  consistent  with  growth-rates  decreasing  following  the  onset  of  sexual  maturity.  The 
suggestion  that  biannual  cohorts  coalesce  in  later  life  to  form  annual  peaks  is  confirmed  by  text-fig.  8,  and  the 
pattern  of  growth  amongst  adults  is  more  discernible  than  in  the  original  length-frequency  histogram.  For 
example,  a 6th  year-class  is  clearly  marked  in  text-fig.  8,  but  could  only  be  inferred  from  text-fig.  4 (i.e.  Table  3). 
However,  the  clearly  differentiated  7th  year-class  in  text-fig.  8 must  be  considered  as  artificial,  as  the  gap  in  the 
histogram  before  the  last  two  specimens  automatically  results  in  a strong  differentiation  from  the  preceding 
groups  of  specimens. 


A NEW  ZOSTEROPHYLL  FROM  THE  LOWER 
DEVONIAN  OF  POLAND 

by  DANUTA  ZDEBSKA 


Abstract.  A new  genus  and  species  Konioria  andrychoviensis  assigned  to  the  Zosterophyllophytina  is  described 
from  the  Lower  Devonian  (Emsian)  of  two  boreholes  in  the  Bielsko-Andrychow  area  of  the  Polish  Western 
Carpathians.  K.  andrychoviensis  possesses  dichotomous  axes  covered  on  their  lower  part  with  long  subulate 
spines  and  on  their  upper  part  with  short  triangular  spines.  Apices  of  axes  form  hooks.  In  addition  to  spines,  the 
axes  show  1-4  longitudinal  wings.  The  reniform  to  rounded  sporangia  are  borne  singly  at  dichotomies.  The 
structure  of  pyritized  axes  shows  a central  exarch  strand  with  scalariform  tracheids  and  a hypodermis.  In 
connection  with  the  unusual  position  of  the  sporangia  the  problem  of  the  evolution  of  the  lycopod  sporophyll 
is  discussed.  Konioria  appears  to  suggest  that  the  lycopod  sporophyll  originated  from  ends  of  fertile  axes, 
in  accordance  with  the  Telome  Theory  of  Zimmermann  (1930).  The  Telome  Theory,  however,  is  based  on 
the  Rhynia- type  of  organization,  while  other  evidence  suggests  that  the  lycopods  originated  from  the 
Zosterophyllophytina. 

This  paper  describes  a single  species  obtained  from  two  bore-holes,  Andrychow  2 and  Andrychow  4 
in  the  Bielsko-Andrychow  area  of  the  Polish  Western  Carpathians  (see  Turnau  1974).  The  depth 
in  Andrychow  2 is  between  2300-8  and  2306-6  m and  in  Andrychow  4 between  2245-5  and  2250-8  m. 
The  age  of  these  rocks  was  determined  as  Lower  Devonian  (Emsian)  on  the  basis  of  their  lithology 
by  Konior  (1965,  1966,  1968,  1969)  and  this  was  confirmed  by  the  miospores  (Konior  and  Turnau 
1973;  Turnau  1974). 

Fragments  of  Drepanophycus  spinaeformis  Goepp.  and  Dawsonites  sp.  have  been  recovered  from 
the  same  cores,  but  have  still  to  be  described.  Preliminary  investigations  of  the  plant  material  were 
carried  out  by  Maria  Reymanowna  who  gave  a short  description  of  Konioria  under  the  name  of 
Psilophyton  sp.  referred  to  in  Konior  (1965).  Well-preserved  fragments  of  the  plant  allow  conclusions 
to  be  drawn  on  its  systematic  position  and  on  the  probable  course  of  evolution  of  the  lycopod 
sporophyll. 


MATERIAL  AND  METHODS 

The  plant  axes  are  preserved  as  coaly  compressions  in  a dark-grey  siltstone.  These  remain  intact  when  removed 
from  the  matrix  with  hydrofluoric  acid.  When  macerated  in  Schulze’s  solution,  most  disintegrated  and  showed 
no  cells.  Several  axes  yielded  tracheid  fragments  with  circular  bordered  pits.  A few  cuticle  fragments  of  axes 
are  in  a state  of  natural  maceration.  These  are  translucent  enough  to  show  both  a central  dark  vascular  strand 
and  stomata.  Certain  other  axes  were  pyritized  and  were  not  very  compressed.  From  these,  thin  sections  were 
prepared  for  reflected  light  microscopy,  using  a modified  method  described  by  Edwards  (1968).  To  prevent 
the  crumbling  of  the  pyritized  axes  during  sectioning,  they  were  embedded  in  dental  plaster  of  paris,  and 
were  sectioned  using  a small  dental  saw.  These  sections  were  fixed  to  glass  slides  with  Canada  balsam  and 
ground  by  hand  in  the  usual  way. 

The  photographs  on  PI.  25,  figs,  5,  6;  PI.  26,  figs.  5,  7,  8;  PI.  27,  figs.  3,  5,  6,  7,  9,  and  PI.  28,  figs.  2,  3,  8,  9 
were  taken  with  the  Zeiss  Photomicroscope  III  Stand,  and  the  micrographs  on  PI.  25,  fig.  3;  PI.  27,  figs.  1,  2,  4; 
and  PI.  25,  fig.  4 with  the  Cambridge  S 600  SEM  in  the  Department  of  Botany,  Birkbeck  College,  University 
of  London. 

The  remaining  photographs  were  taken  with  an  Exacta  camera  and  a Zeiss-Jena  lightmicroscope  in  the 
Botanical  Institute  of  the  Jagiellonian  University,  Krakow. 

(Palaeontology,  Vol.  25,  Part  2,  1982,  pp.  247-263,  pis.  25-28.| 


248 


PALAEONTOLOGY,  VOLUME  25 


SYSTEMATIC  PALAEONTOLOGY 

Order  zosterophyllales 
Family  zosterophyllaceae 
Genus  konioria  gen.  nov. 

Konioria  andrychoviensis  sp.  nov. 

Plant  fragments  extracted  from  the  rock  samples  include  axes  with  both  long  and  short  spines,  axes  with  only 
short  or  long  spines,  sterile  and  fertile  apices,  and  pyritized  axes  (Table  1). 


table  1 . Characters  found  on  separate  parts  of  the  plant  which  indicate  that  they  belong  to  the  same  plant  (cross 
indicates  presence  of  character) 


__PLANT  PASTS 
CHARACTERS  — 

Axis  with 
long  and 
short  spines 

Axis  with 
long  spines 
only 

Axis  with 
short  spines 
only 

Sterile 

apex 

Fertile 

apex 

Pyritised 

axis 

Circinatelly 
coiled  axis 

long  spines 
with 

minute  teeth 

+ 

- 

- 

- 

Short  spines 

- 

- 

- 

* 

Wing 

+ 

- 

- 

+ 

- 

Stomata 

- 

- 

* 

- 

- 

- 

Axes  with  spines.  The  Konioria  axes  are  preserved  on  the  bedding  planes  in  a dark-grey  siltstone.  The  core  is  9 cm 
wide  which  limits  the  length  of  the  specimens;  originally  they  were  longer.  There  are  also  many  pieces  which  were 
broken  into  short  lengths  before  deposition. 

The  axes  dichotomize  into  unequal  or  almost  equal  parts  diverging  at  about  30°  (PI.  25,  figs.  1,  4).  The 
distance  between  successive  dichotomies  is  about  4 cm  towards  the  top  of  the  plant.  Below  dichotomies  the  axes 
widen  and  gradually  change  into  two  branches.  One  axis  was  2 mm  wide  at  1 cm  below  its  first  dichotomy  and  1-5 
mm  at  1 cm  below  its  second  dichotomy.  The  length  of  this  axis  is  over  6 cm.  Other  specimens  suggest  a possible 
length  of  about  15  cm  for  the  largest  axes  which  are  4 mm  wide.  The  length  of  the  plant  is  unknown  but 
some  estimate  is  possible  by  relating  the  extent  of  tapering  along  a single  axis  to  the  extremes  of  axis 
diameter  (0-3-4  0 mm). 

The  longest  unbranched  axis  available  is  7 cm  but  most  branch  at  closer  intervals  than  this  and  near  the  apex 
branching  is  very  close  together.  Successive  dichotomies  are  not  all  in  the  same  place  but  the  precise  angle 
between  them  in  the  lower  part  cannot  be  given.  Near  the  apex  it  is  about  90°.  A characteristic  and  remarkable 
feature  of  the  axes  is  that  they  have  one  to  four  longitudinal  wings.  These  may  lie  in  the  plane  of  compression  (PI. 
25,  fig.  4)  where  they  appear  as  borders  to  the  axis,  or  they  may  be  present  on  the  surface  of  the  axis  (PI.  25,  figs.  5, 


EXPLANATION  OF  PLATE  25 

Fig.  1.  Para  type,  axes  with  two  branches,  S/98/ 14;  x 1-2. 

Fig.  2.  Axes  with  long  spines,  S/98/12b;  x 3. 

Fig.  3.  SEM.  Stem  with  long  and  short  spines;  x 120. 

Fig.  4.  Axes  with  ‘compression  margin’  (wings),  S/98/9a;  x 3. 

Fig.  5.  Surface  of  axis  showing  three  wings  (arrowed)  (specimen  destroyed);  x 126. 

Fig.  6.  Compressed  axis  showing  wing  and  a long  spine  at  an  acute  angle  (specimen  destroyed);  x 6. 


PLATE  25 


ZDEBSKA,  Konioria 


250 


PALAEONTOLOGY,  VOLUME  25 


6;  PI.  26,  figs.  7, 8;  PI.  27,  fig.  6;  PI.  28,  figs.  2, 9).  These  wings  are  present  on  tall  axes.  Occasionally,  the  wings  take 
a spiral  course  (PI.  28,  fig.  9)  which  is  not  the  result  of  twisting  of  the  axis.  Twisted  axes  are  seen  on  PI.  25,  fig.  6. 
On  wings  as  well  as  on  the  axes,  the  spines  may  be  dense  (PI.  26,  fig.  6).  The  author  considered  the  possibility  that 
the  wings  were  mere  compression  borders,  but  decided  they  were  not,  on  the  evidence  given  below: 

1 . In  transverse  sections  of  pyritized  axes  the  wing  is  continuous  with  the  thick,  carbonized  layer  (pi.  26,  fig.  6). 

2.  Cells  of  the  hypodermal  layer  below  the  carbonized  layer  do  not  enter  the  wing  (PI.  26,  fig.  6). 

3.  In  a transverse  section  of  a flattened  pyritized  axis  the  wing  is  not  a continuation  of  the  longer  axis  of  the 
section.  If  the  wing  were  formed  as  a result  of  flattening  of  the  plant,  it  would  occur  in  the  same  plane 
(PI.  26,  fig.  6). 

4.  Often  the  wing  forms  a spiral  on  the  axis  (PI.  28,  fig.  9). 

Axes  bear  spines  of  varying  size  and  frequency.  The  spines  range  from  being  short  and  triangular  to  long  and 
subulate  (PI.  25,  fig.  3).  On  the  rock  the  spines  are  visible  at  the  sides  of  the  axes,  while  on  their  surface  only  their 
bases  are  seen  as  large  and  small  dots.  The  spines  are  distributed  irregularly;  rarely  are  they  very  frequent  on  one 
surface  of  the  axis  and  not  on  the  other  (PI.  26,  figs.  1,  2).  The  length  of  the  spines  is  very  unequal  on  different 
parts  of  the  axis  and  range  from  0T  to  4 mm.  The  longest  spines,  up  to  4 mm  long  and  0-3  mm  wide  at  the  base  are 
on  the  thickest  axes,  but  occasionally  there  may  be  fairly  long  spines  on  the  narrower  axes  (PI.  26,  fig.  3),  and 
occasionally  just  one  long  spine  is  present  (PI.  28,  fig.  3).  Usually  the  narrower  axes  bear  shorter  spines  (PI.  26,  fig. 
5;  text-fig.  1)  or  occasionally  none  at  all.  On  some  axes  (PI.  26,  fig.  2)  both  long  and  short  spines  occur.  Most 
spines  are  at  about  90°  to  the  axis.  The  longer  spines  show  longitudinal  ridges  (PI.  27,  fig.  2)  with  occasional 
minute  teeth  12-16  /xm  long  (PI.  27,  figs.  2, 4).  Teeth  are  not  obvious  on  the  shorter  spines  (PI.  25,  fig.  3).  The  long 
spines  with  teeth  are  most  numerous  on  some  circinately  coiled  axes  (PI.  28,  fig.  8)  which  the  author  thinks 
possibly  represent  early  stages  in  the  development  of  a shoot  (these  are  not  included  in  the  reconstruction).  Some 
spines  show  a dark  core  of  unknown  cellular  structure  (PI.  28,  fig.  3)  suggesting  a vascular  strand.  Such  structures 
are  spines  and  not  sporangial  stalks,  because  they  are  quite  long  and  situated  at  the  side  of  the  axis  and  not  below 
the  dichotomies. 

Cuticle  preparations  of  axis  showed  no  clearly  marked  epidermal  cells  (PI.  25,  fig.  3)  apart  from  dark  stomata. 
The  stomata  are  visible  as  irregularly  distributed  elliptical  dark  dots,  because  they  are  usually  covered  with  a 
dark  substance  (PI.  28,  figs.  1,  7).  They  are  oval  and  orientated  longitudinally,  and  sometimes  show  the  guard 
cells.  The  spines  have  no  stomata. 

Apex  of  axis.  All  apices  whether  sterile  or  fertile  are  much  branched  and  curved  to  form  hooks  (text-fig.  1 ; PI.  26, 
fig.  5;  PI.  27,  fig.  6;  PI.  28,  figs.  1,  2).  These  illustrations  show  all  the  variation  observed  in  both  sterile  and  fertile 
apices.  The  spines  on  the  ends  of  axes  are  never  long,  and  are  sometimes  infrequent  or  absent  (text-fig.  1).  The 
vascular  strand  may  be  visible  by  scanning  electron  microscopy  but  not  after  maceration.  The  dark  strand  in  PI. 
28,  figs.  1,  7,  shows  an  untreated  specimen  photographed  by  transmitted  light. 

Fertile  axes.  The  sporangia  cannot  be  distinguished  on  axes  still  in  the  rock,  probably  because  the  diameter  of  a 
sporangium  is  not  much  larger  than  the  width  of  the  axis  below  the  dichotomy.  They  are  visible  only  on  axes 
removed  with  hydrofluoric  acid.  Most  of  the  distal  branches  are  sterile,  but  in  some  instances  a single  sporangium 
occurs  at  the  final  or  penultimate  dichotomy,  and  perhaps  also  further  down.  The  sporangia  are  mostly  situated 
slightly  below  the  angle  of  the  dichotomy  and  any  stalk  they  possess  must  be  very  short  and  is  concealed  (PI.  26, 
figs.  7,  8;  PI.  28,  fig.  2;  text-figs.  2,  3).  It  appears  that  the  small  mound  visible  on  the  surface  of  most  sporangia  is 


EXPLANATION  OF  PLATE  26 

Figs.  1,  2.  Two  sides  of  a pyritized  axis,  S/98/23;  x 17-5.  1,  surface  with  numerous  short  spines.  2,  surface  with 
bases  of  two  long  spines  and  a few  short  spines. 

Fig.  3.  Fragment  of  axis  with  long  spines  and  wing,  S/98/26;  x 20. 

Fig.  4.  Pyritized  axis  just  below  dichotomy  (specimen  destroyed);  x 30. 

Fig.  5.  Apex  showing  short  spines,  S/98/26;  x 20. 

Fig.  6.  Transverse  section  of  pyritized  axes.  On  the  left-hand  side  is  a wing  with  spines,  S/98/79;  x 25. 

Figs.  7,  8.  Holotype,  fragment  of  axis  with  sporangium  situated  at  the  level  of  a dichotomy  (both  sides), 
S/98/31;  x 30. 


PLATE  26 


ZDEBSKA,  Konioria 


252 


PALAEONTOLOGY,  VOLUME  25 


EXPLANATION  OF  PLATE  27 

Fig.  1.  SEM.  Axis  with  two  flattened  branches  and  a stalked  sporangium  (arrows);  x 30. 

Fig.  2.  SEM.  Surface  of  spine  showing  longitudinal  ridges  with  minute  teeth;  x 300. 

Fig.  3.  Paratype,  transverse  section  of  uncompressed  axis  showing  elliptical  xylem  strand,  S/98/ 15;  x 50. 

Fig.  4.  SEM.  Long  spine  with  round  base  and  minute  teeth;  x 25. 

Fig.  5.  Cells  of  hypodermis  in  longitudinal  section,  S/98/18;  x 125. 

Fig.  6.  Forked  apex  showing  wing  on  under  side  (specimen  destroyed);  x 20. 

Fig.  7.  Protoxylem  and  metaxylem  in  longitudinal  section  showing  scalariform  tracheids,  S/98/17;  x 1 10. 

Fig.  8.  Pyritized  axis  split  longitudinally  and  showing  vascular  strand  with  circular  bordered  pits, 
S/98/49;  x 375. 

Fig.  9.  Metaxylem  and  protoxylem  from  fig.  3;  x 167. 


PLATE  27 


ZDEBSKA,  Konioria 


1 V 


254  PALAEONTOLOGY,  VOLUME  25 

caused  by  the  stalk  (PI.  26,  fig.  8).  Some,  however,  have  a longer  stalk  which  is  clearly  inserted  in  the  angle  of  the 
dichotomy  (PI.  27,  fig.  1). 

The  sporangia  are  flattened  and  composed  of  two  equal  valves  which  separate  along  their  entire  margin.  The 
largest  are  oval  or  slightly  reniform  (PI.  26,  figs.  7, 8;  text-fig.  2)  but  some  are  round  (PI.  28,  fig.  2;  text-fig.  2).  They 
are  typically  about  2-5  mm  wide,  but  the  round  ones  are  smaller  and  one  was  only  0-5  mm  wide.  The  surface  of 
the  outer  valve  may  show  diverging  cells  under  scanning  electron  microscopy.  Often  the  surface  of  a sporangium 
bears  minute  spines  (text-fig.  2).  Nothing  is  known  of  the  deeper  layers  of  the  sporangial  wall.  Along  the  line  of 
dehiscence  the  wall  is  flat  and  appears  thin  (PI.  28,  fig.  4).  Although  the  two  valves  are  always  pressed  together  the 
sporangia  seem  to  have  dehisced  and  shed  all  their  spores.  Maceration  yielded  no  spores  nor  did  it  yield  a central 
mass  which  might  represent  compacted  spores. 

Pyritized axes.  Ten  pyritized  axes  were  studied.  They  are  rather  narrow,  0-5-2  mm  wide.  Some  show  long  spines 
(PI.  26,  fig.  3),  some  short  spines  and  long  broken  spines  (PI.  26,  fig.  2),  and  some  all  short  spines  (PI.  26,  fig.  1). 
These  specimens  were  not  sectioned,  the  sections  being  prepared  from  unfigured  specimens  which  did  show  short 
spines  or  bases  of  long  spines. 


text-fig.  2.  Sporangia  attached  below  the  dichotomy.  Small  spines  present 
on  the  surfaces  of  sporangia. 


EXPLANATION  OF  PLATE  28 

Fig.  1.  Paratype,  naturally  cleared  apex  showing  dark  core  and  stomata  as  dark  spots,  S/98/29;  x 25. 

Fig.  2.  Apex  with  sporangium  (specimen  destroyed);  x 30. 

Fig.  3.  Fragment  of  axis  with  spines;  large  spine  apparently  with  ?vascular  strand  coming  from  the  axis, 

S/98/27;  x 20. 

Fig.  4.  SEM.  Fragment  of  sporangium  valve;  margin  appears  to  be  thin  and  smooth,  but  the  wall  shows 
?papillae;  x 300. 

Fig.  5,  6.  Tracheids  isolated  by  Schulze  maceration  from  carbonized  axis.  Tracheids  show  circular  bordered 
pits.  5,  S/98/75;  x 600;  6;  S/98/76;  x 546. 

Fig.  7.  Naturally  cleared  apices  showing  stomata  as  dark  spots  on  the  epidermis,  S/98/30;  x 24. 

Fig.  8.  Young  coiled  axis  densely  covered  with  long  spines,  S/98/25;  x 20. 

Fig.  9.  One  branch  of  the  dichotomy  (the  other  broken  off")  showing  spiral  wing  (specimen  destroyed);  x 19*5. 


PLATE  28 


ZDEBSKA,  Konioria 


256 


PALAEONTOLOGY,  VOLUME  25 


The  transverse  section  on  PI.  27,  fig.  3,  shows  a black,  coalified  outer  layer  which  can  be  regarded  as  the 
epidermis,  inside  this  are  about  four  layers  of  thick-walled  cells  which  represent  the  hypodermis.  Occasionally 
the  outer  coalified  layer  is  thicker,  suggesting  that  the  hypodermal  cells  are  also  coalified.  In  longitudinal  section 
the  hypodermal  cells  are  elongated  and  show  no  obvious  pits  (PI.  27,  fig.  5).  Length  of  cells  from  1 50  to  450  pm, 
width  from  30  to  85  pm. 


Inside  the  hypodermis,  there  is  a region  in  which  the  cells  are  scarcely  visible,  and  these  surround  a circular  (PI. 
26,  fig.  4)  or  elliptical  (PI.  27,  fig.  3)  xylem  strand,  with  small  tracheids  (regarded  as  protoxylem)  to  the  outside 
and  larger  tracheids  (regarded  as  metaxylem)  to  the  inside  (PI.  27.  fig.  9).  The  metaxylem  shows  well-developed 
scalariform  tracheids  (PI.  27,  fig.  7).  Three  different  stems  showed  exactly  similar  scalariform  tracheids  with  no 
round  pits.  However,  a further  specimen  with  the  surface  fractured  longitudinally  showed  round  pits  in  what 
appears  to  be  a metaxylem  tracheid  (PI.  27,  fig.  8),  but  when  this  stem  was  polished,  no  pits  were  seen.  The  author 
believes  that  the  round  pits  do  not  represent  pyrite  crystals,  which  usually  have  an  angular  outline.  Yet  another 
specimen  (showing  small  spines),  which  was  not  pyritized,  was  unusual  when  macerated  in  Schulze’s  solution,  in 
that  it  yielded  excellent  tracheids  in  longitudinal  view,  which  showed  round  bordered  pits  in  two  rows  in  some  of 
the  tracheids  (PI.  28,  figs.  5,  6).  The  variety  of  xylem  pitting  is  large,  but  such  evidence  does  suggest  that  all  the 
specimens  belong  to  one  species. 

Reconstruction  of  Konioria 

There  are  no  large  specimens  of  this  plant  available,  but  the  separate  fragments  are  linked  by 
common  characters  (Table  1).  The  reconstruction  (text-fig.  4)  is  made  from  the  drawings  of  all  those 
separate  parts. 

Konioria  gen.  nov. 

Type  species.  Konioria  andrychoviensis  sp.  nov. 

Diagnosis.  Erect  axes  slender,  branching  by  more  or  less  equal  dichotomy  in  different  planes.  Lower 
dichotomies  further  apart  than  upper  ones.  Apices  narrow  to  points,  and  are  curved  to  form  hooks. 
Axis  usually  with  1 -4  narrow  longitudinal  wings.  Surface  of  axes  with  spines,  which  range  from  long 
and  subulate  to  short  and  triangular.  Epidermis  of  axis  with  longitudinally  orientated  stomata.  Inside 
the  epidermis  is  a hypodermis  of  up  to  four  layers  of  thick-walled  cells.  The  hypodermal  cells  are 
rounded  in  transverse  section  but  elongated  longitudinally.  The  area  of  the  cortex  and  phloem,  is  not 


ZDEBSKA:  LOWER  DEVONIAN  ZOSTEROPHYLL  KONIORIA 


257 


preserved.  A central  xylem  strand  has  the  smallest  tracheids  to  the  outside.  Metaxylem  tracheids  have 
scalariform  thickening.  Sporangia  are  borne  singly  below  dichotomies,  usually  on  short  stalks.  The 
sporangia  are  round  or  reniform  in  outline,  and  composed  of  two  equal  valves,  with  the  dehiscence 
line  along  the  whole  margin. 


text-fig.  4.  Reconstruction  of  Konioria  andrychoviensis,  x 3 (approx.). 


Konioria  andrychoviensis  sp.  nov. 

Plates  25-28 

1965  Psilophyton  sp.  Reymanowna  in  Konior,  p.  217,  figs.  1,  3. 

Diagnosis.  Axes  0-3-4  mm  wide,  branching  frequently,  with  short  to  long  spines.  Lower  part  of  axis 
bearing  subulate  spines  up  to  4 mm  long  usually  with  minute  lateral  teeth,  the  upper  part  of  the  axis 
has  shorter  and  triangular  spines.  Sporangia  up  to  2-5  mm  wide  but  often  smaller,  frequently  bearing 
minute  spines  near  their  bases. 

Horizon.  Lower  Devonian,  Emsian. 

Locality.  Two  bore-holes  Andrychow  2 and  Andrychow  4 in  the  Bielsko-Andrychow  area  of  the  Western 
Carpathians,  Poland. 


258 


PALAEONTOLOGY,  VOLUME  25 


Type  specimens.  All  specimens  are  deposited  in  the  Palaeobotanical  Museum  of  Institute  of  Botany, 
Jagiellonian  University,  Krakow,  S/98.  Holotype;  S/98/31,  PI.  26,  figs.  7,  8.  Paratypes:  S/98/4,  PI.  25,  fig.  1;  S/98/ 
35,  text-fig.  4;  S/98/ 1 5,  PI.  27,  fig.  3;  S/98/29,  PI.  28,  fig.  1 . 

Derivation  of  name.  The  generic  name  Konioria  is  after  Professor  Konrad  Konior  who  found  the  material  and 
was  kind  enough  to  give  it  to  the  author  for  investigation.  The  specific  name  is  derived  from  the  name  of  the 
locality  Andrychow. 

DISCUSSION 

Comparison  with  Psilophyton  goldschmidtii  and  P.  arcticum 

Konioria  might  be  confused  with  P.  goldschmidtii  (Halle  1916)  and  P.  arcticum  (Hoeg  1942),  because 
of  similar  external  morphology  of  the  axes  as  seen  when  on  the  rock  surface.  A character  common  to 
both  Konioria  and  P.  goldschmidtii  is  the  ‘compression  margin’  (wing),  which  is  distinctly  visible  on 
specimens  seen  on  the  rock  surface.  Halle  (1916)  gives  no  explanation  of  its  nature.  In  Konioria  the 
‘margins’  are  in  fact  wings  running  along  both  sides  of  the  axis.  A second  similar  character  are  the 
subulate  spines  of  variable  length  up  to  4 mm.  The  spines  of  P.  goldschmidtii,  however,  are  less 
numerous.  The  two  plants  differ  in  their  branching.  In  Konioria  the  axes  dichotomize,  whilst  in 
P.  goldschmidtii  the  branching  is  sympodial  and  dichotomous.  The  axes  of  Konioria  are  straight  while 
in  P.  goldschmidtii  they  are  zigzag  shaped  (Nathorst  1913;  Halle  1916;  Hoeg  1967).  In  P.  goldschmidtii 
the  anatomical  structure  of  the  axes  and  the  sporangia  are  unknown.  Although  these  two  plants  show 
several  common  characters,  the  differences  between  them  and  the  different  mode  of  preservation 
suggest  that  the  new  plant  should  not  be  included  in  P.  goldschmidtii.  When  comparing  these  two 
plants,  the  author  was  able  to  study  the  figured  specimen  described  by  Halle  (1916).  The  similarities 
and  differences  between  the  two  plants  mentioned  above  were  confirmed  by  this  material. 

In  1959  Ananiev  described  P.  goldschmidtii  from  the  Devonian  of  south-eastern  Siberia.  Ananiev 
gives  no  description  of  the  plant,  but  his  photographs  (PI.  7,  figs.  1,2, 4;  PI.  8,  fig.  4;  PI.  12, fig.  1;P1. 14, 
figs.  1,  2;  PI.  24,  fig.  2d)  clearly  show  the  differences  between  the  branching  of  Konioria  and 
P.  goldschmidtii. 

In  1932  Lang  described  P.  princeps  and  P.  goldschmidtii  from  the  Strathmore  Beds  in  Scotland 
under  the  name  Psilophyton.  The  Psilophyton  figured  by  Lang  on  PI.  2,  figs.  24,  25,  27,  is  similar  to 
Konioria  in  that  it  shows  axes  with  a ‘compression  margin’  (wing).  His  PI.  2,  figs.  24,  33,  shows  the 
presence  of  short  spines  and  bases  of  broken  long  spines  similar  to  Konioria.  The  differences  in 
branching  between  the  two  plants  are  clearly  seen  when  comparing  Lang’s  PI.  2,  figs.  24,  25, 27,  with 
Konioria.  Lang  describes  the  branching  as  pseudomonopodial,  which  is  clearly  different  from  the 
dichotomous  branching  in  Konioria.  An  investigation  of  Lang’s  material  in  the  British  Museum  of 
Natural  History  in  London,  confirms  the  similarities  and  differences  between  the  two  taxa. 

Konioria  shows  some  external  similarity  to  P.  arcticum  Hoeg  1942,  which  also  possesses  a distinct 
margin  (see  his  PI.  9,  fig.  3;  PI.  12,  fig.  1).  Hoeg  does  not  explain  the  nature  of  this  margin.  Unlike 
Konioria,  however,  P.  arcticum  has  pseudomonopodial  branching  and  hair-like  spines  from  4 to 
6 mm  long. 

Comparison  with  genera  of  the  subdivision  Zoster ophyllophytina 

Attributing  Konioria  to  the  Zosterophyllophytina,  the  author  uses  the  classification  of  the 
‘psilophytes’  given  by  Banks  (1968,  1975).  Konioria  shows  the  characters  of  the  subdivision 
Zosterophyllophytina,  i.e.  the  lateral  arrangement  of  sporangia  and  the  exarch  xylem  strand. 

Comparison  with  Crenaticaulis  and  Gosslingia.  Konioria  possesses  a greater  number  of  characters 
in  common  with  Crenaticaulis  (Banks  and  Davis  1969)  and  Gosslingia  (Edwards  and  Banks  1965; 
Edwards  1970)  than  with  other  genera.  All  three  plants  show  dichotomous  branching,  although 
Crenaticaulis  and  Gosslingia  also  have  pseudomonopodial  branching.  Crenaticaulis  and  Gosslingia 
also  show  scars  below  dichotomies.  In  Crenaticaulis  axillary  branches  are  present  which  were 
compared  by  Banks  and  Davis  (1969)  with  the  rhizophores  of  Selaginella.  Also,  in  Konioria,  the 
sporangia  are  borne  in  the  same  position.  Another  character  in  common  in  these  three  genera  is  that 
the  sporangia  do  not  form  spikes. 


ZDEBSKA:  LOWER  DEVONIAN  ZOSTEROPHYLL  KONIORIA 


259 


Konioria  differs  from  Gosslingia  in  possessing  spines.  Crenaticaulis  shows  characteristic  short 
tooth-like  spines  arranged  in  one  or  two  rows,  while  Konioria  has  irregularly  arranged  spines  ranging 
from  long  and  subulate  to  short  and  triangular.  The  sporangia  of  Gosslingia  and  Konioria  split  into 
two  equal  valves  whilst  in  Crenaticaulis  there  is  a large  abaxial  and  a small  adaxial  valve. 

Comparison  with  Sawdonia  and  Euthursophyton.  Konioria  shows  a certain  similarity  with  Sawdonia 
ornata  (Dawson)  Hueber,  see  for  example,  Dawson  1871;  Hueber  1964,  1971;  Hueber  and  Banks 
1967;  Ananiev  and  Stepanov  1968.  The  two  plants  have  the  same  dichotomous  mode  of  branching 
and  the  lateral  arrangement  of  sporangia  which  split  into  two  equal  valves.  Sawdonia  differs  from 
Konioria  in  having  glandular  spines  and  sporangia  distributed  along  the  axis. 

Konioria  and  Euthursophyton  hamperbachense  Mustafa  (Mustafa  1978)  are  similar  in  their 
dichotomously  branching  axes  covered  with  long  spines,  in  occasionally  showing  circinately  coiled 
axes  and  in  possessing  an  exarch  strand.  In  Euthursophyton , however,  the  sporangia  are  unknown. 
Konioria,  unlike  Euthursophyton,  possesses  differentiated  spines,  from  long  and  subulate  with  minute 
teeth  to  short  and  triangular.  In  addition,  wings  on  the  surface  of  the  axes  are  present  in  Konioria  but 
absent  in  Euthursophyton.  In  the  Euthursophyton  axes  no  hypodermis  has  been  described. 

Comparison  with  Zosterophyllum  and  Rebuchia.  Konioria  differs  from  both  Zosterophyllum  (Croft 
and  Lang  1942;  Edwards  1969a,  b;  Lele  and  Walton  1961)  and  Rebuchia  (Dorf  1933;  Hueber  1970, 
1972a,  b ) in  not  having  the  sporangia  arranged  in  spikes.  In  addition,  these  two  genera  have  smooth 
axes,  which  in  Zosterophyllum  show  in  their  lower  parts  H -shaped  branching.  Common  to 
Zosterophyllum,  Rebuchia,  and  Konioria  are  lateral,  reniform  sporangia  splitting  into  two  equal 
valves.  Edwards  (1969)  demonstrated  exarch  xylem  in  Zosterophyllum  Hanover anum.  In  Rebuchia  the 
xylem  strand  is  not  well  known,  and  is  mentioned  only  by  Lepekhina,  Petrosian,  and  Radchenko 
(1962).  It  is  possible  that  this  is  also  a common  character  of  all  three  genera. 

The  systematic  position  of  Konioria 

The  author  accepts  the  classification  of  the  early  land  plants  into  Rhyniophytina,  Trimerophytina, 
and  Zosterophyllophytina  (see  Banks  1968,  1975).  The  lateral  position  of  sporangia  and  the  exarch 
xylem  strand  of  Konioria  suggest  its  affinity  with  genera  of  the  subdivision  Zosterophyllophytina. 
Comparisons  show  that  Konioria  differs  from  other  Zosterophyllophytina  in  having  sporangia 
attached  below  a dichotomy,  in  showing  wings  on  the  axis,  and  in  possessing  spines  ranging  from 
long  and  subulate  to  short  and  triangular.  These  are  the  characters  of  the  new  genus  and  species 
Konioria  andrychoviensis. 

In  his  classification.  Banks  (1968)  divided  the  order  Zosterophyllales  into  two  families, 
Zosterophyllaceae  and  Gosslingiaceae,  which  differ  in  the  grouping  of  sporangia.  In  a later  paper 
Banks  (1975)  distinguishes  only  one  family,  the  Zosterophyllaceae.  If  this  later  publication  were 
followed,  Konioria  would  become  a member  of  the  Zosterophyllaceae,  fitting  well  the  diagnosis  of 
this  family.  However,  it  would  appear  that  a return  to  the  former  division  of  the  Zosterophyllo- 
phytina into  two  families  would  be  better,  with  the  Zosterophyllaceae  having  sporangia  in  spikes,  and 
the  Gosslingiaceae  having  sporangia  scattered  along  the  axis.  These  two  families  were  also  recognized 
by  Kasper,  Andrews,  and  Forbes  (1974).  Perhaps  Konioria  would  even  merit  the  establishing  of  a 
third  family,  characterized  by  having  sporangia  attached  below  the  dichotomies. 

The  morphology  and  the  anatomical  structure  of  Konioria  provide  additional  data  to  support 
Banks’s  (1968)  classification  of  the  early  land  plants,  in  that  the  morphological  characters  discussed 
above  are  further  evidence  that  the  Zosterophyllophytina  forms  a natural  group.  This  group, 
according  to  the  hypothesis  of  Banks  (1968;  Chaloner  and  Sheerin  1979).  gave  rise  to  the 
Lycophytina. 

Konioria  and  the  evolution  of  the  lycopod  sporophyll 

The  unusual  position  of  the  sporangium  of  Konioria  appears  to  be  related  to  that  in  certain  early 
lycopods.  According  to  Zimmermann  (1930)  the  sporophylls  of  the  lycopods  developed  from 


260 


PALAEONTOLOGY,  VOLUME  25 


ultimate  branchlets  with  sporangia  of  the  Rhynia- type.  However,  according  to  Banks  (1968),  the 
Rhyniophytina  gave  rise  to  all  groups  of  plants,  except  the  lycopods  which  developed  from  the 
Zosterophyllophytina.  This  opinion  is  confirmed  by  the  anatomical  structure  and  the  type  and 
arrangement  of  sporangia  in  the  Zosterophyllophytina.  Accordingly,  the  intermediate  forms  leading 
to  the  sporophyll  of  lycopods  have  been  looked  for  among  the  Zosterophyllophytina.  In  accordance 
with  Zimmermann’s  Telome  Theory,  the  processes  of  Konioria  suggest  an  evolutionary  sequence  of 
the  lycopod  sporophyll  from  the  Zosterophyllophytina.  Text-fig.  5 is  an  attempt  to  arrange  such  a 


text-fig.  5.  Proposed  evolutionary  sequence  of  the  lycopod  sporophyll.  1,  Gosslingia\  2.  Konioria ; 
3,  Colpodexylon',  4.  Leclercqia',  5.  Cyclostigma. 


sequence  of  existing  fossil  plants,  which  could  lead  to  the  sporophylls  of  lycopods.  The  starting-point 
is  Gosslingia  (Edwards  1970),  which  bears  lateral  sporangia  scattered  along  the  axis.  The  next  stage  is 
Konioria  with  lateral  sporangia  borne  below  dichotomies.  Therefore,  Konioria  would  appear  to  be  the 
link  between  the  Zosterophyllophytina  and  such  lycopods  as  Colpodexylon  (Banks  1944)  and 
Leclercqia  (Banks,  Bonamo,  and  Grierson  1972)  which  have  a sporophyll  divided  into  three  or  more 
segments.  The  unequally  dichotomizing  axes  of  Konioria  can  be  regarded  as  the  initial  form  from 
which  the  sporophylls  of  those  lycopods  originated.  It  can  be  assumed  that  the  wider  branch  of  the 
unequal  dichotomy  of  Konioria  would  change  into  the  main  axis,  as  a result  of  the  process  of  over- 
topping. The  other  narrower  branch  of  Konioria  consisting  of  ends  of  axes  with  a sporangium 
attached  at  the  basis  of  the  dichotomy,  could  be  transformed  into  the  lycopod  sporophyll  by  the 
processes  of  planation  and  reduction.  In  Konioria  there  are  unequal  dichotomies  of  the  hook-like 
apices  under  which  the  sporangia  are  attached.  Therefore,  it  is  possible  to  derive  the  sporophylls  of 
Colpodexylon  directly  from  Konioria,  because  in  Konioria  there  are  apices  with  three  or  four  branches 
above  the  sporangium.  It  is  more  difficult  to  derive  the  sporophyll  of  the  lycopod  Leclercqia 
complexa,  which  ends  with  five  segments,  from  Konioria,  because  five  times  divided  apices  of 
Konioria  were  not  found.  In  theory,  however,  such  branching  is  also  possible. 


ZDEBSKA:  LOWER  DEVONIAN  ZOSTEROPHYLL  KONIORIA 


261 


There  is  also  the  theoretical  possibility  of  deriving  the  undivided  sporophylls  of  Cyclostigma 
(Chaloner  1968)  and  similar  lycopods  by  a reduction  from  sporophylls  showing  more  than  one 
division. 

As  a result  of  those  considerations,  the  explanation  by  Zimmermann  (1930)  of  the  origin  of  the 
lycopod  sporophyll  from  sterile  and  fertile  axes  by  overtopping  and  reduction  appears  justified,  and 
can  be  envisaged  directly  from  the  Zosterophyllophytina,  but  not  from  the  Rhyniophytina.  The 
sequence  proposed  in  the  present  paper  appears  to  confirm  the  axial  origin  of  the  lycopod  sporophyll. 

In  the  light  of  new  facts  established  about  the  position  of  sporangia  in  Drepanophycus  and 
Protolepidodendron,  these  two  genera  are  excluded  from  the  sequence.  Drepanophycus  ( Protolycopo - 
dites)  devonicus  is  not  placed  here,  because  Schweitzer  and  Giesen  (1980)  established  that  sporangia 
of  this  plant  do  not  occur  on  sporophylls,  but  on  the  axis  among  microphylls.  According  to 
Schweitzer  and  Giesen,  the  sporangia  in  D.  spinaeformis  Goepp.  also  have  a similar  position,  and  do 
not  occur  on  the  sporophyll  as  described  by  Krausel  and  Weyland  (1930).  Also,  Protolepidodendron 
scharianum  Potonie  and  Bernard  which  according  to  Krausel  and  Weyland  (1932)  possessed 
bifurcating  sporophylls  is  not  shown  in  the  sequence.  According  to  Schweitzer  and  Giesen  (1980),  its 
sporophylls  show  a double  dichotomy  with  a sporangium  present  below  each  dichotomy.  Schweitzer 
and  Giesen  think  that  P.  wahnbachense  Krausel  and  Weyland  also  shows  this  type  of  sporophyll  (see 
their  reconstruction  (text-fig.  12,  p.  15)).  However,  there  exists  a controversy  about  this  reconstruc- 
tion, because  according  to  Fairon-Demaret  (1979)  these  sporophylls  bear  several  sporangia,  and  she 
tentatively  attributes  this  plant  to  the  Sphenopsida  and  not  to  the  Lycopsida. 

In  summary,  the  above  considerations  demonstrate  the  possibility  that  the  lycopod  sporophyll 
may  have  originated  through  a change  in  the  ends  of  axes  into  sporophylls  divided  into  segments,  and 
to  a reduction  of  the  number  of  these  segments  (text-fig.  5). 

Acknowledgements.  I would  like  to  express  my  sincere  thanks  to  Professor  J.  Dyakowska  from  the  Botanical 
Institute  of  the  Jagiellonian  University  for  her  kind  care  during  the  preparation  of  my  doctor’s  thesis  and  to  the 
late  Professor  M.  Kostyniuk  for  his  comments.  To  Dr.  M.  Reymanowna  from  the  Botanical  Institute  of  the 
Polish  Academy  of  Sciences  my  thanks  for  consultation  and  critical  remarks.  My  gratitude  is  due  to  Professor 
K.  Konior  from  the  Geological  Institute  (Carpathian  Section)  in  Krakow  for  supplying  the  material  and  for  his 
stimulating  remarks.  I wish  to  thank  Professor  W.  G.  Chaloner  for  my  stay  during  1973  in  his  Department  in 
Birkbeck  College  of  the  University  of  London,  for  discussion,  and  the  opportunity  to  take  microphotographs 
and  SEM  micrographs.  I thank  Professor  T.  M.  Harris  for  his  remarks  on  the  paper  and  Professor  H.  P.  Banks 
and  Dr.  D.  Edwards  for  consultation.  To  Professor  B.  Lundblad  from  the  Swedish  Museum  of  Natural  History  I 
am  indebted  for  the  loan  of  specimens  of  Psilophyton  goldschmidtii.  I thank  the  British  Museum  (Natural 
History)  for  the  opportunity  to  study  the  W.  H.  Lang  collection. 


REFERENCES 

ananiev,  a.  r.  1959.  Die  wichtigsten  Fundstellen  von  Devonftora  im  Ssajan-Altaj Berggebiete,  Tomsk.  [In  Russian: 
German  summary.] 

— and  stepanov,  s.  a.  1968.  Nachodki  organov  sporonosenia  u Psilophyton princeps  Dawson  emend.  Halle  v 
niznem  devonie  Yuzno-Minusinskoi  Kotloviny  (Zapadnia  Sibir).  Trudy  tomsk.  gos.  Univ.  202,  31-46.  [In 
Russian.] 

banks,  h.  p.  1944.  A new  Devonian  lycopod  genus  from  southeastern  New  York.  Am.  J.  Bot.  31,  649-659. 

— 1968.  The  early  history  of  land  plants.  Pp.  73-107.  In  drake,  e.  t.  (ed.)  Evolution  and  environment , Yale 
University  Press,  New  Haven. 

1975.  Reclassification  of  psilophyta.  Taxon , 24,  401-413. 

— bonamo,  p.  M.  and  Grierson,  j.  d.  1972.  Leclercqia  complexa  gen.  et  sp.  nov.,  a new  lycopod  from  the  late 
Middle  Devonian  of  eastern  New  York.  Rev.  Palaeobot.  Palynol.  14,  19-40. 

— and  davis,  m.  r.  1969.  Crenaticaulis,  a new  genus  of  Devonian  plants  allied  to  Zosterophyllum,  and  its 
bearing  on  the  classification  of  early  land  plants.  Am.  J.  Bot.  56,  436-449. 

chaloner,  w.  g.  1968.  The  cone  of  Cyclostigma  kiltorkense  Haughton,  from  the  Upper  Devonian  of  Ireland.  J. 
Linn.  Soc.  61,  25-36. 


262 


PALAEONTOLOGY,  VOLUME  25 


chaloner,  w.  G.  and  sheerin,  a.  1979.  Devonian  macrofloras.  Special  papers  in  Palaeontology,  23,  145-161. 
croft,  w.  N.  and  lang,  w.  h.  1942.  The  Lower  Devonian  flora  of  the  Senni  Beds  of  Monmouthshire  and 
Breconshire.  Phil.  Trans.  R.  Soc.  231  B,  131-163. 

dawson,  j.  w.  1871.  The  fossil  plants  of  the  Devonian  and  Upper  Silurian  formations  of  Canada.  Geol.  Surv. 
Can.  1-92. 

dorf,  e.  1933.  A new  occurrence  of  the  oldest  known  terrestrial  vegetation,  from  Beartooth  Butte,  Wyoming. 
Bot.  Gaz.  95,  240-257. 

edwards,  d.  1968.  A new  plant  from  the  Lower  Old  Red  Sandstone  of  South  Wales.  Palaeontology,  11,  683- 
690. 

— 1969a.  Further  observations  on  Zoster ophy llum  llanoveranum  from  the  Lower  Devonian  of  South  Wales. 
Am.  J.  Bot.  56,  201-210. 

— 1969ft.  Zosterophyllum  from  the  Lower  Old  Red  Sandstone  of  South  Wales.  New  Phytol.  68,  923-931. 

— 1970.  Further  observations  on  the  Lower  Devonian  plant.  Gosslingia  breconensis  Heard.  Phil.  Trans.  R. 

Soc.  258  B,  225-243. 

— and  banks,  H.  p.  1965.  Branching  in  Gosslingia  breconensis.  Am.  J.  Bot.  52,  636. 
fairon-demaret,  m.  1979.  Estinnophyton  wahnbachense  (Krausel  and  Weyland)  comb,  nov.,  une  plante 

remarquable  du  Siegenien  d’Allemagne.  Rev.  Palaeobot.  Palynol.  28,  145-160. 
halle,  t.  G.  1916.  Lower  Devonian  plants  from  Roragen  in  Norway.  K.  svenska  Vetens.  Akad.  Handl.  57, 
1-46. 

hoeg,  o.  A.  1942.  The  Devonian  and  Downtonian  flora  of  Spitsbergen.  Norges  Svalb.  Ishars-Unders.  Skrifter, 
83,  1-228. 

— 1967.  Psilophyta.  Pp.  191-352.  In  boureau,  e.  (ed.),  Traite  de  Paleobotanique.  II.  Masson  et  Cie,  Paris. 
hueber,  F.  M.  1964.  The  Psilophytes  and  their  relationship  to  the  origin  of  ferns.  Bull.  Torrey  bot.  Club, 
21,  5-9. 

1970.  Rebuchia:  a new  name  for  Bucheria  Dorf.  Taxon,  19,  822. 

— 1971.  Sawdonia  ornata:  a new  name  for  Psilophyton  princeps  var.  ornatum.  Ibid.  20,  641-642. 

— 1972a.  Rebuchia  ovata,  it  vegetative  morphology  and  classification  with  the  Zosterophyllophytina.  Rev. 
Palaeobot.  Palynol.  14,  113-127. 

1972ft.  Early  Devonian  land  plants  from  Bathurst  Island,  District  of  Franklin.  Geol.  Surv.  Pap.  Can. 

71-28,  1-17. 

and  banks,  H.  p.  1967.  Psilophyton  princeps'.  the  search  for  organic  connection.  Taxon,  16,  81-85. 

kasper,  a.  e.,  Andrews,  h.  n.  and  forbes,  w.  h.  1974.  New  fertile  species  of  Psilophyton  from  the  Devonian  of 
Maine.  Am.  J.  Bot.  63,  339-359. 

konior,  k.  1965.  Le  Devonien  inferieur  dans  la  base  des  Karpates  bordurales  de  la  region  Cieszyn-Andrychow. 
Bull.  Acad.  Pol.  Sc.  ser.  geol.  geogr.  13,  215-219. 

— 1966.  Remarques  sur  le  developpement  et  Page  du  Devonien  inferieur  du  substratum  de  la  region 
Bielsko-Andrychow.  Ibid.  15,  231-232. 

— 1968.  Lower  Devonian  in  borehole  Andrychow  4.  Kwart.  geol.  12,  827-842.  [In  Polish:  English  summary.] 

— 1969.  The  Lower  Devonian  from  boreholes  in  the  Bielsko-Andrychow  region.  Acta  geol.  pol.  19, 177-220. 
[In  Polish:  English  summary.] 

— and  turnau,  e.  1973.  Preliminary  study  of  microflora  from  Lower  Devonian  deposits  in  the  area  of 
Bielsko-Wadowice.  Ann.  Soc.  geol.  Pol.  43,  273-282. 

krausel,  r.  and  weyland,  h.  1930.  Die  Flora  des  deutschen  Unterdevons.  Abh.  preuss.  geol.  Landesanst.  n.f. 
131, 1-92. 

— 1932.  Pflanzenreste  aus  dem  Devon.  IV.  Protolepidodendron  Krejci.  Senckenbergiana,  14,  391-403. 
lang,  w.  H.  1932.  Contributions  to  the  study  of  the  Old  Red  Sandstone  flora  of  Scotland.  VIII.  On  Arthrostigma. 
Psilophyton  and  some  associated  plant-remains  from  the  Strathmore  Beds  of  the  Caledonian  Lower  Old  Red 
Sandstone.  Trans.  R.  Soc.  Edinb.  57,  491-521. 

— and  cookson,  i.  c.  1935.  On  a flora,  including  vascular  land  plants,  associated  with  Monograptus , in  rocks 
of  Silurian  age,  from  Victoria,  Australia.  Phil.  Trans.  R.  Soc.  224  B,  421-449. 

lele,  k.  m.  and  walton,  j.  1961.  Contribution  to  the  knowledge  of  ‘ Zosterophyllum  myretonianum'  Penhallow 
from  the  Lower  Old  Red  Sandstone  of  Angus.  Trans.  R.  Soc.  Edinb.  64,  469-475. 
lepekhina,  v.  G.,  Petrosian,  n.  M.  and  Radchenko,  G.  p.  1962.  The  most  important  Devonian  plants  of  the  Altai- 
Sayan  mountain  region.  Trudy  vses.  nauchno-issled.  geol.  Inst.  70,  61-189.  [In  Russian.] 
mustafa,  h.  1978.  Beitrage  zur  Devonflora,  III.  Argumenta  palaeobot.  5,  91-132. 

nathorst,  a.  G.  1913.  Die  Pflanzenreste  der  Roragen  Ablagerung.  In  Goldschmidt,  v.  m.  (ed.),  Das  Devongebiet 
am  Roragen  bei  Roros,  Kristiania.  Pp.  25-27. 


ZDEBSKA:  LOWER  DEVONIAN  ZOSTEROPHYLL  KONIORIA 


263 


Schweitzer,  h.  j.  and  geesen,  p.  1980.  Uber  Taeniophyton  inopinatum,  Protolycopodites  devonicus  und 
Cladoxylon  scoparium  aus  dem  Mitteldevon  von  Wuppertal.  Palaeontographica,  173  B,  1-25. 
turnau,  E.  1974.  Microflora  from  core  samples  of  some  Paleozoic  sediments  from  beneath  the  flysch 
Carpathians  (Bielsko-Wadowice  area,  southern  Poland).  Ann.  Soc.  geol.  Pol.  64,  143-169. 
zimmermann,  w.  1930.  Die  Phylogenie  der  Pflanzen.  1.  Gustav  Fischer.  Jena. 


DANUTA  ZDEBSKA 


Typescript  received  21  June  1980 
Revised  typescript  received  29  April  1981 


Institute  of  Botany 
Jagiellonian  University 
ul.  Lubicz  46 
31-512  Krakow,  Poland 


TWO  S ALENIOID  ECHINOIDS 
IN  THE  DANIAN  OF  THE 
MAASTRICHT  AREA 

by  J.  F.  GEYS 


Abstract.  Two  species  of  salenioid  echinoids,  Salenia  minima  Agassiz  and  Desor,  1846  and  Hyposalenia 
heliophora  (Agassiz  and  Desor  1846)  from  the  Danian  strata  in  the  Maastricht  area,  are  described  and  discussed. 
Biometrical  parameters  are  statistically  treated  and  compared  with  those  of  some  other  salenioids.  It  is  shown 
that  both  species  are  probably  good  index  fossils  for  strata  of  Danian  age. 


Previous  systematic  treatments  of  the  Danian  echinoderm  fauna  in  the  Maastricht  area  were  made 
exclusively  by  investigators  studying  the  Cretaceous  Maastricht  Chalk.  Lambert  (1911)  was  the  first 
to  study  the  echinoid  fauna  of  the  Maastricht  Cretaceous  as  a whole.  The  fauna  was  systematically 
revised  by  Smiser  (1935).  Both  these  authors  essentially  studied  the  important  collections  of  the 
Koninklijk  Belgisch  Instituut  voor  Natuurwetenschappen  in  Brussels  (K.B.I.N.).  The  so-called  post- 
Maastrichtian  (or  Danian)  echinoids  from  the  Maastricht  area,  are  less  well  represented  in  these 
collections,  which  may  explain  why  the  characteristic  Danian  species  were  not  revised  by  both  these 
authors.  The  collections  of  the  K.B.I.N.,  as  well  as  those  of  the  Natuurhistorisch  Museum  at 
Maastricht  (N.H.M.M.),  were  revised  by  M.  Meijer.  Unfortunately,  his  results  were  only  partly 
published,  but  Meijer  (1965)  was  the  first  to  draw  attention  to  the  important  differences  between  the 
echinoid  faunas  of  the  Houthem  Formation  (post-Maastrichtian)  and  the  underlying  chalk  beds  (Ma 
to  Md),  belonging  to  the  Maastricht  Chalk.  As  a result  he  separated  his  ‘zone  III'  from  the  echinoid 
assemblages  of  the  Maastricht  Chalk,  and  he  presumed  that  this  zone  was  of  Dano-Montian  age. 
Meijer’s  view  proved  to  be  correct  when  Moorkens  (1972)  correlated  the  post-Maastrichtian 
Geulhem  Chalk  (Houthem  Formation)  with  the  Ciply  Tuffaceous  Chalk,  which  had  been  previously 
correlated  with  the  type  Danian  at  Fakse  (Denmark)  by  Rasmussen  (1965). 

Two  of  the  most  characteristic  species  from  the  post-Maastrichtian  Houthem  Formation  in  the 
Maastricht  area  are  discussed  in  this  paper.  They  occur  in  the  Ciply  Tuffaceous  Chalk,  in  the  vicinity 
of  Mons  (Belgium),  but  they  are  absent  in  overlying  and  in  underlying  strata.  In  the  Maastricht  area, 
both  species  seem  to  be  confined  to  strata  which  Felder  (1975)  termed  the  Geulhem  Chalk  of  the 
Houthem  Formation.  A correlation  between  the  Geulhem  Chalk  and  the  Ciply  Tuffaceous  Chalk  is 
thus  confirmed. 

The  specimens  studied  belong  to  the  collections  of  the  Natuurhistorisch  Museum  at  Maastricht  (N.H.M.).  In 
some  specimens,  a few  dimensions  were  measured  by  means  of  calipers  (absolute  error:  OT  mm).  Statistical 
calculations  were  carried  out,  using  the  formulas  and  symbols  proposed  by  Till  (1974). 

The  following  abbreviations  are  used: 

D:  ambital  diameter  of  the  test  III-5; 
h:  total  height  of  the  test; 

ds:  diameter  of  the  apical  system,  taken  between  the  centres  of  the  distal  borders  of  ocular  III  and  genital  5; 
dp:  diameter  of  the  peristome  III-5. 


IPalaeontology,  Vol.  25,  Part  2,  1982,  pp.  265-276,  pi.  29.| 


266 


PALAEONTOLOGY,  VOLUME  25 


SYSTEMATIC  DESCRIPTIONS 

Class  echinoidea  Leske,  1778 
Subclass  euechinoidea  Bronn,  1860 
Superorder  echinacea  Claus,  1876 
Order  salenioida  Delage  and  Herouard,  1903 
Family  saleniidae  Agassiz,  1838 
Subfamily  saleniinae  Agassiz,  1838 
Genus  salenia  Gray,  1835 

(=  Cidarelle  Desmoulins,  1835;  = Bathy salenia  Pomel,  1838) 

Type  species.  Cidarites  scutigera  Munster  in  Goldfuss  1826,  by  original  designation. 

Salenia  minima  Agassiz  and  Desor,  1 846 
Plate  29,  figs.  1-4 

*.1846  Salenia  minima,  Agassiz  and  Desor,  p.  342. 

.1850  Salenia  minima,  d’Orbigny,  p.  273. 

.1857  Salenia  minima,  Desor,  p.  151. 

1857  Salenia  minima.  Bosquet,  no.  839. 

1859  Salenia  minima,  Binkhorst  van  den  Binkhorst,  p.  120. 

.1864  Salenia  minima,  Cotteau,  pp.  171-173,  pi.  1040,  figs.  1-10. 

1879  Salenia  minima,  Ubaghs,  p.  228. 

1881  Salenia  minima,  Mourlon,  p.  125. 

.1910  Salenia  minima,  Lambert  and  Thiery,  p.  21 1. 

.1935  Salenia  minima,  Mortensen,  p.  369,  fig.  195 d. 
v.1965  Salenia  minima,  Meijer,  p.  23. 

.1979  Salenia  minima,  Geys,  p.  320. 
non  1935  Salenia  minima,  Smiser,  p.  28,  pi.  2,  fig.  6 a-d. 

Type  locality.  Ciply,  Hainaut,  Belgium.  Ciply  Tuffaceous  Chalk,  Danian. 

Studied  specimens  from  the  Maastricht  area  {Geulhem  Chalk).  Vroenhoven,  Belgian  Limburg:  88  specimens; 
Geulhem,  Dutch  Limburg:  18  specimens. 

Dimensions.  D:  1-7-9-2  mm;  h:  0-9-6  0 mm;  h/D  ratio:  0-48-0-70;  ds:  1-7-6-8  mm;  ds/D  ratio:  0-60-1-00;  dp: 
0-8-4-8  mm;  dp/D  ratio:  0-36-0-63. 

Description.  The  adoral  side  of  the  test  is  flat;  the  peristome  is  circular  to  subpentagonal,  not  sunken.  Gill  slits  are 
very  small.  The  perignatic  girdle  consists  of  small,  spatulate  auricles,  not  in  contact  with  each  other,  and  low 
arcuate  apophyses. 

The  apical  system  is  large  and  covers  most  of  the  aboral  side  of  the  test  in  young  specimens.  It  is  slightly  convex 
and  consists  of  eleven  smooth  plates,  separated  by  distinct  sutures.  Sutural  depressions  are  very  small  or  absent. 
The  ocular  plates  are  triangular  and  a little  concave.  The  genital  plates  and  the  suranal  plate  are  nearly  flat.  The 
genital  pores  have  an  excentric  position  in  the  genital  plates,  being  nearer  to  the  apex.  The  madreporite  can  easily 
be  distinguished  by  its  irregular,  more  or  less  triangular  poriferous  depression.  The  periproct  is  oval  or 
subtrigonal. 

Each  ambulacral  series  shows  ten  to  twelve  crenulate,  non-perforate  primary  tubercles.  The  ambulacra  are 
straight.  The  bigeminate  character  of  its  plates  is  very  regular.  The  tubercles  are  close  together  and 
extrascrobicular  granulation  is  almost  completely  absent.  The  axes  of  the  pore  pairs  have  an  inclination  of 
about  45°. 

Interambulacral  tubercles  are  crenulate,  non-perforate.  Five  of  them  make  a series.  The  areoles  are  smooth, 
large,  and  close  together.  Sometimes  they  are  confluent,  in  the  vicinity  of  the  ambitus.  A ring  of  six  to  eight 
scrobicular  tubercles  surrounds  them  on  all  but  the  adradial  sides.  Interradial  extrascrobicular  surfaces  are 
coarsely  granulated. 

Variability,  dp,  ds,  and  h were  plotted  against  D.  Some  parameters  and  the  reduced  major  axis  lines  (rmal)  were 
computed  for  each  of  these  graphs,  using  the  formulas  proposed  by  Till  (1974). 


GEYS:  SALENIOID  ECHINOIDS 


267 


dp  ( 
5- 
4- 
3- 
2 
1 


mm) 


Salenia  minima 


r = °-97  . 
slope  = 25°30' 

rmal  : dp=  0,06  + 0,48  D 


123456789  D(mm) 
text-fig.  1.  dp-D  plot  of  Salenia  minima,  with  reduced  major  axis  line  (rmal). 


a.  The  rmal  of  the  dp-D  plot  is  given  by 

dp  = 0-06  + 048  D (1)  (Text-fig.  1). 

The  95%  confidence  intervals  of  intercept  and  slope  are  respectively 
1 -96  sa(D/dp)  = +0-16  mm  (2) 

1 96  sb(D/dp)=  ±0-04  (3). 

From  (1)  and  (2)  one  can  conclude  that  the  origin  of  the  plot  is  included  in  the  confidence  band  of  the  rmal,  and 
that  the  dp/D  ratio  can  be  considered  constant.  Moreover,  the  slope  (b  = 0-48  ± 0-04)  does  not  differ  significantly 
from  the  mean  dp/ratio  (=  0-50),  as  follows  from  (1)  and  (3).  The  relative  size  of  the  peristome  is  not  influenced 
by  over-all  size  or  age  of  the  specimen. 

b.  The  rmal  of  the  h-D  plot  is  given  by 

h = - 0-3 1 + 0-68  D (4)  (Text-fig.  2) 

The  95%  confidence  intervals  of  intercept  and  slope  are  respectively 
1-96  sa(D/h)  = +0-13  mm  (5) 

1-96  sb(D/h)=  ±0-03  (6). 

Obviously  the  origin  is  not  included  in  the  confidence  band  of  the  rmal,  as  follows  from  (4)  and  (5).  (6)  indicates 
that  the  slope  (b  = 0-68  ± 0-03)  differs  significantly  from  the  mean  h/D  ratio  ( = 0-59).  One  can  thus  conclude  that 
the  h/D  ratio  is  not  constant,  though  the  relationship  between  h and  D is  linear.  Larger  specimens  tend  to  be  less 
flattened  than  smaller  individuals. 

c.  The  rmal  of  the  ds-D  plot  is  given  by 

ds  = 0-62  + 0-64  D (7)  (Text-fig.  3), 
with  95%  confidence  intervals  of  intercept  and  slope  respectively 
1 -96  sa(D/ds)  = + 0T  8 mm  (8) 

1-96  sb(D/ds)=  ±0-02  (9). 

On  one  hand,  one  can  conclude  from  (7)  and  (8)  that  the  origin  is  not  included  in  the  confidence  band  of  the  rmal; 
on  the  other,  (7)  and  (9)  indicate  that  the  slope  (b  = 0-64  + 0-02)  differs  significantly  from  the  mean  ds/D  ratio 
(=  0-82).  Larger  individuals  are  shown  to  have  a relatively  smaller  apical  system  than  smaller  individuals.  This 
could  be  put  as  follows:  the  apical  system  seems  to  grow  less  fast  than  the  entire  animal. 

The  relationship  is  clearly  linear  for  specimens  with  D > 4-5  mm  and  for  specimens  with  D < 4-5  mm 


268 


PALAEONTOLOGY,  VOLUME  25 


h (mm) 


text-fig.  3.  ds-D  plot  of  Salenia  minima , with  reduced  major  axis  line  (rmal). 


GEYS:  SALENIOID  ECHINOIDS 


269 


separately,  but  not  for  the  population  as  a whole.  This  is  clearly  demonstrated  by  computing  the  rmal  for  both 
partial  populations  separately. 

For  large  specimens  (D>4-5  mm)  one  obtains 
ds=  -1-18  + 0-87  D 
with  95%  confidence  intervals 
1-96  sa(D/ds)  = ±0-61  mm 
1 -96  sb(D/ds)  = ±0-04. 

For  small  specimens  (D<4-5  mm)  one  obtains  analogously 
ds  = 0-39  + 0-73  D 
with  95%  confidence  intervals 
1 96  sa(D/ds)  = ±0-23  mm 
1 96  sb(D/ds)=  ±0-07. 

From  these  equations  we  can  conclude  that  the  intercept  is  significantly  different  in  large  and  small  individuals. 
The  difference  in  slope  between  the  rmals  of  both  populations  is  much  less  important.  It  seems  as  if  the  rate  of 
growth  of  the  apical  system  is  drastically  reduced  when  the  specimens  attain  4 to  5 mm  over-all  diameter.  After 
crossing  the  D = 5 mm  threshold,  the  rate  of  growth  of  the  apical  system  seems  to  be  more  or  less  restored. 

Discussion.  Salenia  minima  is  superficially  similar  to  other  small  salenioids,  such  as  Salenidia pygmaea 
(Hagenow  1940)  and  Salenidia  maestrichtensis  (Schliiter  1892).  The  lack  of  sutural  depressions  in  its 
smooth  apical  system  makes  Salenia  minima  easy  to  recognize. 

Smiser  (1935)  confused  Salenia  minima  and  Goniopygus  minor.  What  he  figured  as  Salenia  minima 
is  in  fact  a Goniopygus  minor  Sorignet,  1850.  This  error  is  discussed  in  a forthcoming  paper  of  mine 
(Geys  1981). 

A very  similar  species  to  Salenia  minima  has  been  described  from  the  Danian  of  Denmark: 
Salenidia  danica  Ravn,  1928.  Both  species  probably  differ  in  the  structure  of  their  ambulacra  and  in 
the  shape  of  their  ocular  plates. 

For  both  Salenidia  pygmaea  and  Salenidia  maestrichtensis  h-D,  ds-D,  and  dp-D  plots  were 
statistically  analysed,  respectively  by  Nestler  (1965)  and  by  Geys  (1979).  A comparison  with  Salenia 
minima  is  instructive.  The  flattening  in  shape  of  larger,  hence  older  specimens,  as  demonstrated  in 
Salenia  minima,  exists  in  Salenidia  pygmaea  but  not  in  Salenidia  maestrichtensis.  The  relative 
shrinking  of  the  peristome,  demonstrated  in  Salenidia  pygmaea  and  Salenidia  maestrichtensis,  is  not 
indicated  in  Salenia  minima.  The  growth  of  the  apical  system,  however,  shows  similar  patterns  in  all 
three  of  the  species.  Very  small,  very  young  specimens  have  apical  systems  covering  the  complete 
adapical  surface.  In  Salenidia  maestrichtensis  a non-linear  relationship  between  D and  ds  was 
demonstrated.  For  Salenidia  maestrichtensis  it  was  established  that  the  beginning  of  the  final  linear 
segments  in  its  d-D  plot  (D  >4-5  mm)  coincides  with  the  attainment  of  sexual  maturity.  It  is  not  clear 
whether  the  same  is  true  for  Salenia  minima. 

Subfamily  hyposaleniinae  Mortensen,  1934 
Genus  hyposalenia  Desor,  1856 

( = Peltastes  Agassiz,  1838,  non  Peltastes  Rossi,  1807;  Peltosalenia  Quenstedt,  1874) 

Type  species.  Echinus  acanthoides  Desmoulins  1837,  by  subsequent  designation  of  Mortensen  (1935). 

Hyposalenia  heliophora  (Agassiz  and  Desor,  1 846) 

Plate  29,  figs.  5-8 

v*.1846  Salenia  heliophora,  Agassiz  and  Desor,  p.  342. 
v.1850  Salenia  heliophora,  d’Orbigny,  p.  273. 
v.1856  Hyposalenia  heliophora,  Desor,  p.  148. 

1857  Hyposalenia  heliophora.  Bosquet,  no.  842. 
v.  1 864  Peltastes  heliophorus,  Cotteau,  pp.  122-124,  pi.  1029,  figs.  1-7. 


270 


PALAEONTOLOGY,  VOLUME  25 


1874  Peltastes  heliophorus,  Cotteau,  p.  642. 

1879  Hyposalenia  heliophora,  Ubaghs,  p.  228. 

1881  Hyposalenia  heliophora,  Mourlon,  p.  125. 

.1892  Peltastes  cf.  heliophorus,  Schliiter,  pp.  152-154. 

.1910  Peltastes  heliophorus,  Lambert  and  Thiery,  p.  209. 
v.1928  Hyposalenia  heliophora,  Lambert  and  Jeannet,  p.  203. 

.1935  Peltastes  cf.  heliophorus,  Kongiel,  p.  31,  pi.  2,  fig.  5 a-c. 

.1935  Hyposalenia  heliophora,  Mortensen,  p.  344,  fig.  188g. 

.1939  Hyposalenia  heliophora,  Kongiel,  p.  20,  pi.  3,  figs.  19-21. 

.1966  Hyposalenia  heliophora.  Fell  and  Pawson,  p.  U379,  fig.  277-1/ 
v.1966  Hyposalenia  heliophora,  Meijer,  p.  23. 

.1979  Hyposalenia  heliophora,  Geys,  p.  320. 

Type  Locality.  Maastricht,  Dutch  Limburg,  the  Netherlands.  Geulhem  Chalk,  Houthem  Formation,  Danian. 

Other  occurrences.  Ciply,  Hainaut,  Belgium:  Ciply  Tuffaceous  Chalk,  Danian  (Agassiz  and  Desor  1846).  Ciply, 
Hainaut,  Belgium:  Malogne  Gravel,  Danian  (Cotteau  1874).  Berlin,  Germany:  ‘Geschiebe’,  reworked  in 
Pleistocene  (Schliiter  1892).  Gora  Pulawska,  Poland:  Siwak,  Lower  Danian  (Kongiel  1935,  1939). 

Studied  specimens  from  the  Maastricht  area  ( Geulhem  Chalk).  Vroenhoven,  Belgian  Limburg:  41  specimens; 
Geulhem,  Dutch  Limburg:  225  specimens. 

Dimensions.  Forty  specimens,  chosen  at  random,  were  measured.  D:  1-7-12-0  mm;  h:  1 -0-7-5  mm;  h/D  ratio: 
0-5-0-7;  ds:  1 -6-7-7  mm;  ds/D  ratio:  0-62-1 -00;  dp:  0-6-5-2  mm;  dp/D  ratio:  0-26-0-58. 

Description.  The  adoral  side  is  slightly  convex,  the  peristome  a little  sunken.  The  peristome  is  distinctly 
decagonal  and  shows  rather  large  gill  slits.  These  slits  are  surrounded  by  low,  blunt  ridges.  The  auricles  are  short, 
rectangular,  and  spatulate.  They  are  not  in  contact  over  the  perradial  suture. 

The  apical  system  is  large  and  pentagonal,  covering  most  of  the  adapical  surface.  It  is  conically  convex  and 
consists  of  eleven  plates,  separated  by  very  fine,  hardly  visible  sutures.  Real  sutural  depressions  are  absent.  The 
ocular  plates  are  pentagonal,  with  more  or  less  straight  distal  margins.  The  genital  plates  are  hexagonal,  the 
suranal  plate  is  pentagonal.  All  the  plates  are  covered  by  a conspicuous  sculpture  of  ridges  and  furrows,  which 
radiate  from  the  centres  of  the  plates,  and  which  are  perpendicular  to  the  sutures.  The  genital  pores  have  a central 
position  in  the  plates.  The  periproct  is  oval  or  subtrigonal.  It  is  situated  between  the  apex  and  genital  plate  5,  so 
that  the  test  is  symmetrical  with  respect  to  the  III-5  plane. 

Ambulacral  series  consist  of  twelve  or  thirteen  crenulate,  non-perforate  primary  tubercles.  The  tubercles  just 
below  the  ambitus  are  the  largest  in  size.  From  there  upwards,  their  size  diminishes  abruptly.  The  ambulacra  are 
straight.  The  plates  are  very  regularly  bigeminate.  The  poriferous  zones  are  uniserial  throughout.  The  pores  of 
each  pair  are  separated  by  a low  ridge.  The  axes  of  the  pore  pairs  have  an  inclination  of  45°  below,  and  less  than 
45°  above,  the  ambitus.  Very  small  scrobicular  tubercles,  or  granules,  surround  the  primaries. 

The  interambulacral  tubercles  are  crenulate,  non-perforate.  There  are  four  or  five  in  each  series.  The  areoles 
are  smooth,  large,  and  confluent.  Scrobicular  rings  are  hence  discontinuous.  Granulation  is  coarse  on  the 
interradial  miliary  surfaces. 

Variability,  dp,  ds,  and  h were  plotted  against  D.  Reduced  major  axis  lines  (rmal)  and  some  parameters  were 
computed  for  each  of  these  graphs. 


EXPLANATION  OF  PLATE  29 

Figs.  1 -4.  Salenia  minima  Agassiz  & Desor;  NHMM-MM  895;  Vroenhoven,  Belgian  Limburg;  Geulhem  Chalk. 

1 , adapical  view,  x 4.  2,  adoral  view,  x 4.  3,  lateral  view,  x 4.  4,  detail  of  ambulacrum  at  the  ambitus,  x 12. 
Figs.  5-8.  Hyposalenia  heliophora  (Agassiz  & Desor);  NHMM-MM  899;  Geulhem,  Dutch  Limburg;  Geulhem 
Chalk.  5,  detail  of  ambulacrum  at  the  ambitus,  x 12.  6,  adapical  view,  x4.  7,  adoral  view,  x4.  8,  lateral 
view,  x 4. 


PLATE  29 


GEYS,  salenioid  echinoids 


272  PALAEONTOLOGY,  VOLUME  25 


a.  The  rmal  of  the  dp-D  plot  is  given  by 

dp  = — 0 06  + 0-38  D (text-fig.  4) 
with  95%  confidence  intervals  of  intercept  and  slope,  respectively 
1 -96  sa(dp/D)  = + 0-1 9 mm 
1 -96  sb(dp/D)  = ±0-04. 

From  these  equations  can  be  concluded  that  the  origin  of  the  plot  is  included  in  the  confidence  band  of  the 
rmal,  and  that  the  mean  dp/D  ratio  ( = 0-37)  does  not  differ  significantly  from  the  slope  (b  = 0-38  + 0 04).  Hence, 
there  is  no  difference  in  relative  size  of  the  peristome,  between  young  (small)  and  old  (large)  specimens.  The  dp/D 
ratio  is  constant. 

b.  The  rmal  of  the  h-D  plot  is  given  by 

h = 0-16  + 0-57  D (text-fig.  5) 

with  95%  confidence  intervals 
1 -96  sa(h/D)  = ±0-11  mm 
1 96  sb(h/D)  = 0-01. 

The  origin  of  the  plot  is  clearly  not  included  in  the  confidence  band  of  the  rmal.  Moreover,  the  mean  h/D 
ratio  ( = 0-62)  differs  significantly  from  the  slope  (b  = 0-57±0-01).  Though  the  relationship  between  h and 
D is  linear,  the  h/D  ratio  is  not  constant.  Older  (larger)  specimens  are  slightly  more  flattened  than  young 
(small)  ones. 

c.  The  rmal  of  the  ds-D  plot  is  given  by 

ds  = 0-60  + 0-62  D (Text-fig.  6) 
with  95%  confidence  intervals 
1 -96  sa(ds/D)  = ±0-12  mm 
1 -96  sb(ds/D)  = +0-01. 

Obviously  the  origin  of  the  graph  is  not  included  in  the  confidence  band  of  the  rmal.  The  mean  ds/D  ratio 
(=  0-78)  differs  significantly  from  the  slope  (b  = 0-62  + 0-01).  The  ds/D  ratio  is  thus  not  constant.  The  apical 
system,  covering  the  entire  adapical  surface  in  very  young  individuals,  grows  relatively  smaller  in  older 
specimens.  One  could  say  that  the  apical  system  has  a rate  of  growth  less  fast  than  that  of  the  animal  as  a whole. 
Unlike  the  other  salenioids,  discussed  above,  the  relationship  between  ds  and  D is  linear  in  Hyposalenia 
heliophora. 


h(mm) 


Hyposalenia  heliophora 


rmal  : h = 0,16  ♦ 0,57  D 
slope  = 30°30' 
r = 0,99 


1 23456789  10  11  D(mm) 

text-fig.  5.  h-D  plot  of  Hyposalenia  heliophora  with  reduced  major  axis  line  (rmal). 


ds(mm) 


Hyposalenia  heliophora 


rmal  : ds  = 0,60  + 0,62  D 
slope  = 32° 
r = 0,99 


D(mm) 


text-fig.  6.  ds-D  plot  of  Hyposalenia  heliophora  with  reduced  major  axis  line  (rmal). 


274 


PALAEONTOLOGY,  VOLUME  25 


Discussion.  It  is  clear  from  the  synonymy  list  that  little  confusion  has  arisen  with  repect  to  H. 
heliophora.  Cotteau  (1864)  describes  this  species  as  being  easily  recognized,  owing  to  the  character- 
istic structure  of  its  ambulacra  and  of  its  apical  system.  Little  can  be  added  to  that  statement.  Like 
Salenia  minima  Agassiz  and  Desor,  1 846,  and  Salenidia pygmaea  (Hagenow  1 840),  H.  heliophora  shows 
a slight  tendency  to  flatten  in  shape  with  age.  Like  Salenia  minima,  but  unlike  Salenidia  pygmaea  and 
Salenidia  maestrichtensis  (Schliiter  1892),  H.  heliophora  has  a constant  dp/D  ratio:  the  relative  size  of 
its  peristome  does  not  change  with  over-all  size  or  age.  The  growth  pattern  of  the  apical  system  in  H. 
heliophora  shows  some  characteristics,  common  to  most,  if  not  all,  Salenioids:  a strong  reduction  in 
relative  size  with  age  or  with  over-all  size.  Unlike  the  other  species  which  were  statistically  analysed, 
H.  heliophora  has  a constant  rate  of  growth  in  its  apical  system. 


text-fig.  7.  Detail  of  the  apical  system,  with  periproct,  suranal  plate  and  surrounding 
genital  plates  in  Salenia  minima  (a)  and  Hyposalenia  heliophora  ( b ).  Enlarged,  x 8. 


CONCLUSIONS 

On  various  occasions  Salenia  minima  and  H.  heliophora  have  been  reported  from  the  Upper 
Cretaceous  in  Belgium  and  in  the  Netherlands  (Bosquet  1857;  Binkhorst  1859;  Ubaghs  1879; 
Mourlon  1881).  These  reports  are  not  really  erroneous,  because  the  Houthem  Formation,  as  well  as 
the  Ciply  Tuffaceous  Chalk,  were  included  within  the  Maastrichtian  or  ‘Senonian’  at  that  time.  The 
strati  graphical  ranges  of  both  species  were  precisely  documented  by  Meijer  (1966)  in  Dutch  Limburg: 
the  ‘uppermost  part  of  the  Maastricht  Chalk’  (now  called  Houthem  Formation),  which  he  called 
‘zone  III’  and  was  then  considered  to  be  of  Dano-Montian  age.  Among  Cretaceous  salenioids 
in  various  Dutch  and  Belgian  collections  (Geys  1979),  I could  not  trace  any  specimen  of  either  of  the 
species  described  here,  which  was  of  true  Cretaceous  age. 

Specimens  of  Salenia  minima  and  of  H.  heliophora,  in  the  Natuurhistorisch  Museum  at  Maastricht, 
were  collected  exclusively  in  the  so-called  ‘post-Maastrichtian’.  These  strata,  included  in  zone  III  of 
Meijer  (1966),  and  called  Geulhem  Chalk  of  the  Houthem  Formation  by  Felder  (1975),  were 
considered  of  Danian  age  by  Meijer  (1959).  In  the  Mons  Basin  the  same  two  species  seem  to  be 
restricted  to  the  Ciply  Tuffaceous  Chalk  and  the  underlying  Malogne  Gravel.  The  Ciply  Tuffaceous 
Chalk  was  shown  to  be  of  Danian  age  by  Rasmussen  (1965). 

The  debate  whether  the  Mesozoic-Cenozoic  boundary  is  situated  at  the  base  or  at  the  top  of  the 
Danian  is  not  yet  completely  resolved.  Some  authors  still  include  the  Danian  with  the  Upper 
Cretaceous  (e.g.  Davies  1975).  By  far  the  most  widespread  point  of  view  is  to  include  the  Danian  into 
the  Lower  Tertiary  (Berggren  1963;  Moorkens  1972).  The  evidence  is  mainly  based  on  various 


GEYS:  SALENIOID  ECHINOIDS 


275 


palaeontological  arguments:  foraminifera  and  calcareous  nannoplankton  (Cepek  and  Moorkens 
1979),  ostracods  (Deroo  1966),  brachiopods  (Kruytzer  and  Meijer  1958).  Both  Salenia  minima  and 
H.  heliophora  seem  to  be  index  fossils  for  Danian  strata.  Their  presence  in  Maastrichtian  or  in 
Montian  deposits  has  not  yet  been  demonstrated. 

Acknowlegements.  I am  indebted  to  Dr.  D.  G.  Montagne  (Natuurhistorisch  Museum,  Maastricht,  the 
Netherlands),  to  Drs.  X.  Misonne  and  P.  Sartenaer  (Koninklijk  Belgisch  Instituut  voor  Natuurweten- 
schappen,  Brussels,  Belgium),  to  Mrs.  D.  Gaspard  (Universite  de  Paris-Sud,  Orsay,  France)  and  to  Professor 
Dr.  J.  Remane  (Universite  de  Neuchatel,  Switzerland)  for  permission  and  facilities  to  study  the  collections  in 
their  care.  I also  wish  to  thank  Ing.  P.  J.  Felder  (Natuurhistorisch  Museum,  Maastricht,  the  Netherlands)  and 
Dr.  T.  Moorkens  (Deutsche  Texaco  A.G.,  Wietze,  Federal  Republic  of  Germany)  for  critically  reading  the 
manuscript  and  for  suggesting  some  improvements. 

REFERENCES 

agassiz,  l.  and  desor,  e.  1846.  Catalogue  raisonne  des  families,  des  genres  et  des  especes  de  la  classe  des 
echinodermes  I.  Annls.  Sci.  nat.  (5)  Zoologie,  6,  305-374. 
berggren,  w.  A.  1963.  Problems  of  Paleocene  stratigraphic  correlation.  Revue  Inst.  fr.  Petrole,  18,  1448-1457. 
binkhorst  van  de  binkhorst,  j.  t.  1859.  Esquisse  geologique  et  paleontologique  des  couches  Cretacees  du 
Limbourg,  et  plus  specialement  de  la  craie  tuffeau,  avec  carte  geologique, coupes,  plan  horizontal  des  carrieres  de 
St.  Pierre,  etc.,  268  pp.  Maastricht:  Van  Osch-America. 
bosquet,  j.  1857.  Fossiele  fauna  en  flora  van  het  Krijt  van  Limburg.  In  w.  staring,  De  Bodem  van  Nederland, 
376-388.  Haarlem:  Kruseman. 

Cepek,  p.  and  moorkens,  t.  1969.  Cretaceous-Tertiary  Boundary  and  Maastrichtian-Danian  Biostratigraphy 
(Coccoliths  and  Foraminifera)  in  the  Maastrichtian  type  area.  In  w.  k.  Christensen  and  t.  birkelund  (eds.), 
Cretaceous-Tertiary  Boundary  Events  Symposium  II.  Proceedings,  137-142.  Kobenhavn. 
cotteau,  G.  1862-1867.  Paleontologie  franqaise.  Description  des  animaux  invertebres  commencee  par  Alcide 
d'Orbigny.  Terrain  Cretace,  VIII.  Echinides,  892  pp.,  pis.  1007-1204.  Paris:  Masson. 

1874.  Note  sur  les  echinides  cretaces  de  la  province  du  Hainaut.  Bull.  Soc.  geol.  Fr.  (3)  2, 368-660,  pi.  19-20. 

davies,  A.  m.  1975.  Tertiary  faunas.  A textbook  for  oilfield  Paleontologists  and  students  of  Geology.  II.  The 
sequence  of  Tertiary  faunas,  447  pp.  London:  Allen  & Unwin. 
deroo,  G.  1966.  Cytherea  (Ostracodes)  du  Maastrichtien  de  Maastricht  et  des  regions  voisines.  Resultats 
stratigraphiques  et  paleontologiques  de  leur  etude.  Meded.  Geol.  St.  (C,  V,  2)  2,  197  pp. 
desor,  e.  1855-1859.  Synopsis  des  Echinides  fossiles,  64  + 490  pp.,  44  pis.  Paris:  Reinwald. 
felder,  w.  m.  1975.  Litostratigrafie  van  het  Boven  Krijt  en  het  Dano-Montiaan  in  Zuid  Limburg  en  het 
aangrenzend  gebied.  Toelichtingen  Geologische  Overzichtskaart  van  Nederland,  63-71.  Haarlem:  Rijksgeolo- 
gische  Dienst. 

geys,  J.  f.  1979.  Salenioid  Echinoids  from  the  Maastrichtian  (Upper  Cretaceous)  of  Belgium  and  the 
Netherlands.  Palaont.  Z.  53,  296-322,  6 pis. 

— 1981.  Arbacioid  Echinoids  from  the  Maastrichtian  (Upper  Cretaceous)  of  Belgium  and  the  Netherlands. 
Ibid.  55,  257-270. 

kongiel,  r.  1935.  W sprawie  wieku  ‘siwaka’  w okolicach  Pulaw.  Prace  Tow.  Przyj.  Nauk  Wilnie,  9,  57  pp.,  8 pis. 

1939.  Materjaly  do  znajomosci  polskich  jezowcow  kredowych.  I.  Jezowcow  regularne.  Ibid.  13,  54  pp.,  3 pi. 

kruytzer,  e.  m.  and  meijer,  m.  1958.  On  the  occurrence  of  Crania  brattenburgica  (v.  Schlotheim,  1820)  in  the 
region  of  Maastricht  (the  Netherlands).  Natuurhist,  Maandbl.  11958)  11-12,  135-141. 

Lambert,  j.  1911.  Description  des  echinides  cretaces  de  la  Belgique.  II.  Echinides  de  l’etage  Senonien.  Mem.  Mus. 
r.  Hist.  Nat.  Belg.  16,  81  pp.,  3 pis. 

— and  jeannet,  a.  1928.  Nouveau  catalogue  des  moules  d’echinides  fossiles  du  Musee  d’Histoire  Naturelle  de 
Neuchatel,  executes  sous  la  direction  de  L.  Agassiz  et  E.  Desor.  Denkschr.  Schweiz.  Naturf.  Gesell.  64, 
83-233,  2 pis. 

and  thiery,  p.  1909-1925.  Essai  de  nomenclature  raisonnee  des  echinides,  607  pp.,  15  pi.  Chaumont: 

Ferriere. 

meijer,  m.  1959.  Sur  la  limite  superieure  de  l’etage  Maastrichtien  dans  la  region  type.  Bull.  Acad.  roy.  Beige 
Cl.  Sci.  45,  316-338. 

— 1965.  The  stratigraphical  distribution  of  echinoids  in  the  Chalk  and  Tuffaceous  Chalk  in  the 
neighbourhood  of  Maastricht  (Netherlands).  Meded.  Geol.  St.  ( N.S. ),  17,  21-25. 
moorkens,  t.  1972.  Foraminiferen  uit  het  stratotype  van  het  Montiaan  en  uit  onderliggende  lagen  van  de  boring 


276 


PALAEONTOLOGY,  VOLUME  25 


te  Obourg  (Met  een  overzicht  van  de  stratigrafie  van  het  Paleoceen  in  Belgie).  Natuurwet.  Tijdschr. 
54,  117-127. 

mortensen,  t.  1935.  A monograph  on  the  Echinoidea.  II.  Bothriocidaroida,  Melonechinoida,  Lepidocentroida  and 
Stirodonta,  647  pp.,  89  pis.  Kobenhavn:  Reitzel. 
mourlon,  M.  1881.  Geologie  de  la  Belgique,  317  + 392  pp.  Bruxelles:  Hayez. 

orbigny  a.  d’  1850.  Prodrome  de  Paleontologie  stratigraphique  universelle  des  animaux  mollusques  et  rayonnes, 
faisant  suite  au  Corns  elementaire  de  Paleontologie  et  de  Geologie  stratigraphique,  II,  427  pp.  Paris:  Masson. 
ravn,  J.  p.  J.  1928.  De  regulaere  echinider  i Danmarks  Kridtaflejringer.  Kgl.  Danske  Vidensk.  Selsk.  Skrifter, 
Naturv.  og  Math.  Afd.  9 (7)  1,  5-62,  pis.  1-6. 

Rasmussen,  H.  w.  1965.  The  Danian  affinities  of  the  Tuffeau  de  Ciply  in  Belgium  and  the  ‘post-Maastrichtian’  in 
the  Netherlands.  Meded.  Geol.  St.  ( N.S. ),  17,  33-38,  pis.  8-9. 
schluter,  c.  1892.  Die  regularen  Echiniden  der  norddeutschen  Kreide.  II.  Cidaridae.  Saleniidae.  Abh.  Kon. 
Preuss.  geol.  Landesamt  ( N.F .)  5,  243  pp.,  21  pis. 

smiser,  j.  1935.  A monograph  of  the  Belgian  Cretaceous  Echinoids.  Mem.  Mus.  roy.  Hist.  Nat.  Belg. 
68,  98  pp.,  9 pi. 

till,  r.  1974.  Statistical  methods  for  the  Earth  Scientist,  154  pp.  London:  Macmillan. 

ubaghs,  c.  1879.  Description  geologique  et  paleontologique  du  sol  du  Limbourg,  275  pp.  Roermond:  Romer. 


J.  F.  GEYS 


Typescript  received  7 July  1980 

Revised  typescript  received  8 December  1980 


Laboratory  for  Mineralogy,  Geology  and  Physical  Geography 
State  University  Centre  (RUCA) 
Groenenborgerlaan  171,  B-2020  Antwerpen 
Belgium 


DEVONIAN  MIOSPORE  ASSEMBLAGES 
FROM  FAIR  ISLE,  SHETLAND 

by  J.  E.  A.  MARSHALL  and  K.  C.  ALLEN 


Abstract.  Miospore  assemblages  have  been  isolated  from  a Devonian  sequence  of  Old  Red  Sandstone  facies,  on 
Fair  Isle,  Shetland.  The  special  problems  encountered  in  processing  these  palynomorphs  with  their  high 
carbonization  levels  and  subsequent  darkening  are  mentioned.  Thirty  miospore  species  are  recorded  and  their 
taxonomic  problems  and  stratigraphical  significance  are  discussed.  Comparisons  with  similar  assemblages  from 
the  northern  hemisphere  indicate  a Givetian  (in  parts  specifically  late  Givetian)  age  for  the  Fair  Isle  material.  The 
genus  Rhabdosporites  Richardson  1960  is  emended,  and  R.  langii  Richardson  1960  and  R.  parvulus  Richardson 
1965  are  combined.  The  ecological  significance  of  Geminspora  (Balme)  Owens  1971  is  discussed. 

Shetland  contains  one  of  the  most  complex  set  of  continental  Devonian  rocks  in  Britain,  with 
different  rock  sequences  juxtaposed  by  major  transcurrent  faults.  These  complex  structural 
relationships  have  proved  very  difficult  to  elucidate,  and  this  has  not  been  helped  by  the  poor 
biostratigraphic  control  between  the  different  basins.  It  is  hoped  that  palynological  contributions  will 
give  the  biostratigraphic  basis  for  a comparison  of  the  time  relations  in  these  sedimentary  basins,  and 
will  yield  information  on  the  timing  of  the  movements  along  the  major  transcurrent  faults,  about 
which  there  is  still  much  controversy  (Smith  1977).  This  paper  deals  with  the  sequence  found  in  the 
small  (5  km  x 3 km),  isolated  Devonian  outlier  of  Fair  Isle,  which  lies  39  km  south-west  of  the 
southern  tip  of  mainland  Shetland  (text-fig.  1).  Emphasis  is  placed  on  stratigraphic  palynology;  little 
attempt  has  been  made  to  formally  modify,  or  add  to  existing  taxonomy.  This  is  because  of  serious 
preservational  problems  encountered  in  the  palynological  studies,  which  are  discussed  later. 


THE  DEVONIAN  SUCCESSION  ON  FAIR  ISLE 

Although  Fair  Isle  has  a relatively  small  Devonian  outcrop,  its  geographical  position  (text-fig.  1)  is 
of  importance  because  it  is  generally  considered  (Mykura  and  Young  1969;  Mykura  1972a)  to  lie  just 
to  the  east  of  the  Walls  Boundary  Fault,  which  is  a possible  extension  of  the  Great  Glen  Fault.  It  has 
also  been  suggested  (Mykura  1976;  Donovan,  Archer,  Turner,  and  Tarling  1976)  that  it  is  part  of  the 
same  sedimentary  basin  as  that  in  which  the  Walls  Sandstone  was  deposited,  and  movements  along 
this  fault  have  placed  it  in  its  present  position.  Any  indication  therefore  of  the  precise  age  could  help 
clarify  these  palaeogeographic  relationships. 

The  most  recent  and  complete  accounts  of  the  geology  of  Fair  Isle  are  those  of  Mykura  (1972a, 
19726,  1976),  and  this  brief  synopsis  is  drawn  from  them.  The  Fair  Isle  sedimentary  sequence  (for 
which  neither  the  top  nor  the  base  is  seen),  is  composed  of  over  3000  m of  dominantly  clastic 
terrigenous  sediments,  all  steeply  dipping  east-south-east.  The  rocks  have  been  subdivided  (Mykura’s 
nomenclature  and  notation  followed  here)  into  four  stratigraphic  subdivisions  (see  text-figs.  1,  2),  on 
the  basis  of  lithological  differences.  These  units  can  be  traced  with  varying  degrees  of  accuracy  across 
the  island,  but  the  presence  of  two  major  east-north-east  trending  faults  create  local  correlation 
difficulties  which  are  only  partially  solved  by  lithological  mapping. 

The  lowest  unit  is  the  Ward  Hill  Group  (la  and  lb  in  Mykura’s  notation),  composed  of  over 
1600  m of  conglomerates  and  sandstones.  This  is  succeeded  by  the  Observatory  Group  (2a,  2b,  and 
2c)  with  over  900  m of  sandstones  and  finer  sediments  including  dolomitic  siltstones  and  mudstones. 
Above  this,  the  Vaasetter  Group  (3a  and  3b)  which  is  only  exposed  in  the  central  fault  block,  is 
composed  of  about  600  m of  conglomerates  and  sandstones,  most  of  which  are  inaccessible.  The  top 


IPalaeontology,  Vol.  25,  Part  2,  1982,  pp.  277-312,  pis.  30-33.| 


278 


PALAEONTOLOGY,  VOLUME  25 


» 36  Sample  Location 

3b  Stratigraphic  Unit  ( See  column)  0 Macro- Plants  (Chaloner '72 ) 

text-fig.  1.  Geological  map  of  Fair  Isle  (after  Mykura),  showing  location  of  palynological  samples 
and  plant  macrofossil  sites. 


MARSHALL  AND  ALLEN:  DEVONIAN  MIOSPORES 


279 


NORTHERN 

SUCCESSION 


SAMPLE 

LOCATION 


Svalbardia  »cotica0  24i 


D.  roskiliensisO 
c.f.  Th.  milleri  O 


Q co  < 

q o ^ 

o 

= 1 a 0 

o 

O ° o 

0o°o  ^ 

O o 


CENTRAL 

SUCCESSION 


BU  NESS 
GROUP 


» S 

r 


3b 

Sheep  Craig  Sndst. 

^ o o 

If 

o - J 

2 

Huni 

1b 

to 

VAASETTER 

GROUP 


OBSERVATORY 

GROUP 


□ 

□ 


1^  ol 


Fine-grained  Sediment 


Pebbly  Sandstone 


Conglomerate 


O Macro-  Plants 


1 Sample  Location 


text-fig.  2.  Stratigraphical  section  of  Fair  Isle  (after  Mykura)  showing  position  of 
palynological  samples. 


280 


PALAEONTOLOGY,  VOLUME  25 


unit,  the  Bu  Ness  Group  (4a  and  4b),  consists  of  approximately  600  m of  conglomerates,  sandstones, 
and  siltstones;  is  seen  only  in  the  northern  succession  of  the  island  where  it  is  faulted  against  the 
Observatory  Group  (2a,  2b,  and  2c). 

Increasing  tectonic  disturbance  is  seen  in  the  rocks  towards  the  southern  end  of  the  island  where 
they  exhibit  prominent  cleavage,  minor  folding,  mineralization,  and  the  emplacement  of  small  dyke 
intrusions.  The  cause  of  this  disturbance  has  been  discussed  by  Mykura  (1972a)  who  suggested  a 
possible  granitic  intrusion  lying  off  the  south-east  coast  of  the  island. 

Previous  palaeontological  records  are  very  restricted,  and  largely  limited  to  the  plant  macrofossil 
remains  described  by  Chaloner  (1972),  for  which  the  localities  are  shown  on  text-fig.  1.  The  fossil 
plants  clearly  indicate  a Devonian  age.  Chaloner  tentatively  suggests  an  age  not  older  than  middle 
Siegenian  for  the  Observatory  Group  (Unit  2a)  based  on  the  presence  of  Dawsonites  roskiliensis 
Chaloner  1972,  and  an  age  not  older  than  Eifelian,  but  more  likely  Middle  Devonian,  for  the  Bu 
Ness  Group  (Unit  4a),  based  on  the  occurrence  of  Svalbardia  scotica  Chaloner  1972.  Other  plants 
include  cf.  Thursophyton  milleri  (Salter)  Nathorst  191 5 in  the  Observatory  Group  and  hostinellid  axes 
and  cf.  Prototaxites  Dawson  1859  in  the  Bu  Ness  Group. 

The  Bu  Ness  Group  yields  the  only  animal  fossils  yet  found,  which  include  dipnoan  scales,  an 
arthrodire  plate  (' ICoccosteus  Agassiz  1844)  and  the  branchiopod  Asmussia  Pacht,  1849.  These  also 
favour  a Middle  Devonian  age  for  this  group. 


MATERIAL  AND  METHODS 

Four  samples  were  initially  provided  by  the  Institute  of  Geological  Sciences  (Edinburgh),  and  when 
two  of  these  were  found  to  contain  miospores  a field  trip  was  made  to  the  island,  and  seventy-eight 
samples  were  collected  from  fine-grained  dark-grey  to  black  clastic  rocks  for  palynological  analysis. 
Of  these,  sixty-one  were  processed,  but  only  ten  gave  assemblages  of  sufficient  quantity  and 
preservation  to  merit  further  study.  The  best-preserved  assemblages  were  from  the  Observatory 
Group;  followed  by  the  Bu  Ness  and  Ward  Hill  Groups.  Very  poor  preservation  was  found  in  the 
argillaceous  units  (2a)  of  the  Furse  Argillaceous  Beds  where  there  was  some  cleavage  development. 
The  lack  of  miospores  in  the  southern  part  of  the  island  is  of  interest,  as  it  is  in  accord  with  the 
increased  deformation  reported  by  Mykura  (1972a).  Megaspores  were  also  rare  and  usually 
fragmentary.  This  can  be  related  to  the  orthogonal  crack  sets  seen  on  some  of  the  larger  specimens, 
which  may  result  from  an  incipient  cleavage  (or  shrinkage),  causing  their  fragmentation  during 
deformation  or  in  the  processing  (Burmann  1969). 

The  samples  were  firstly  demineralized  with  hydrochloric  and  hydrofluoric  acids,  then  screen- 
washed  through  a twenty  micron  nylon  sieve,  before  being  cleaned  of  insoluble  fluorides  with 
repeated  hot  HC1  treatments  to  give  a kerogen  concentrate.  The  miospore  assemblages  were  very 
highly  carbonized  (black  in  colour),  and  the  only  oxidizing  medium  found  capable  of  clearing  them 
was  a fuming  Schulze  mixture.  The  oxidations  were  carried  out  in  a porosity-2  sinter  funnel/buchner 
flask  system  linked  with  a low-pressure  air  line  to  give  continuous  aeration  (Neves  and  Dale  1963). 
The  Schulze  mixture  was  made  up  with  fuming  nitric  acid  either  at  full  strength  or  freshly  diluted  with 
water,  but  always  at  a concentration  greater  than  70%.  The  time  taken  for  the  miospores  to  be 
oxidized  to  a level  suitable  for  transmitted  light  microscopy  varied  from  5 to  30  minutes.  In  addition, 
a marked  preferential  clearing  was  seen,  with  some  of  the  thicker  walled  (e.g.  Hystricosporites  spp.) 
miospores  never  attaining  more  than  a low  level  of  translucency.  Samples  from  different  horizons 
showed  very  different  oxidation  characteristics.  Those  from  the  Furse  Argillaceous  Beds  (2a), 
required  only  5 minutes’  oxidation  time  and  a relatively  weak  fuming  Schulze,  to  reveal  a poorly 
preserved  assemblage;  whilst  samples  from  the  Bu  Ness  Group  needed  stronger  and  longer  oxidation, 
but  gave  much  better-preserved  assemblages. 

After  oxidation  the  miospores  showed  the  phenomenon  of  redarkening  (to  opacity)  in  1 to  3 days, 
with  the  thicker-walled  miospores  deteriorating  much  more  rapidly.  This  degradation  was 
accelerated  by  heating  and  water  removal,  so  that  an  inert  non-water  miscible  plastic  mounting 
medium  could  not  be  used,  as  dehydration  was  impossible  under  normal  conditions.  Silicone  oil  was 


MARSHALL  AND  ALLEN:  DEVONIAN  MIOSPORES 


281 


also  tried,  but  the  dehydration  step  involving  tertiary  butyl  alcohol,  produced  a rapid  darkening 
reaction  in  the  cleared  sample.  Eventually,  glycerol  jelly  was  used  as  the  mounting  medium,  and  this 
gave  assemblages  which  could  be  studied  for  up  to  3 days.  The  breakdown  and  darkening  of  the 
exines  when  mounted  in  glycerol  jelly  was  noticeable,  with  degradational  products  imparting  a yellow 
stain  to  the  mounting  medium.  Miospores  with  thicker  exines  (e.g.  Hystricosporites,  Geminospora) 
showed  a more  rapid  return  to  opacity,  and  had  to  be  studied  during  the  first  day. 

A subsequent  study  of  palynological  assemblages  from  the  Devonian  of  the  south-east  mainland 
of  Shetland  also  showed  this  oxidation  problem.  There,  it  was  much  more  serious,  with  viability  times 
of  about  5 minutes  in  glycerol  jelly.  During  this  time  the  miospores  could  be  seen  to  redarken  and 
start  to  dissolve.  It  was  then  that  a special  technique  was  developed  which  involved  a matching  of  the 
oxidizing  media  strength  to  a miospore  assemblage  preservation,  and  the  use  of  a rapid  drying 
technique  which  dehydrated  the  miospores  faster  than  they  could  redarken,  so  that  an  inert  plastic 
mountant  could  be  used  (Marshall  1980). 

The  assemblages  from  Fair  Isle  were  all  studied  from  repeated  oxidation,  with  fresh  assemblage 
slides  being  made  up  every  2 to  3 days.  The  study  was  therefore  limited  by  the  amount  of  organic 
residue  per  sample;  some  giving  enough  material  for  seventy  to  eighty  slides,  others  only  having 
sufficient  for  four  or  five.  All  the  slides  have  been  kept,  and  co-ordinates  given  for  illustrated 
specimens  in  the  event  that  advances  in  techniques,  such  as  infra-red  microscopy,  will  enable 
palynomorphs  to  be  studied  without  further  recourse  to  oxidative  clearing. 

One  difficulty  in  dealing  with  such  highly  carbonized  assemblages  (‘vitrinite’  reflectance 
measurements  from  Fair  Isle  give  values  of  approx.  4 to  5%,  pers.  comm.  Dr.  J.  M.  Jones),  is  that  the 
identification  of  reworked  components  by  colour  differences  is  not  possible.  Three  species  are 
described  ( Emphanisporites  rotatus,  Camptozonotriletes  aliquantus,  and  Grandispora  Inaumovii) 
which  occur  in  a very  low  proportion,  and  it  is  possible  that  these  may  represent  reworked 
components  (see  Clayton,  Higgs,  and  Keegan  1977  for  a discussion  of  Emphanisporites).  However, 
until  one  is  better  able  to  recognize  features  of  reworking  such  as  breakage  and  erosion  (as  in  Birks 
1970),  both  this  possibility  and  that  of  their  continued  presence  as  minor  and  rare  elements  in  the 
flora  must  be  considered. 


PREVIOUS  PALYNOLOGICAL  STUDIES  FROM  SHETLAND  AND 
ADJACENT  AREAS 

The  pioneer  work  on  Old  Red  Sandstone  palynology  in  Britain  was  carried  out  by  Lang  (1925)  as  part 
of  his  study  of  the  Orcadian  Basin  flora.  Later,  Richardson  produced  a series  of  papers  (1960,  1962, 
1965)  on  the  classical  Orcadian  area.  Since  then,  very  little  work  has  been  published  except  for  brief 
taxonomic  lists  as,  for  example,  Donovan,  Collins,  Rowlands,  and  Archer  (1978),  who  give  an 
account  of  a miospore  assemblage  from  the  island  of  Foula  (western  Shetland).  Two  doctoral 
theses  have  included  a certain  amount  of  Orcadian  palynology.  Fannin  (1970),  in  conjunction  with 
Richardson,  has  tabulated  information  on  the  palaeoecology  and  stratigraphic  distribution  of 
miospore  assemblages  from  Orkney,  whilst  Fletcher  (1976)  studied  the  megaspores  of  the  Melby 
Fish  Beds.  Although  a useful  amount  of  data  has  accumulated  on  the  taxonomy  of  Devonian 
spores  from  the  Orcadian  Basin,  very  little  is  known  of  their  detailed  stratigraphic  distribution, 
and  this  has  been  a major  handicap  to  any  precise  correlation. 


STATISTICS  AND  DATA  HANDLING 

The  treatment  of  the  simple  bivariate  statistical  data,  such  as  the  exoexine  and  intexine  diameters  of 
miospores,  presents  problems  for  data  handling  and  statistical  testing.  Classical  regression  analysis  is 
not  applicable,  as  there  are  no  dependent  and  independent  variates;  furthermore,  there  is  no  easy 
solution  to  fit  a best  line  to  this  type  of  two-error  data.  Various  iterative  methods  such  as  those  described 
by  Williamson  ( 1 968),  Y ork  ( 1 966),  and  comparatively  reviewed  by  Brooks,  Hart,  and  Wendt  ( 1 972), 


282 


PALAEONTOLOGY,  VOLUME  25 


demand  statistical  estimates  for  the  variances  of  each  data  point.  These  are  certainly  not  worth  the 
extra  time  involved  for  the  quality  of  data  produced  on  the  Fair  Isle  material.  An  approximate 
method  is  the  Reduced  Major  Axis  line  as  described  by  Kermack  and  Haldane  (1950),  and  further 
documented  by  Till  ( 1 974)  and  others.  The  statistical  validity  of  the  line  produced  is  in  doubt,  but  as  it 
appears  correct  (i.e.  passes  through  the  centre  of  the  experimental  scatter),  it  is  a useful  means  of 
comparing  populations,  providing  not  too  much  reliance  is  placed  on  any  subsequent  statistical  tests 
based  upon  it. 

The  data  in  the  bivariate  plots  were  in  fact  computed  for  logarithmic  and  linear  fits,  using  both 
classical  regression  and  the  reduced  major  axis  line  to  find  the  best  fit.  The  latter  was  found  to  be  more 
successful  in  describing  a line  through  a set  of  data  points.  However,  the  logarithmic  fit  did  not  give 
much  advantage  over  the  linear  model.  When  data  values  were  taken  from  published  graphical  plots 
(Richardson  1965),  an  electronic  digitizing  machine  (D-MAC)  was  used  to  provide  more  accurate 
results  and  known  estimates  of  error. 

NOMENCLATURE  AND  SYSTEMATICS 

The  morphological  terminology  used  is  that  of  Smith  and  Butterworth  ( 1 967),  and  their  classification 
system  is  followed  except  for  the  inclusion  of  Hystricosporites,  Ancyrospora,  and  Geminospora  in  an 
incertae  sedis  group,  following  the  practice  of  Streel  (in  Becker,  Bless,  Streel,  and  Thorez  1974).  No 
categories  higher  than  infraturmae  are  used,  and  the  retusoid  miospores  are  retained  in  the  Laevigati 
and  Apiculati,  because  haptotypic  features  are  not  here  regarded  as  being  of  major  classificatory 
importance. 

Figured  material  is  housed  in  the  palaeobotanical  collection  of  the  Department  of  Botany,  Bristol 
University.  Co-ordinates  given  refer  to  a Leitz  Orthoplan  microscope  no.  715334.  A ringed  reference 
slide  is  also  provided.  Each  sample  number  (i.e.  Fair  66)  is  followed  by  a strew  slide  number  and  then 
the  appropriate  co-ordinates. 

Infraturma  laevigati  Bennie  and  Kidston  emend.  Potonie  and  Kremp  1954 
Genus  trileites  Erdtman  ex  Potonie  1956 

Type  species.  Trileites  spurius  Dijkstra  emend.  Potonie  1956 

Trileites  langii  Richardson  1965 
Plate  30,  fig.  1 

Dimensions  (two  specimens).  Maximum  equatorial  diameters  141  and  160  p.m. 

Lithostratigraphic  Range.  Ward  Hill  Sandstone  (lb);  sample  Fair  66  (see  text-fig.  2). 


EXPLANATION  OF  PLATE  30 
All  figures  x 400  unless  otherwise  stated. 

Fig.  1.  Trileites  langii  Richardson  1965.  Fair  66.2,  48.9,  98.7 

Figs.  2, 5.  Acinosporites  lindlarensis  Riegel  1968  var.  minor  McGregor  and  Camfield  1976.  Fair  37.80, 37.4, 104.8, 
x 1000  Distal  and  proximal  surfaces  respectively. 

Figs.  3,  4.  Camptozonotriletes  aliquantus  Allen  1965.  3,  Distal  view.  4,  x 1000  Distal  surface  showing 
sculpture.  Both  Fair  66.2,  17.5,  109.8 

Fig.  6.  Retusotriletes  rotundus  (Streel)  Lele  and  Streel  1969.  Fair  37.79,  19.9,  106.0. 

Fig.  7.  Calamospora  atava  (Naumova)  McGregor  1964.  Fair  37.79,  26.9,  96.5 
Fig.  8.  Emphanisporites  rotatus  (McGregor)  McGregor  1973.  Fair  28.32,  1 14.6,  27.2 
Fig.  9.  Chelinospora  concinna  Allen  1965.  x 1000.  Fair  37.1,  25.9,  1 15.3 
Fig.  10.  Convolutispora  disparalis  Allen  1965.  x 1000  Fair  24.40,  50.0,  100.3 


PLATE  30 


284 


PALAEONTOLOGY,  VOLUME  25 


Remarks.  Similar  to  the  population  described  by  Richardson  (1965),  but  smaller  in  over-all  diameter. 
However,  it  is  comparable  with  the  miospores  described  by  Chi  and  Hills  (1976),  and  is  here  referred 
to  Trileites  langii. 


Genus  retusotriletes  Naumova  1953  emend.  Streel  1964 
Type  species.  Retusotriletes  simplex  Naumova  1953 

Retusotriletes  rotundus  (Streel)  Lele  and  Streel  1969 
Plate  30,  fig.  6 

Synonymy,  see  McGregor  (1973,  p.  20). 

Dimensions  (eighteen  specimens).  Maximum  equatorial  diameter  56-80  /urn  (mean  67  /nm). 

Lithostratigraphic  range.  Ward  Hill,  Observatory,  and  Bu  Ness  Groups  (lb  to  4a).  Found  in  all  samples 
examined. 

Remarks.  This  is  an  example  from  the  wide  variety  of  retusoid  forms  which  occur  in  the  Fair  Isle 
assemblages. 

Genus  calamospora  Schopf,  Wilson,  and  Bentall  1944 
Type  species.  Calamospora  hartungiana  Schopf,  Wilson,  and  Bentall  1944 

Calamospora  atava  (Naumova)  McGregor  1964 
Plate  30,  fig.  7 

Dimensions  (twenty-four  specimens).  Maximum  equatorial  diameter  27-71  fan  (mean  56  /j.m). 

Lithostratigraphic  range.  Ward  Hill  Group  to  Bu  Ness  Group  (lb  to  4a).  In  all  samples  examined  except  Fair  33 
(see  text-fig.  2). 

Remarks.  Compares  well  with  the  emended  species  described  by  McGregor  (1964),  but  exhibits  a 
slightly  greater  size  range. 

Infraturma  murornati  Potonie  and  Kremp  1954 
Genus  convolutispora  Hoffmeister,  Staplin,  and  Malloy  1955 

Type  species.  Convolutispora  florida  Hoffmeister,  Staplin,  and  Malloy  1955 

Convolutispora  dispar alis  Allen  1965 
Plate  30,  fig.  10 

Dimensions  (six  specimens).  Maximum  equatorial  diameter  36-54  jun  (mean  46  ^m). 

Lithostratigraphic  range.  Bu  Ness  Group  (4a);  samples  Fair  24  and  28. 

Remarks.  As  stated  in  Allen  (1965)  it  is  thought  likely  that  sculptural  elements  of  this  type  result  from 
corrosion  of  the  exine.  It  is  noticeable  that  a wide  range  of  sculpture  is  present,  with  transitional 
forms  resembling  Raistrickia  Potonie  and  Kremp  1954. 

Genus  emphanisporites  McGregor  1961 
Type  species.  Emphanisporites  rotatus  McGregor  1961 


MARSHALL  AND  ALLEN:  DEVONIAN  MIOSPORES 


285 


Emphanisporites  rotatus  McGregor  emend.  McGregor  1973 
Plate  30,  fig.  8 

Synonymy,  see  McGregor  (1973,  p.  46). 

Dimensions  (three  specimens).  Maximum  equatorial  diameter  33,  38,  and  50  /xm. 

Lithostratigraphic  range.  Observatory  and  Bu  Ness  Groups  (4a  and  2c);  samples  Fair  28  and  37. 

Remarks.  Clayton  et  al.  (1977)  have  documented  a series  of  sporadic  occurrences  of  Emphanisporites 
spp.  from  the  later  Devonian  and  early  Carboniferous  of  southern  Ireland.  Similar  sporadic 
occurrences  were  also  noted  by  Richardson  (1965)  from  the  Eday  Flags  of  the  Orcadian  Basin.  It 
seems  likely  that  the  Fair  Isle  occurrences  are  only  rare  examples  of  a minor  but  persistent  element  in 
the  flora.  The  decision  when  to  regard  elements  of  an  assemblage  as  reworked  or  rare  is  often  difficult 
(see  Captozonotrilites  aliquantus),  and  should  be  based  on  features  such  as  stratigraphic  persistence, 
association  with  other  possible  reworked  elements,  and  obvious  signs  of  physical  reworking  (see 
Birks  1970). 


Genus  acinosporites  Richardson  1965 
Type  species.  Acinosporites  acanthomammillatus  Richardson  1965 

Acinosporites  lindlarensis  Riegel  1968  var.  minor  McGregor  and  Camfield  1976 
Plate  30,  figs.  2,  5 

Dimensions  (one  specimen).  Maximum  equatorial  diameter  48  /xm. 

Lithostratigraphical  range.  Observatory  Group  (2c);  sample  Fair  37. 

Remarks.  The  Fair  Isle  miospore  closely  resembles  specimens  described  by  McGregor  and  Camfield 
(1976)  from  sediments  of  Emsian  to  Givetian  age  from  the  Moose  River  Basin,  Ontario.  Although 
this  species  has  a Geminospora  organization,  the  authors  follow  McGregor  and  Camfield  (1976),  by 
placing  it  in  Acinosporites. 

Infraturma  crassiti  Bharadwaj  and  Venkatachala  1962 
Genus  aneurospora  (Streel)  Streel  1967 

Type  species.  Aneurospora  goensis  Steel  1964 

Aneurospora  greggsii  (McGregor)  Streel  in  Becker  et  al.  1974 
Plate  31,  fig.  1 

Synonymy,  see  Streel  in  Becker  et  al.  (1974,  p.  24). 

Dimensions  (eight  specimens).  Maximum  equatorial  diameter  68-116  /xm  (mean  84  /xm) 

Lithostratigraphic  range,  restricted  to  the  Bu  Ness  Group  (4a);  samples  Fair  24,  28,  31,  and  33. 

Comparisons.  Similar  problems  to  those  noted  by  Lele  and  Streel  (1969)  and  Streel  (1972)  were 
encountered  in  separating  this  genus  from  Geminospora.  A possible  synonymy  is  with  Archaeo- 
zonotriletes  nalivkinii  Naumova  as  figured  by  Chibrikova  (1977,  pi.  XIX,  fig.  11),  but  since  no 
description  was  given,  it  is  impossible  to  make  a detailed  comparison.  Certain  specimens  also  show 
similarities  with  Geminospora  svalbardiae  Allen  1965. 

Infraturma  cingulicavati  Smith  and  Butterworth  1967 
Genus  camptozonotriletes  Staplin  1960 

Type  species.  Camptozonotriletes  vermiculatus  Staplin  1960 


286 


PALAEONTOLOGY,  VOLUME  25 
Camptozonotriletes  aliquantus  Allen  1965 
Plate  30,  figs.  3,  4 

Dimensions  (one  specimen).  Maximum  exoexine  diameter  77  /un,  maximum  intexine  diameter  57^m. 
Lithostratigraphic  range.  Ward  Hill  Group  (lb),  sample  Fair  66. 

Remarks.  The  known  stratigraphic  occurrences  for  this  species  are  Siegenian  to  Lower  Eifelian  (Allen 
1967);  Upper  Siegenian  to  Lower  Emsian  (Massa  and  Moreau-Benoit  1976),  and  lower  Eifelian 
(Riegel,  1973,  1974).  These  occurrences  are  significantly  different  from  the  Givetian  age  assigned  to 
the  Fair  Isle  succession,  and  its  appearance  may  be  the  result  of  reworking. 

Genus  denosporites  Berry  emend.,  Potonie  and  Kremp  1954 
Type  species.  Densosporites  covensis  Berry  1937 

Remarks.  Butterworth  et  coll.  1964  (and  in  Staplin  and  Jansonius  1964)  in  an  emendation  of  the 
densospore  group  of  miospores,  provided  unified  limits  (based  largely  on  Carboniferous  material)  for 
the  subdivision  of  generic  groups.  These  genera  have  not  proved  applicable  for  the  Devonian 
densospores  from  Fair  Isle,  but  until  a more  unified  revision  is  carried  out,  we  feel  they  should  be 
retained. 

Densosporites  devonicus  Richardson  1960 
Plate  31,  fig.  12 

Dimensions  (thirty-six  specimens).  Maximum  equatorial  diameter  55-120  /xm  (mean  85  //.m). 

Lithostratigraphic  range.  Ward  Hill  to  Bu  Ness  Group  (lb  to  4a).  In  all  samples  examined. 

Remarks.  Richardson  (1965),  in  his  study  of  Middle  Devonian  miospores  from  the  Orcadian  Basin, 
gave  the  sculptural  details,  and  the  relative  widths  of  the  light  and  dark  zones  of  the  cingulum  as 
criteria  for  distinguishing  Densosporites  devonicus  from  D.  orcadensis.  These  same  criteria  were  used 
in  Fair  Isle  in  an  attempt  to  substantiate  the  differences  between  the  two  species,  but  no  systematic 
variation  of  these  characters,  as  claimed  by  Richardson  (1965,  p.  581),  was  found.  McGregor  and 
Camfield  (1976)  and  McGregor  (19796)  also  had  difficulty  in  distinguishing  between  the  two  species. 
This  raises  doubts  as  to  the  significance  of  the  stratigraphic  distribution  of  the  two  species  as  recorded 
by  Richardson  (1965). 


Genus  samarisporites  Richardson  1965 
Type  species.  Samarisporites  orcadensis  (Richardson)  Richardson  1965. 


EXPLANATION  OF  PLATE  31 

All  figures  x 400  unless  otherwise  stated. 

Fig.  1.  Aneurospora  greggsii  (McGregor)  Streel  1974.  Fair  31.8,  28.5.  100.6 
Fig.  2.  Cirratriradites  avius  Allen  1965.  Fair  66.2,  12.8,  95.2 

Fig.  3.  Samarisporites  orcadensis  (Richardson)  Richardson  1965.  Fair  31.5,  12.1,  104.5 
Figs.  4,  7.  Cirratriradites  sp.  A.  4,  x 1000.  Detail  of  distal  sculpture.  7,  x400.  Fair  66.9,  13.9,  110.6 
Figs.  5,  6.  Samarisporites  mediconus  (Richardson)  Richardson  1965.  5,  Proximal  view.  6,  Distal  view.  Fair 
37.71,  13.5,  101.2 

Figs.  8,  9.  Samarisporites  conannulatus  (Richardson)  Richardson  1965.  8,  Distal  view.  9,  Proximal  view.  Fair 
37.80,  5.1,  100.3 

Fig.  10.  Auroraspora  macromanifestus  (Hacquebard)  Richardson  1960.  Fair  37.82,  14.8,  1 12.4 
Fig.  1 1 . Auroraspora  micromanifestus  (Hacquebard)  Richardson  1960.  Fair  37.82,  7.1,  111.0 
Fig.  12.  Densosporites  devonicus  Richardson  1960.  Fair  66.3,  4.0,  95.7 


PLATE  31 


MARSHALL  and  ALLEN,  Devonian  miospores 


288 


PALAEONTOLOGY,  VOLUME  25 


Remarks.  The  generic  status  of  Samarisporites  is  similar  to  that  of  the  Devonian  Densosporites,  in 
being  an  element  of  the  densospore  complex.  The  genus  Samarisporites,  as  originally  proposed  by 
Richardson  (1965),  was  to  accommodate  zonate  spores  with  a variety  of  solely  distal  sculptural 
elements.  The  justification  for  the  erection  of  this  genus  was  that  its  initial  assignation  to 
Cristatisporites  (Potonie  and  Kremp  1954)  was  invalid,  because  the  latter  has  both  proximal  and 
distal  sculpture.  However,  it  was  suggested  by  Playford  (1971),  that  with  the  emendation  of 
Cristatisporites  (Butterworth  et  coll.  1964)  as  part  of  a general  reorganization  of  the  densospore 
group,  that  the  use  of  Samarisporites  as  a distinct  generic  category  could  be  abandoned.  The 
emendation,  however,  still  includes  the  possible  presence  of  a proximal  sculpture  in  the  form  of  a ring 
of  setae.  It  also  restricts  the  distal  sculpture  to  being  dominantly  mammoid  in  type,  and  showing  no 
differentiation  in  form.  The  present  usage  of  Samarisporites  includes  forms  with  a wide  variety  of 
distal  sculpture  (e.g.  coni,  cristae,  verrucae)  which  cannot  be  accommodated  within  Cristatisporites 
(sensu  Butterworth  et  coll.  1964).  It  is  proposed  therefore  to  use  Samarisporites  for  these  species,  until 
a more  unified  set  of  limits  for  the  densospore  group  is  proposed,  which  accommodates  both 
Devonian  and  Carboniferous  representatives. 

Richardson  (1960,  1965)  erected  three  species  of  Samarisporites  based  largely  on  sculptural 
differences  and  their  distribution  on  the  distal  surface  (e.g.  central  packing  in  S.  mediconus , ring 
development  in  S.  conannulatus).  The  Fair  Isle  populations  contain  intermediate  forms,  and  there  is 
a complete  morphological  transition  series  between  Richardson’s  S.  mediconus,  S.  orcadensis,  and  S. 
conannulatus,  which  can  be  considered  as  occupying  distinct  positions  on  the  various  trends.  It  would 
be  interesting  to  speculate  whether  the  continuous  variation  this  plexus  of  species  exhibits,  could  be 
treated  in  the  same  way  as  in  the  morphon  concept  recently  outlined  by  Van  der  Zwan  (1979,  1980). 
However,  not  enough  individuals  have  yet  been  found  to  systematically  describe  the  variation  both  in 
a morphological  and  stratigraphical  sense,  either  to  record  separate  species  or  varieties  in  a morphon, 
or  to  combine  them  as  a single  species. 


Samarisporites  mediconus  (Richardson)  Richardson  1965 
Plate  31,  figs.  5,6 

Dimensions  (eight  specimens).  Maximum  equatorial  diameter  73-130  ^ m (mean  108  ^m),  cingulum  10-26  ^m 
wide  (mean  17^m). 

Lithostratigraphic  range.  Ward  Hill  to  Bu  Ness  Group  (lb  to  4a);  samples  Fair  31,  37,  and  66. 

Remarks.  Closely  resembles  the  type  species,  except  that  there  is  a greater  variety  of  labra 
morphology. 


Samarisporites  orcadensis  (Richardson)  Richardson  1965 
Plate  31,  fig.  3 

Dimensions  (three  specimens).  Maximum  equatorial  diameter  88,  97,  and  146  /j,m.  Cingulum  width  15-20  /xm. 
Lithostratigraphic  range.  Observatory  and  Bu  Ness  Groups  (2c  and  4a);  samples  Fair  31,  33,  and  36. 

Remarks.  The  three  individuals  found  compare  closely  with  Richardson’s  holotype.  However,  they 
show  greater  variety  of  labra  morphology  in  both  length  and  height. 


Samarisporites  conannulatus  (Richardson)  Richardson  1965 
Plate  31,  figs.  8,  9 

Dimensions  (five  specimens).  Maximum  equatorial  diameter  104-120  (mean  113  ^m).  Cingulum  14-21 
wide,  maximum  width  interradially. 

Lithostratigraphic  range.  Ward  Hill  to  Bu  Ness  Groups  (lb  to  4a);  samples  Fair  31,  36,  37,  and  64. 


MARSHALL  AND  ALLEN:  DEVONIAN  MIOSPORES 


289 


Remarks.  Compares  closely  with  the  type  species  except  that  the  labra  show  a greater  amount  of 
variation  in  length  and  height. 

Genus  cirratriradites  Wilson  and  Coe  1940. 

Type  species.  Cirratriradites  saturnii  (Ibrahim)  Schopf,  Wilson,  and  Bentall  1944 
Cirratriradites  avius  Allen  1965 
Plate  31,  fig.  2 

Dimensions.  Maximum  equatorial  diameter  80-130  /xm  (mean  105  /xm),  twenty-eight  species  measured. 
Maximum  intexine  diameter  58-78  /xm  (mean  67  /xm),  eight  specimens  measured. 

Lithostratigraphic  range.  Bu  Ness  and  Ward  Hill  Groups  (4a  and  lb);  samples  Fair  24,  28,  33,  64,  and  66. 

Remarks.  Although  the  type  species  of  this  genus  is  characterized  by  having  distinctive  fovea,  the 
formal  designation  of  the  genus  considered  this  feature  to  be  of  little  importance.  The  assignment  by 
Allen  (1965)  of  Cirratriradites  avius  to  this  genus,  was  made  on  the  presence  of  the  reduced  sculpture 
and  apparently  thin  equatorial  flange.  As  it  may  seem  desirable  to  restrict  the  use  of  Cirratriradites  to 
miospores  with  distinctive  fovea,  it  may  in  the  future  be  necessary  to  refer  this,  and  other  species  of 
similar  organization,  to  a new  genus  within  the  densospore  complex.  The  intexine  is  only  seen  in 
overmacerated  specimens.  Hymenozonotriletes punctomonogrammos  Arkhangelskaya  (in  Filiminova 
and  Arkhangelskaya  1963),  from  the  Mosolovian  of  the  Central  Devonian  Field  is  clearly  similar. 

Cirratriradites  sp.  A 
Plate  31,  figs.  4,  7 

Description.  Miospores  trilete;  camerate;  amb  triangular.  Suturae  indistinct.  Exine  two-layered,  intexine  thin 
and  closely  appressed  to  the  exoexine.  Exoexine  infrapunctuate,  distally  sculptured  with  muri  (2-6  /xm  high) 
fused  into  a cristoreticulate  pattern,  more  dense  in  central  area.  Cingulum  with  an  apparent  thin  margin,  but 
possessing  a thicker  inner  zone. 

Dimensions  (two  specimens).  Maximum  equatorial  diameter  125  and  135  /xm. 

Lithostratigraphic  range.  Ward  Hill  Sandstone  (lb);  sample  Fair  66  only. 

Remarks.  Differs  from  Cirratriradites  avius  in  having  a dense  distal  sculpture  of  muri.  Hymenozono- 
triletes monogrammos  Arkhangelskaya  (1963)  recorded  from  the  Vorobyevskian,  Starskoolian, 
Chernoyarian,  and  Mosolovian  beds  (Eifelian  to  Givetian)  of  the  Russian  Platform,  Volga-Urals, 
and  Karatau  (Filimonova  and  Arkhangelskaya  1963,  Arkhangelskaya  1974,  Raskatova  1969, 
Chibrikova  1977)  is  very  similar,  having  a distal  network  of  muri  and  may  prove  to  be  synonymous. 

Infraturma  patinati  Butterworth  and  Williams  emend.  Smith  and  Butterworth  1967 
Genus  chelinospora  Allen  1 965 

Type  species.  Chelinospora  concinna  Allen  1965 

Chelinospora  concinna  Allen  1965 
Plate  30,  fig.  9 

Dimensions  (eight  specimens).  Maximum  equatorial  diameter  36-65  /xm  (mean  45  /xm). 

Lithostratigraphic  range.  Observatory  Group  (3c);  samples  Fair  36  ad  37. 

Infraturma  monopseudosacciti  Smith  and  Butterworth  1967 
Genus  auroraspora  Hoffmeister,  Staplin,  and  Malloy  1955 

Type  species.  Auroraspora  solisortus  Hoffmeister,  Staplin,  and  Malloy  1955. 


290 


PALAEONTOLOGY,  VOLUME  25 


Population  variation  in  Auroraspora.  Four  species  of  Auroraspora  have  been  described  from  the 
Middle  Devonian  of  the  Orcadian  Basin,  separated  on  features  such  as  size,  shape,  and  the  relative 
dimensions  of  the  intexine  and  exoexine  layers.  The  exoexine  and  intexing  diameters  for  the  type- 
species  populations  of  Auroraspora  micromanifestus  Richardson  1960,  A.  macromanifestus  Richard- 
son 1960,  A.  minuta  Richardson  1965,  and  A.  aurora  Richardson  1960  (with  holotype  positions 


40  60  80  100  120  140  160  180  200  220  240 

DIAMETER  OF  EXOEXINE  (pm) 


text-fig.  3.  Graphical  plot  of  exoexine  and  intexine  diameters  for  Auroraspora  spp. 
from  Fair  Isle.  The  stars  mark  the  holotype  positions  of  species  recorded  by 
Richardson  1960,  1965.  The  regression  line  is  for  all  Auroraspora  from  Fair  Isle. 


marked),  are  plotted  in  text-fig.  3,  and  as  can  be  readily  seen,  the  degree  of  separation  is  not  sufficient 
to  delimit  the  species  purely  by  this  method.  The  differences  could  be  related  to  ontogenetic  variation 
in  a sporangium,  with  a changing  ratio  of  exoexine  to  intexine  diameters  as  the  miospores  increase  in 
size.  Other  characters  cited  as  important  in  delimiting  species,  such  as  shape  and  eccentricity,  are 
probably  not  mutually  exclusive,  and  can  be  related  to  the  presence  or  absence  of  labra  and  their 
degree  of  development,  which  are  again  themselves  a poor  classificatory  character.  Prominent  labra 
on  a small  miospore  will  give  a triangular  shape,  whereas  small  labra  on  a large  miospore  often  results 
in  a circular  shape  which,  when  compressed,  can  increase  the  eccentricity  of  the  two  exine  layers.  A 
histogram  of  the  exoexine  size  distribution  (text-fig.  4)  also  provides  valuable  information,  as  a 
skewed  population  is  seen,  which  is  reminiscent  of  the  pattern  produced  during  the  development  of 
heterospory  (see  Chaloner  1967).  Size  distribution  is  also  given  for  individual  collection  samples,  and 
this  shows  the  effects  of  some  sedimentary  sorting.  However,  bimodal  populations  are  apparent  in 
three  of  the  four  assemblages,  as  well  as  in  the  five  point  moving  average  line. 

It  is  believed  that  the  species  differences  are  ontogenetic  and  perhaps  reinforced  by  an  incipient 
heterospory,  but  more  information  is  needed  on  miospore  populations  found  both  dispersed  and  in 
situ  from  single  sporangia,  before  combining  certain  miospore  species.  A note  of  caution  is  also 
necessary  in  case  the  lumping  of  species  causes  the  loss  of  potentially  valuable  morphological  data 
which  may  be  of  stratigraphic  value  as  populations  of  Auroraspora  succeed  each  other,  with  slightly 
different  degrees  of  size  distribution  (viz.  the  possible  biostratigraphic  value  of  changing  over-all 
diameters  in  Retusotriletes  from  the  Russian  Platform  as  shown  by  Naumova  1953,  p.  18,  text-fig.  6). 
A sensible  course  in  handling  species  such  as  those  of  Auroraspora , is  to  give  precise  descriptions  and 
details  of  dimensions  as  illustrated  in  text-figs.  3 and  4. 


MARSHALL  AND  ALLEN:  DEVONIAN  MIOSPORES 


291 


text-fig.  4.  Frequency  distribution  plot  for  Auroraspora.  Note  skewed  distribution  present  in  different  samples 
and  remains  on  five  point  moving  average  lines. 


Auroraspora  macromanifestus  (Hacquebard)  Richardson  1960 
Plate  31,  fig.  10 

Dimensions  (seventeen  specimens).  Maximum  exoexine  diameters  120-200  ^m  (mean  153  /xm),  maximum 
intexine  diameters  60-115  /xm  (mean  83  /xm).  See  text-fig.  3 for  size  distributions  and  ratios  of  central-body 
diameter  to  whole-body  diameter  (mean  1 : 4). 

Lithostratigraphic  range.  Ward  Hill  Group  to  Bu  Ness  Group  (lb  to  4a);  samples  Fair  8,  31,  37,  and  66. 

Remarks.  Differs  from  specimens  described  by  Richardson  (1960)  from  the  Middle  Devonian  of  the 
Cromarty  area  (Scotland)  only  in  size  range. 

Auroraspora  micromanifestus  (Hacquebard)  Richardson  1960 
Plate  31,  fig.  11 

Dimensions  (one  hundred  and  two  specimens).  Maximum  exoexine  diameter  65-120  /xm  (mean  89  /xm), 
maximum  intexene  diameter  35-85  /im  (mean  56  /x m).  See  text-fig.  3 for  size  distribution  and  ratio  of  central- 
body  diameter  to  whole-body  diameter  (mean  1 : 2). 

Lithostratigraphic  range.  Ward  Hill  to  Bu  Ness  Group  (lb  to  4a);  samples  Fair  8, 24, 28,  31,  33, 36, 37, 64, 66,  and 
67. 

Genus  grandispora  (Hoffmeister,  Staplin,  and  Malloy)  Neves  and  Owens  1966  sensu  Playford  1971 
Type  species.  Grandispora  spinosa  Hoffmeister,  Staplin,  and  Malloy  1955. 

Grandispora  Inaumovii  (Kedo)  McGregor  1973 
Plate  32,  fig.  5 

Description.  Miospore  trilete,  camerate,  amb  ovate.  Suturae  accompanied  by  wavy  labra  up  to  6-5  /xm  in  height 
and  individually  2 /xm  wide.  Exine  two-layered,  intexine  laevigate,  indistinct,  exoexine  1 /xm  thick,  shagreenate, 
distally  sculptured  with  sparse,  gently  tapering  spines.  There  are  eleven  spines  around  the  equatorial  periphery 
which  measure  up  to  20  /xm  in  length. 


292 


PALAEONTOLOGY,  VOLUME  25 


Dimensions  (one  specimen).  Exoexine  diameter  130  x 106  /xm,  intexine  diameter  80  x 62  /xm. 

Lithostratigraphic  range.  Observatory  Group  (2c);  sample  Fair  37. 

Comparisons.  This  miospore  closely  resembles  Grandispora  Inaumovii  described  by  McGregor  and 
Camfield  (1976)  from  the  Middle  Devonian  of  the  Hudson  Bay  area  (Canada)  and  by  McGregor 
(1973)  from  the  Middle  Devonian  of  Gaspe  (Canada). 

Grandispora  velata  (Eisenack)  Playford  1971 
Plate  32,  fig.  2 

Synonymy,  see  Owens  1971,  p.  46 

Dimensions  (nine  specimens).  Maximum  exoexine  diameter  94-123  /xm  (mean  111  /xm),  maximum  intexine 
diameter  57-80  /xm  (mean  70  /xm). 

Lithostratigraphic  range.  Ward  Hill  to  Bu  Ness  Groups  (lb  to  4a);  samples  Fair  24,  36,  37,  and  66. 

Comparisons.  The  Fair  Isle  population  differs  from  the  type  population  in  having  a slightly  smaller 
sculptural  size  range. 


Grandispora  protea  (Naumova)  Moreau-Benoit  1980 
Plate  32,  figs.  3,  4 

Dimensions  (ten  specimens).  Maximum  exoexine  diameter  84-184  /xm  (mean  136  /xm),  maximum  intexine 
diameter  39-97  /xm  (mean  70  /xm). 

Lithostratigraphic  range.  Ward  Hill  to  Bu  Ness  Groups  (lb  to  4a);  samples  Fair  24,  36,  37,  and  66. 

Genus  velamisporites  Bharadwaj  and  Venkatachala  1962 
Type  species.  Velamisporites  rugosus  Bharadwaj  and  Venkatachala  1962. 

Remarks.  Evans  (1970)  has  interpreted  the  organization  of  the  genus  Perotrilites  Erdtman  ex  Couper 
1953  as  being  zonate  and  not  perinate  as  previously  supposed  (see  Playford  1971).  Prior  to  this, 
Palaeozoic  perinate  miospores  were  placed  in  Perotrilites,  but  these  are  now  removed  to 
Velamisporites,  which  was  originally  thought  to  be  a junior  synonym,  but  now  forms  a satisfactory 
genus  for  miospores  with  such  an  organization. 

Velamisporites  sp.  A 
Plate  32,  fig.  1 

Description.  Miospores  trilete,  camerate,  amb  triangular  to  rounded.  Suturae  accompanied  by  labra  commonly 
1 -5-2-5  /xm  high  (maximum  4 /xm)  which  continue  as  folds  on  to  the  perine  layer.  Exine  three-layered;  intexine 


EXPLANATION  OF  PLATE  32 
All  figures  x 400  unless  otherwise  stated. 

Fig.  1.  Velamisporites  sp.  A.  x 1000.  Fair  66.2,  19.5,  102.2 

Fig.  2.  Grandispora  velata  (Eisenack)  Playford  1971.  Fair  37.64,  14.3,  100.7 

Figs.  3, 4.  Grandispora  protea  (Naumova)  Moreau-Benoit  1980.  3,  detail  of  distal  sculpture  x 1000.  4,  x400. 
Fair  37.71,49.5,  108.1 

Fig.  5.  Grandispora  Inaumovii  (Kedo)  McGregor  1973.  Fair  37.80,  12.5,  93.2 

Figs.  6,  7.  Rhabdosporites  sp.  A.  6,  x400.  7,  x 1000:  detail  showing  three  wall  layers.  Fair  37.71,  44.4,  95.1 
Figs.  8-10.  Rhabdosporites  langii  (Eisenack)  comb.  nov.  8,  ‘ parvulus ’ type  Fair  37.62,  19.4,  94.60.  9,  7 angii' 
type  Fair  37.80,  40.2,  105.5.  10,  ‘ parvulus ’ type,  compare  with  Geminospora.  Fair  37.82,  40.2,  105.5 


PLATE  32 


MARSHALL  and  ALLEN,  Devonian  miospores 


294 


PALAEONTOLOGY,  VOLUME  25 


laevigate,  often  indistinct,  closely  appressed  to  the  exoexine.  Exoexine  commonly  infrapunctate,  1-5-4  /xm  thick, 
with  an  interradial  maximum.  Perine  wrinkled  (especially  on  the  contact  areas,  where  it  forms  muroid  folds) 
0-25-0-5  (im  thick,  sparsely  sculptured  with  parallel  sided,  flat-topped  rods  (up  to  1 /xm  high)  mixed  with  grana 
and  coni  (up  to  0-5  /xm  high). 

Dimensions  (seven  specimens).  Maximum  perine  diameter  61-76  /xm  (mean  66  /xm),  maximum  exoexine  diameter 
46-67  /xm  (mean  55  /xm).  Separation  between  these  two  layers  5-17  /xm  (mean  1 1 /xm). 

Lithostratigraphic  range.  Ward  Hill  to  Bu  Ness  Groups  (lb  to  4a);  samples  31,  37,  and  66. 

Remarks.  Many  of  the  described  species  of  Devonian  Perotrilites  resemble  the  Fair  Isle  material. 
However,  all  differ  in  size  and  minor  sculptural  details.  Perotrilites  pannosus  Allen  1965  has  a thick 
perine  with  more  folds,  and  coni  with  bifurcating  tips.  P.  conatus  Richardson  1965  has  a denser 
sculpture  of  cones  and  a different  ratio  of  wall-layer  dimensions.  P.  aculeatus  Owens  1971  has  a 
thinner  perine  and  a sculpture  of  cones.  P.  selectus  (Arkhangelskaya)  McGregor  and  Camfield  1976 
is  larger  and  has  cones. 


Genus  rhabdosporites  Richardson  emend. 

Type  species.  Rhabdosporites  langii  (Eisenack)  Richardson  1 960. 

Emended  diagnosis.  Miospores  radial,  trilete,  amb  circular  to  triangular.  Camerate,  with  separate 
exoexine  and  intexine  attached  at  the  proximal  pole.  Exoexine  with  or  without  limbus,  sculptured 
with  low  rods,  coni,  and  grana.  Intexine  laevigate. 

Discussion.  In  1960  the  genus  Rhabdosporites  was  erected  to  accommodate  certain  distinctive 
pseudosaccate  miospores  previously  described  by  Eisenack  (1944),  Lang  (1925),  and  others.  Its  main 
distinguishing  features  according  to  Richardson  (1960)  were  the  proximally  attached  non-limbate 
bladder  and  the  evenly  distributed  over-all  sculpture  of  closely  packed  rods  with  parallel  sides  and  flat 
tops.  Subsequently,  the  genus,  and  particularly  the  type  species  Rhabdosporites  langii,  has  proved 
to  be  one  of  the  most  common  Devonian  miospores,  showing  a wide  stratigraphic  range  and  geo- 
graphic distribution.  Later  workers,  however,  have  not  kept  to  the  original  diagnosis,  placing  in 
Rhabdosporites  limbate  miospores  as  well  as  those  with  sculptural  elements  containing  grana  and 
coni.  There  have  been  some  good  arguments  for  this;  Owens  (1971)  pointed  out  a discernible  limbus 
in  the  illustration  of  the  type  species,  and  Lele  and  Streel  (1969,  p.  102)  commented  on  some  well- 
preserved  specimens  from  the  same  horizons  as  Richardson’s  material,  with  coni  often  as  a dominant 
element,  as  well  as  rods  with  truncated  tips.  It  would  seem  sensible  therefore  to  widen  the  scope  of  the 
genus  to  include  similar  spores,  with  both  a limbus  and  a more  variable  ornament  of  reduced  size. 
Camerate  spores  with  coarser  sculpture  are  included  in  other  genera  (e.g.  Grandispora).  In  widening 
the  generic  concept  of  Rhabdosporites,  we  are  aware  of  the  close  similarity  between  small  specimens 
of  Rhabdosporites  and  Geminospora  (Balme)  Owens  1971,  which  has  been  mentioned  by  previous 
authors  (e.g.  Lele  and  Streel  1969,  p.  103).  Geminospora  is  typified  by  a thin-walled  intexine 
either  closely  appressed  to,  or  showing  a variable  degree  of  separation  from,  a sculptured  exoexine 
with  a thickened  distal  surface.  This  organization  is  difficult  to  distinguish  (without  recourse  to 
microtome  sections)  from  a limbate  pseudosaccate  miospore  (see  PI.  32,  fig.  8)  with  a high 
intexine -exoexine  cavity  ratio  as  seen  in  small  Rhabdosporites.  Various  indirect  criteria  were  used  in 
attempts  to  distinguish  the  forms,  based  on  the  presence  of  the  thickened  distal  surface  which  is  a 
characteristic  of  the  Geminospora  group.  It  might  be  expected  that  this  would  hold  the  miospore  rigid, 
with  the  intexine  and  proximal  exoexine  positioned  in  the  shaped  distal  exoexine  ‘cup’,  thus  limiting 
the  eccentricity  between  the  two  bodies  when  compressed.  The  thickened  distal  exoexine  would  also 
deflect  and  restrict  exoexine  folding,  to  give  different  folding  characteristics  on  the  proximal  and 
distal  surfaces.  Further  evidence  for  the  similarity  (or  inability  to  easily  distinguish)  between 
Rhabdosporites  and  Geminospora  comes  from  the  study  of  in  situ  spores  (see  Allen  1980),  where 
spores  assignable  to  Rhabdosporites  and  Geminospora  are  recorded  from  closely  related  progymno- 
sperms. 


MARSHALL  AND  ALLEN:  DEVONIAN  MIOSPORES 


295 


Rhabdosporites  langii  (Eisenack)  comb.  nov. 

Plate  32,  figs.  8-10 

1925  Type  B Lang,  p.  256,  pi.  1,  figs.  3-6. 

1944  Triletes  langii  Eisenack,  p.  112,  pi.  12,  fig.  4. 

1960  Rhabdosporites  langii  (Eisenack)  Richardson,  p.  54,  pi.  14,  figs.  8,  9. 

1963  Rhabdosporites  firmus  Guennel,  p.  256,  fig.  12. 

1965  Rhabdosporites  parvulus  Richardson,  p.  588,  pi.  93,  figs.  5-7. 

1971  Rhabdosporites  micropaxillus  Owens,  p.  49,  pi.  15,  figs.  3-7. 

1973  Rhabdosporites  sp.  Hamid,  p.  202,  pi.  10,  no.  1. 

For  additional  synonymies,  see  Moreau-Benoit,  1980,  pp.  29  and  30. 

Dimensions  (one  hundred  and  sixty-five  specimens).  Maximum  exoexine  diameter  45-166  (mean  90  /un), 
maximum  intexine  diameter  39-146  ^m  (mean  65  pm). 

Lithostratigraphic  range.  Ward  Hill  to  Bu  Ness  Group  (lb  to  4a)— samples  Fair  8,  24,  28,  31,  33,  36,  37,  64,  66, 
and  67. 

Remarks.  Rhabdosporites  parvulus  Richardson  1965  was  associated  with  R.  langii  in  all  the  samples 
studied,  and  attempts  were  made  to  discriminate  between  the  two  species  as  did  Richardson  (1965, 
p.  588),  using  a graphical  plot  of  exoexine  and  intexine  diameters.  A graph  of  the  data  from  Fair  Isle 
(text-fig.  5)  shows  no  obvious  separation  into  two  populations,  but  a gradual  change  in  the  ratio  of 


160 


140 


El  20 
1 

wlOO 


40 


40  60  80  100  120  140  160  180 

DIAMETER  OF  EXOEXINE  (pm  ) 


text-fig.  5.  Size  variation  in  Rhabdosporites  langii  from  Fair  Isle  compared  with  published 
data  from  the  Orcadian  Basin.  Reduced  major  axis  lines  are  seen  to  be  closely  comparable 
between  Fair  Isle  and  a combined  R.  langii  and  R.  parvulus  population  from  the  Orcadian 
Basin.  Separate  R.M.A.  lines  for  R.  langii  Richardson  1960  and  R.  parvulus  Richardson  1965 
only  apply  to  parts  of  the  population  which  is  clearly  seen  to  be  continuous.  The  boxes  show 
the  limits  of  the  type-species  population,  and  the  stars  show  the  holotype  position.  Larger 
points  refer  to  two  or  three  individuals. 


296 


PALAEONTOLOGY,  VOLUME  25 


text-fig.  6.  Frequency  plots  for  Rhabdosporites  exoexine  diameters. 


the  exoexine  and  intexine  diameters  as  the  spore  size  increases.  Further,  the  data  from  Richardson’s 
graph  were  replotted  as  a single  population,  and  reduced  major  axis  lines  calculated  for  the  combined 
data.  It  can  be  seen  from  this  result  that  there  is  no  obvious  difference  between  regression  lines  for  [ 
combined  populations  of  R.  langii  and  R.  parvulus  and  similar  lines  calculated  for  the  separate  species 
as  defined  by  Richardson.  A regression  line  calculated  for  the  Fair  Isle  population  shows  a strong  j 
similarity  to  the  combined  Richardson  plot.  A frequency  distribution  plot  of  the  exoexine  diameters 
(see  text-fig.  6)  gives  a smooth  population  curve  and  no  sign  of  bimodality,  which  might  be  expected  if  i 
two  distinct  species  were  present.  The  stratigraphic  range  chart  of  Richardson  (1965,  opposite  p.  590)  j 
shows  a disjunct  appearance  for  R.  parvulus  only  in  the  Eday  Group,  whilst  the  text  (p.  588)  does 
indicate  its  appearance  further  down  the  sequence,  such  that  it  is  coincident  with  R.  langii.  \ 
Rhabdosporites  has  also  been  identified  as  an  in  situ  miospore  in  both  Tetraxylopteris  schmidtii  (Beck) 
Bonamo  and  Banks  1967  and  in  Rellimia  thompsonii  Leclercq  and  Bonamo  1973  (in  Leclercq  and 
Bonamo  1971).  These  occurrences  also  provide  further  information  on  the  similarity  between  i 
Rhabdosporites  langii  and  R.  parvulus , as  both  of  these  descriptions  also  comment  on  the  presence  of 
miospores  with  a size  range  including  both  R.  langii  and  R.  pavulus  in  the  same  sporangium.  These 
authors  also  consider  that  the  differences  described  by  Richardson  are  ontogenetic,  and  as  miospores  S 
matured  inside  a sporangium,  the  exoexine  expanded  more  rapidly  than  the  intexine,  showing  an 
increasing  ratio  between  the  two  wall  diameters  as  well  as  increased  sculpture  size.  It  is  suggested  that 
the  continua  seen  in  the  dispersed  species  can  be  attributed  to  this  ontogenetic  variation,  and  that 
with  the  size  distribution  data  collected  for  these  in  situ  miospores  to  confirm  this  similarity  there  is 
now  no  reason  to  consider  langii  and  parvulus  as  separate  species.  If  the  enlarged  concept  of  the  genus 
to  include  limbate  spores  with  sculptural  elements  of  grana  and  coni  as  well  as  rods  is  accepted,  then 
R.  micropaxillus  Owens  1 97 1 is  intermediate  between  R.  langii  and  R.  parvulus  and  can  be  regarded  as 
synonymous;  as  are  R.  firmus  Guennel  1963  and  Rhabdosporites  sp.  Hamid  1973,  which  similarly 
differ  only  in  the  presence  of  grana  and  coni  amongst  the  sculptural  elements. 


MARSHALL  AND  ALLEN:  DEVONIAN  MIOSPORES 


297 


Rhabdosporites  sp.  A 
Plate  32,  figs.  6,  7 

Description.  Miospores  trilete;  amb  circular;  exine  three-layered.  Suturae  length  equal  to  inner  body  radius, 
accompanied  by  simple  labra  (each  1 /xm  wide).  The  two  inner  wall  layers  laevigate,  homogeneous;  the  outer  wall 
infrapunctate,  sculptured  with  coni,  rounded-tipped  rods,  and  grana  (0-25-1  high).  The  three  walls  are 
variable  in  thickness  and  carry  folds  which  show  them  to  be  separated  except  on  the  proximal  surface.  Outer  wall 
appears  limbate. 

Dimensions  (three  specimens).  Maximum  diameters:  outer  wall  90  /xin,  101  /urn,  96  jim;  middle  wall  74  (im,  67  /im, 
80  ^m;  inner  wall  58  jum,  43  ^m,  61  ^m.  Optically  discernible  wall  thicknesses:  outer  wall  2-0  /u.m,  2-5  /xm,  0-5  /xm; 
middle  wall  0-25  /xm,  TO  /xm,  0-5  /xm;  inner  wall  2-0  /xm,  1-5  /xm,  2-0  /im. 

Lithostradgraphic  range.  Observatory  Group  (2c);  sample  Fair  37. 

Remarks.  Rhabdosporites  sp.  A of  Richardson  1965  is  similar  in  having  three  layers,  and  differs  only 
on  size  and  minor  sculptural  features.  ICalyptosporites  sp.  A Mortimer  and  Chaloner  1972  appears  to 
be  conspecific,  and  these  authors  also  suggest  a possible  further  synonymy  with  a miospore  figured  by 
Ozolin'a  (1960a,  pi.  1,  fig.  29)  and  identified  as  Archaeozonotriletes  micromanifesus  var.  minor 
Naumova.  Hemer  and  Nygreen  (1967)  illustrate  a triwalled  miospore  as  Rhabdosporites  sp.  but  as  no 
description  was  provided,  further  comparison  is  impossible.  Massa  and  Moreau-Benoit  (1976) 
record  the  presence  of  Rhabdosporites  sp.  A Richardson,  from  their  palynozones  6 and  7,  which  are 
dated  as  late  Givetian  and  early  Frasnian. 

Similar  in  situ  miospores  have  been  recovered  by  Tetraxylopteris  schmidtii  Bonamo  and  Banks 
1967,  where  they  occurred  in  subordinate  numbers  to  specimens  similar  to  Rhabdosporites  langii  and 
R.  parvulus.  Bonamo  and  Banks  (1967)  included  a personal  communication  from  Richardson,  which 
expressed  the  opinion  that  the  extra  layer  could  have  arisen  by  a splitting  of  one  of  the  two  wall  layers 
of  Rhabdosporites.  The  middle  wall  of  the  Fair  Isle  specimens  appears  to  be  a more  definite  layer  than 
in  similar  miospores  illustrated  from  T.  schmidtii.  A possible  explanation  for  this  difference  is  that  the 
third  layer  may  be  a teratological  feature,  and  a possible  way  of  showing  this,  as  opposed  to  the  wall 
splitting  hypothesis,  is  to  sum  the  various  thicknesses  of  the  wall  layers.  It  might  be  expected  that  any 
wall  splitting  would  provide  a total  equal  to  the  mean  intexine  plus  exoexine  value  of  a normal 
Rhabdosporites.  Unfortunately  the  data  are  not  available  to  show  this,  as  no  wall-thickness 
measurements  are  provided  for  the  Cromarty  populations  of  these  spores.  The  presence  of  a limbate 
wall  feature  would  also  create  difficulties  and  uncertainty  in  measurement. 

INCERTAE  SEDIS 

Genus  hystricosporites  McGregor  1 960 
Type  species.  Hystricosporites  delectabilis  McGregor  1960. 

Hystricosporites  cf.  corystus  Richardson  1960 
Plate  33,  figs.  7-9 

Description.  Miospore  trilete;  amb  circular  to  triangular,  ovate  in  lateral  compression.  Suturae  accompanied  by 
prominent  labra  (19-36  /xm  high,  56-116  /xm  in  length  as  seen  in  lateral  compression).  Exine  two-layered, 
intexine  laevigate,  closely  appressed  to  the  exoexine.  Exoexine  with  paired  radial  muri  on  the  contact  areas 
(length  20-28  /xm,  width  4-7  /xm),  proximo-equatorially  and  distally  sculptured  with  5-14  grapnel- tipped  spines 
(20-52  /xm  long). 

Dimensions.  Maximum  equatorial  diameter  60-130  /xm  (mean  96  /xm),  eleven  specimens  measured.  Maximum 
height  (excluding  labra)  in  lateral  compression  65-122  pm,  six  specimens  measured. 

Lithostradgraphic  range.  Ward  Hill  to  Bu  Ness  Groups  (lb  to  4d);  samples  Fair  8,  24,  31,  33,  36,  37,  64,  66, 
and  67. 


298 


PALAEONTOLOGY,  VOLUME  25 


Remarks.  The  Fair  Isle  population  is  very  similar  to  Hystricosporites  corystus  Richardson  1962, 
differing  only  in  gross  size.  Richardson’s  Hystricosporites  cf.  corystus  should  not  be  confused  with  the 
Fair  Isle  miospores. 

The  proximal  radial  muri  and  intexine  are  seen  only  in  overmacerated  specimens.  Frequently  only 
the  bases  of  the  spines  are  complete,  but  when  intact  the  grapnel-tips  are  of  the  laterally  extended 
form  (see  Owens  1971,  p.  87).  Difficulties  were  encountered  in  relating  spores  with  intact  spines  to 
those  showing  proximal  lip  and  intexine  detail,  because  the  high  maceration  levels  needed  to  clear  the 
miospores  resulted  in  severe  erosion  of  the  grapnel-tips. 

Genus  ancyrospora  Richardson  emend.  Richardson  1962 
Type  species.  Ancyrospora  grandispinosa  Richardson  1960. 

Ancyrospora  ancyrea  (Eisenack)  Richardson  1 962 
Plate  33,  figs.  3,  10 

Dimensions  (seventeen  specimens).  Maximum  equatorial  diameter  81-165  /un  (mean  120  (xm). 

Lithostratigraphic  range.  Ward  Hill  to  Bu  Ness  Groups  (lb  to  4a),  samples  8,  24,  28,  31,  33,  36,  37,  64,  66, 
and  67. 

Remarks.  Placed  in  Ancyrospora  ancyrea  and  not  A.  ancyrea  var.  ancyrea  Richardson  1962  because  it 
does  not  show  the  wide  flange  development  which  characterizes  the  variety.  It  conforms  to  text-fig.  5 
(p.  178)  of  Richardson  (1962),  and  is  morphologically  average  for  the  three  described  varieties. 

The  same  maceration  problems  were  encountered  in  studying  this  species  as  with  Hystricosporites 
cf.  corystus , in  that  it  was  rare  to  find  completely  cleared  spores  in  which  the  grapnel-tipped  spines 
were  preserved. 


Ancyrospora  ancyrea  cf.  var.  brevispinosa  Richardson  1962 
Plate  33,  figs.  4-6 

Description.  Miospores  trilete;  amb  triangular  to  subtriangular.  Suturae  accompanied  by  labra  (1  ^m  high). 
Exine  two-layered,  intexine  laevigate,  thin,  closely  appressed  to  the  exoexine.  Exoexine  shagreenate,  with  a dark 
halo,  or  triangular  darkening  in  the  proximal  polar  area,  contact  areas  often  with  sinuous  folds.  Proximo- 
equatorially  and  distally  sculptured  with  grapnel-tipped  spines  (1-10  /un  long),  the  tips  1 -2  across  have  little 
discernible  detail  (PI.  33,  fig.  6).  Flange  development  variable  (0-12  ^ m wide)  with  some  very  narrow  forms 
(PI.  33,  fig.  5)  and  others  interradially. 

Dimensions  (one  hundred  and  fifty  specimens).  Maximum  equatorial  diameter  39-95  ^m  (mean  65  /urn). 
Lithostratigraphic  range.  Ward  Hill  to  Bu  Ness  Group  (lb  to  4a);  samples  8,  24,  28,  31,  33,  36,  37, 64,  66,  and  67. 


EXPLANATION  OF  PLATE  33 
All  figures  x 400  unless  otherwise  stated. 

Fig.  1.  Geminospora  sp.  B.  Fair  31.5,  6.3,  101.1 
Fig.  2.  Geminospora  sp.  A.  Fair  31.5,  36.6,  101.7 

Figs.  3,  10.  Ancyrospora  ancyrea  (Eisenack)  Richardson  1962.  3,  Fair  37.64,  29.3,  107.1.  10,  detail  of  spine 
process,  x 1000. 

Figs.  4-6.  Ancyrospora  ancyrea  cf.  var.  brevispinosa  Richardson  1962.  4,  wide  flange  form.  Fair  8.1,  11.4, 106.1. 

5,  narrow  flange  form.  Fair  37.64,  14.4,  96.9.  6,  detail  of  spine  process  from  5,  x 1800. 

Figs.  7-9.  Hystricosporites  cf.  corystus  Richardson  1960.  7,  detail  of  spine  process  from  9.  8,  well  oxidized 
specimen.  Fair  67.30,  100.6,  33.4.  9,  dark  specimen,  lateral  compression.  Fair  37.63,  9.1,  96.9. 

Fig.  11.  Geminospora  tuberculata  (Kedo)  Allen  1965.  x 1000.  Fair  37.64,  22.4,  99.2 
Fig.  12.  Geminospora  svalbardiae  (Vigran)  Allen  1965.  x 1000.  Fair  28.20,  50.2,  1 14.3 


PLATE  33 


300 


PALAEONTOLOGY,  VOLUME  25 


Remarks.  This  spore  is  closely  similar  to  Ancyrospora  ancyrea  var.  brevispinosa  Richardson  1962 
except  in  gross  diameter,  which  seems  to  reflect  a genuine  difference,  as  opposed  to  modifications  in 
the  population  based  in  sedimentary  sorting  processes.  The  evidence  for  this  (see  text-fig.  7)  is  that 
whilst  the  size  frequency  distributions  for  Ancyrospora  ancyrea  from  both  Fair  Isle  and  the  Orcadian 
Basin  (see  Richardson  1962,  p.  185)  clearly  parallel  each  other,  those  for  var.  brevispinosa  (this  paper) 
are  quite  different.  It  is  postulated  on  a simple  model  that  any  sorting  process  which  could  modify  the 
size-frequency  distributions  of  the  variety  brevispinosa  and  cf.  brevispinosa  to  this  extent,  would  also 
show  a difference  in  the  Ancyrospora  ancyrea  populations,  which  is  not  so. 


2 50 


_A. ancyrea  var.  brevispinosa^ 
type  pop.  Rich.  ’62 


N>  “ 

8 5 


70  90 

MIOSPORE  DIAMETER  (pm) 


text-fig.  7.  Frequency  plot  for  Ancyrospora  ancyrea  cf.  var.  brevispinosa.  There  is  a difference  in 
population  size  range  compared  with  that  of  Richardson  1 962,  which  is  not  reflected  in  A . ancyrea. 


The  much-reduced  flange  on  the  exoexine  (see  PI.  33,  fig.  5)  produces  some  problems  in  using  the 
genus  Ancyrospora,  which,  as  defined  by  Richardson  (1962)  in  his  emended  diagnosis,  has  an 
extended  equatorial  flange  of  pseudoflange.  The  evidence  from  the  Fair  Isle  sequence  suggests  that  a 
continua  exists  from  obviously  flanged  miospores  to  almost  flangeless  ones,  and  therefore  in  part 
outside  the  generic  diagnosis.  The  presence  of  sinuous  surface  folds  on  the  exoexine  and  the 
occasional  appearance  of  a membranous  top  exine  layer,  suggests  the  possibility  that  a third  layer  is 
developed,  similar  to  A.  fallacia  Urban  (1970),  and  frequently  lost  on  oxidation.  Another  point  of 
interest  is  a possible  relationship  with  Perotrilites  bifur catus  Richardson  1962,  which  could  be 
interpreted  as  an  overmacerated  Ancyrospora  with  the  spalling  off  of  a third  outer  exine  layer 
normally  closely  appressed  to  the  exoexine. 

geminospora  (Balme)  Owens  1971 
Type  species.  Geminospora  lemur ata  Balme  1962 

Discussion.  As  well  as  the  difficulties  discussed  earlier  in  separating  Rhabdosporites  from  Gemino- 
spora, there  is  an  apparent  morphological  transition  series  between  Geminospora  and  Grandispora  in 
our  material.  The  problem  has  previously  been  described  in  other  assemblages  by  Neves  and  Owens 
(1966)  and  Playford  (1971).  Where  one  draws  the  dividing  line  between  truly  camerate  miospores 
such  as  Grandispora  which  show  a clear  separation  of  intexine  and  exoexine  (sensu  Playford  1976 
and  not  McGregor  1973),  and  Geminospora  which  has  both  widely  separated  and  closely  appressed 


MARSHALL  AND  ALLEN:  DEVONIAN  MIOSPORES 


301 


wall  layers  in  the  same  population,  is  at  present  arbitrary.  A further  difficulty  is  that  increased 
separation  of  wall  layers  may  result  from  the  oxidative  processes  used  in  clearing  the  highly 
carbonized  Fair  Isle  miospore  assemblages. 

Geminospora  tuberculata  (Kedo)  Allen  1 965 
Plate  33,  fig.  1 1 

Dimensions  (eleven  specimens).  Maximum  equatorial  diameter  42-67  /xm  (mean  56  /am). 

Lithostradgraphic  range.  Ward  Hill  to  Bu  Ness  Groups  (lb  to  4a);  samples  Fair  24, 28,  31,  33,  36,  37,  66,  and  67. 

Remarks.  The  Fair  Isle  material  differs  only  from  that  described  by  Allen  (1965)  in  having  slightly 
smaller  sculptural  elements  and  greater  variation  of  haptotypic  features.  The  emendation  of 
Archaeozonotrilites  tuberculatus  Kedo  by  Allen  (1965)  to  accommodate  the  Spitsbergen  population 
of  Geminospora  is  open  to  some  modification  as  more  Soviet  miospores  of  closely  related  species  have 
now  been  described,  into  which  parts  of  the  population  could  be  accommodated.  We  believe  that 
there  are  problems  of  synonymy  between  Geminospora  and  Archaeozonotriletes  (sensu  Naumova) 
and  a thorough  separate  revision  is  preferable  to  minor  amendments  made  here. 

Geminospora  svalbardiae  (Vigran)  Allen  1965 
Plate  33,  fig.  12 

Dimensions  (nine  specimens).  Maximum  equatorial  diameter  47-72  (im  (mean  58  jam). 

Lithostradgraphic  range.  Ward  Hill  to  Bu  Ness  Groups  (lb  to  4a);  samples  Fair  8,  31,  33,  36,  37,  and  66. 

Remarks.  The  population  described  by  Allen  (1965)  is  very  similar  to  the  Fair  Isle  material,  differing 
only  in  being  of  larger  over-all  size  and  having  a more  reduced  sculpture.  Possible  synonymies  include 
Geminospora  maculata  Taugourdeau-Lantz  1967  and  Geminospora  plicata  Clayton  and  Graham 
1 974  (nomen  illeg.  —junior  homonym  for  G.  plicata  Owens  1971).  Both  these  forms,  as  described,  are 
similar,  and  differ  only  in  minor  features  of  size  and  type  of  sculpture. 

Geminospora  sp.  A 
Plate  33,  fig.  2 

Description.  Miospore  trilete;  amb  circular  to  triangular.  Suturae  straight  and  extending  to  the  intexine  margin. 
Exine  two-layered;  intexine  laevigate,  thin  with  conspicuous  folds,  and  of  variable  separation  from  the  exoexine. 
Exoexine  5 /xm  thick,  with  a sculpture  of  coni  (2  jam  high),  some  of  which  are  arcuate  or  biform,  typical 
height  2 jam. 

Dimensions  (one  specimen).  Maximum  equatorial  diameter  70  /x m. 

Lithostradgraphic  range.  Observatory  Group  (2c);  sample  Fair  37. 

Remarks.  Similar  miospores  have  been  described  from  the  Soviet  Union  as  Archaeozonotriletes 
visendus  Chibrikova,  A.  visendus  var.  echinatus  Chibrikova,  A.  egregius  Naumova,  A.  tuberculatus 
var.  aculeatus  Raskatova,  A.  acutus  Raskatova,  A.  pensus  Kedo,  A.  lasius  var.  minor  Naumova. 
Unfortunately  its  single  occurrence  in  the  Fair  Isle  sequence  limits  its  value  at  present. 

Geminospora  sp.  B 
Plate  33,  fig.  1 

Description.  Miospores  trilete;  amb  roughly  triangular.  Suturae  simple  or  accompanied  by  thin  labra  commonly 
1 -2  jam  high  (maximum  5 /am).  Exine  two-layered  with  exoexine  and  intexine  layers  separate,  attached  only  at  the 
proximal  pole.  Intexine  laevigate,  thin,  showing  a variation  in  degree  of  separation.  Exoexine  2-5—1 4 /am  thick 
(mean  6-5  /xm)  with  individuals  showing  an  interradial  maximum  (up  to  2-5  /am  thicker).  Exoexine  sculpture 


302 


PALAEONTOLOGY,  VOLUME  25 


(except  for  contact  areas)  with  cones  and  fine  tapering  spines  1 -5  /xm  long  (mean  2-5  jim),  denser  and  of  greater 
length  distally  than  equatorially  or  proximo-equatorially. 

Dimensions  (twenty-seven  specimens).  Maximum  exoexine  diameter  65-140  fxm  (mean  89  /xm). 
Lithostratigraphic  range.  Ward  Hill  and  Observatory  Groups  (lb  to  2c);  samples.  Fair  37  and  66. 

Remarks.  Archaeozonotriletes  comptus  Naumova  (1953),  A.  comptus  var.  expletivus  Chibrikova 
(1959)  (see  also  Raskatova  1969  and  Chibrikova  1962,  1977),  A.  tuber culatus  Kedo  var.  minor  Kedo 
(1976),  and  A.  tuber  culatus  Kedo  var.  triangulatus  Kedo  1976  are  all  miospores  of  a similar 
construction,  sculpture,  and  size.  However,  their  descriptions  are  not  detailed  enough  to  confidently 
place  Geminospora  sp.  B into  any  one  of  these  species. 


SEQUENCE  AND  STRATIGRAPHIC  SIGNIFICANCE  OF  THE 
MIOSPORE  ASSEMBLAGES 

The  number  of  miospore-bearing  horizons  has  proved  to  be  small,  but  fortunately  of  wide 
distribution.  This  is  shown  for  selected  miospore  taxa  in  text-fig.  8.  There  is  no  major  palynostrati- 
graphic  change  in  the  succession  until  the  Bu  Ness  Group,  which  is  characterized  by  the  incoming  of 
Aneurospora  greggsii  and  Convolutispora  disparalis. 

The  general  problems  of  Devonian  palynostratigraphy  have  been  discussed  at  length  by  several 
authors  (Owens  and  Richardson  1972;  Richardson  1974;  and  McGregor  1979a)  so  will  only  be  briefly 
alluded  to  in  this  stratigraphical  synthesis.  The  total  stratigraphical  ranges  for  some  of  the  taxa  from 
Fair  Isle  are  plotted  in  text-fig.  9 and  show  long  ranging  distributions.  This  type  of  compilation 
suffers  from  the  lack  of  miospore  assemblages  securely  dated  by  independent  means  (notably 
Ammonoids,  Conodonts),  which  can  be  compared  accurately  with  the  type  sections  of  the  traditional 
stages.  The  data  suggests  a Givetian  age  for  the  Fair  Isle  assemblages  but  does  not  place  reliance  on 
the  restricted  ranges  of  several  poorly  known  species.  A more  precise  correlation  is  possible  by 
drawing  comparisons  with  selected  areas  containing  well-dated  assemblages. 

Comparisons  with  European  Assemblages  ( excluding  the  Soviet  Union  but  including  Spitsbergen) 

The  spore  sequences  described  by  Richardson  (1965)  for  the  Orcadian  Basin  are  now  geographically 
the  closest  described  Devonian  palynofloras  to  Fair  Isle.  The  abundance  in  Fair  Isle  of  Ancyrospora 
ancyrea  cf.  var.  brevispinosa  with  the  first  appearance  of  Grandispora  Inaumovii  suggest  possible 
equivalence  with  the  Eday  Beds.  An  interesting  comparison,  and  possibly  significant,  is  the  very  rare 
occurrences  of  Emphanisporites  spp.  in  both  Fair  Isle  and  the  Orcadian  Basin,  which  are  not  thought 
to  be  the  result  of  reworking  (Clayton  et  al.  1977).  A major  difference  between  the  Fair  Isle  beds  and 
the  Eday  Group  on  Orkney  is  the  complete  absence  of  Geminospora  spp.  in  the  latter,  compared  with 
its  relative  abundance  and  diversity  in  Fair  Isle.  It  could  be  argued  that  this  is  of  stratigraphic 
importance,  except  that  Geminospora  spp.  is  known  from  independently  dated  Givetian  beds  from 
over  a wide  area  including  Illinois  (Sanders  1968),  Spitsbergen  (Allen  1965),  and  southern  England 
(Mortimer  and  Chaloner  1972).  It  has  been  suggested  (Richardson  1967)  that  the  control  is 
ecological,  and  this  will  be  discussed  later. 

The  spore  assemblages  described  from  borehole  material  of  southern  England  by  Mortimer  and 
Chaloner  (1972)  compare  quite  closely,  but  add  little  to  the  specific  age  assignment  apart  from 
confirming  it  as  Givetian.  Detailed  comparisons  of  the  Fair  Isle  spore  assemblages  can  also  be  made 
with  the  sequences  described  by  Streel  and  associates  (reviewed  in  Streel  1972)  from  Belgium  and 
northern  France.  Streel  in  a discussion  of  marker  spores  for  the  Givetian-Frasnian  boundary 
considers  the  first  appearances  of  Aneurospora  greggsii,  Ancyrospora  langii,  and  Samarisporites 
hesperus  to  be  of  major  significance.  The  occurrence  of  Aneurospora  greggsii  in  the  Bu  Ness  Group 
does  suggest  that  this  higher  part  of  the  Fair  Isle  sequence  is  late  Givetian  in  age  and  close  to  the 
Givetian-Frasnian  boundary.  Streel  (1972)  also  gave  the  range  of  Grandispora  velata  and 
IRhabdosporites  langii  (closely  comparable  with  the  Fair  Isle  form)  from  the  Tournai  borehole  as 


MARSHALL  AND  ALLEN:  DEVONIAN  MIOSPORES 


1 


1' 


O W > O 


3 3 3 § 3 

2 5'  5'  & 2 


(/)(/)(/></) 


3 2. 


text-fig.  8.  Stratigraphical  distribution  of  selected  miospores  in  Fair  Isle. 


overlapping  the  range  of  Aneurospora  greggsii,  but  the  information  was  not  sufficient  to  suggest  that 
this  was  of  any  widespread  stratigraphic  value.  Loboziak  and  Streel  (1980),  in  a study  of  Givetian  and 
Frasnian  rocks  from  Boulonnais  (France),  considered  several  species  to  be  of  importance  in 
delimiting  beds  of  late  Givetian  age,  including  Chelinospora  concinna,  Ancyrospora  ancyrea  var. 
brevispinosa,  and  Convolutispora  disparalis,  all  found  in  the  Fair  Isle  section. 

In  Spitsbergen,  Allen  ( 1 967)  described  a zonal  scheme  of  three  assemblages  for  the  Devonian  rocks. 
The  Fair  Isle  sequence  has  many  similarities  to  the  triangulatus  assemblage  with  Geminospora 
tuberculata,  Cirratriradites  avius,  Convolutispora  disparalis,  Densosporites  devonicus,  Chelinospora 
concinna,  and  Grandispora  protea  in  common.  The  triangulatus  assemblage  was  assigned  to  the 


Sampless  sS  8S 


Gp.  Obs.  Gp. 


Hill  Gp. 


Convolutispora  disparalis 
Aneurospora  greggsii 
Samarisporites  orcadensis 
Grandispora?  naumovii 
-•  Velamisporites  sp.  A 

-•  Grandispora  velata 

"•  Samarisporites  mediconus 
Grandispora  protea 
"•  Geminospora  svalbardiae 
Geminospora  tuberculata 
— • Samarisporites  conannulatus 
Auroraspora  macromanifestus 
Retusotrilites  rotundus 
Calamospora  atava 
Hystricosporites  cf.  corystus 
Oensosporites  devonicus 
Cirratriradites  avius 

-+-+  Ancyrospora  ancyrea  cf.  var.  brevispinosa 
Ancyrospora  ancyrea 
Geminospoa  sp.  B 
Rhabdosporites  langii 
-+-•  Auroraspora  micromanifestus 


i 

I 

.3 

1 

1 


2 


304 


PALAEONTOLOGY,  VOLUME  25 


text-fig.  9.  Stratigraphical  distribution  of  selected  taxa  which  occur  on  Fair  Isle.  * Refers  to  comparison  record. 


Givetian  by  Allen  (1967)  although  the  lower  part  is  without  Densoporites  devonicus  and  is  considered 
to  be  of  possible  Eifelian  age  (McGregor  and  Camfield  1976). 

A detailed  study  has  been  made  by  Beju  (1972)  on  rocks  of  Devonian  age  from  the  Moesian 
platform  of  Romania.  The  Eifelian  subzone  (D2a)  has  the  long-ranging  Grandispora  protea  and 
Calamospora  atava  in  common  with  Fair  Isle.  The  Givetian  subzone  (D2b)  is  more  similar,  with 
Ancyrospora  ancyrea,  Rhabdosporites  langii,  Grandispora  velata,  Auroraspora  macromanifestus,  and 
Densosporites  devonicus  in  common.  These  last  three  are  considered  by  Beju  to  be  specifically 
characteristic  of  this  subzone.  The  Frasnian  zone  (D3)  contains  Geminospora  svalbardiae,  G. 
tuberculata,  Rhabdosporites  parvulus,  and  Auroraspora  micromanifestus. 

Comparisons  with  North  American  Assemblages 

The  Frasnian  Ghost  River  spore  assemblage  described  from  Alberta  by  McGregor  (1964)  compares 
with  the  Bu  Ness  Group  of  the  Fair  Isle  sequence  only  in  the  presence  of  Aneurospora  greggsii. 
McGregor  (in  McGregor  and  Uyeno  1972)  described  a sequence  of  spore  assemblages  from  the 
Canadian  Arctic  Archipelago,  and  of  these  the  closest  to  the  Fair  Isle  material  is  assemblage  D of  late 
Givetian  age,  characterized  by  the  appearance  of  Grandispora  protea  and  Chelinospora  concinna.  The 
disappearance  of  Densosporites  devonicus  is  not  held  to  be  of  major  significance  as  its  replacement  by 
D.  orcadensis  in  the  Orcadian  sequence  may  be  anomalous  in  view  of  the  apparent  synonymy  which 
exists  between  the  two  species  (see  McGregor  and  Camfield  1976;  McGregor  1979a  and  this  paper). 

Devonian  spores  described  by  Owens  (1971)  from  Middle  and  Upper  Devonian  rocks  of  the 
Canadian  Arctic  Archipelago,  suggest  that  broad  similarities  can  be  drawn,  but  his  work  is  largely 
based  on  new  species  from  a few  horizons,  and  is  of  limited  value  for  detailed  comparisons. 


MARSHALL  AND  ALLEN:  DEVONIAN  MIOSPORES 


305 


Sanders  (1968)  described  the  miospores  from  an  acritarch-dominated  assemblage  of  Givetian  age, 
with  Rhabdosporites  langii  (and  R.  parvulus),  Geminospora  tuberculata , and  examples  of  Ancyrospora 
spp.  and  Velamisporites  spp.  ( Perotrilites ). 

Comparisons  with  North  African  and  the  Middle-East  Assemblages 

Massa  and  Moreau-Benoit  (1976)  described  a series  of  miospore  zones  covering  the  Devonian  strata 
of  western  Libya,  and  limited  comparisons  can  be  drawn  with  the  Fair  Isle  succession.  Palynozones 
4 and  5 are  the  most  similar,  having  Auroraspora  micromanifestus,  Emphanisporites  rotatus, 
Rhabdosporites  parvulus  (R.  langii),  Grandispora  velata,  Geminospora  svalbardiae,  and  Samarisporites 
mediconus  in  common,  and  these  are  dated  as  Couvinian  and  lower  Givetian  respectively. 

Comparisons  with  miospore  assemblages  from  the  Russian  Platform 

General  comparisons.  Table  1 lists  some  miospore  taxa  from  the  Russian  Platform  together  with  their 
stratigraphical  ranges,  which  are  considered  to  be  broadly  comparable  with  taxa  from  Fair  Isle. 
Comparability  is  variable,  ranging  from  likely  direct  equivalence  to  part  synonymy  or  closely 
similar.  Some  species,  for  example  Hymenozonotriletes  protea,  can  be  referred  to  several  western 
species. 

In  the  compilation  of  Table  1 it  was  hoped  to  keep  the  stratigraphic  information  in  the  form  of  the 
local  units  used  over  the  Russian  Platform,  but  correlation  difficulties  made  it  impossible  to  construct 
a single  table  for  the  data,  so  that  a more  generalized  scheme  was  necessarily  produced.  The  placing 
of  the  Eifelian-Givetian  and  the  Givetian-Frasnian  boundaries  is  open  to  question,  with  some 
amendments  having  been  made  in  certain  areas.  Hymenozonotriletes  punctomonogrammos  and  H. 
monogrammos  (Arkhangelskaya  1974),  for  example,  are  found  in  the  Mosolovian  and  Chernoyarian 
beds  of  the  Central  and  Eastern  Russian  Platform,  and  are  part  of  the  polymorphos-monogrammos 
zone  which  has  many  taxa  similar  to  the  Fair  Isle  material,  but  is  placed  in  the  Eifelian  by 
Arkhangelskaya.  A similar  problem  is  seen  in  the  Givetian-Frasnian  beds  of  the  Baltic  region,  where 
the  Soviet  workers  draw  the  boundary  at  the  base  of  the  Gauja  formation  (e.g.  Ozolin'a  1963), 
whereas  recently  compiled  data  (Westoll  1979  and  pers.  comm.),  places  it  at  the  base  of  the  Snetogor, 
which  gives  this  succession  great  importance  as  being  continuous  across  the  Middle/Upper  Devonian 
boundary  and  containing  good  vertebrate  faunas.  Unfortunately  although  this  section  appears 
to  be  well  documented  (e.g.  Ozolin'a  1955,  1960a,  1960 b,  1961,  1963;  Kedo  1966)  the  miospore 
illustrations  and  descriptions  are  poor,  making  comparisons  difficult.  The  compilation  of  data 
presented  in  Table  1 shows  a Givetian  age  for  the  Fair  Isle  assemblages  but  it  is  impossible  to  give 
greater  precision  with  this  type  of  synthesis. 

Comparisons  with  zonal  schemes  from  the  Russian  Platform 

Several  zonal  schemes  have  been  published  which  cover  the  relevant  stratigraphic  interval,  and 
specific  comparisons  are  made  in  an  attempt  to  give  a more  detailed  correlation.  Kedo  (1966) 
produced  a zonal  scheme  for  the  western  part  of  the  Russian  Platform  and  although  there  are  some 
taxa  in  common,  it  is  very  difficult  to  make  a precise  correlation,  except  to  record  an  influx  of 
Archaeozonotriletes  (sensu  Naumova  = Geminospora  spp.)  in  the  upper  part  of  the  Givetian,  which 
compares  with  Fair  Isle,  but  contrasts  with  the  Orcadian  Givetian  which  does  not  contain 
Geminospora.  This  indicates  either  younger  Givetian  rocks  in  Fair  Isle  than  in  the  Orcadian  Basin,  or 
the  possibility  of  an  ecological  control  (see  below). 

Raskatova  (1969,  1974)  has  produced  a more  compatible  scheme  which  again  shows  an  influx 
of  Archaeozonotriletes  in  the  upper  Givetian  (zones  IV  to  VIII)  Vorbyevkian  to  Staryi’  oskolian. 
Hymenozonotriletes  punctomonogrammos  Arkhangelskaya  and  H.  monogrammos  Arkhangelskaya 
are  also  present,  but  occur  lower  down  in  the  uppermost  Eifelian  as  well  as  the  Givetian,  so  that  the 
age  assignment  is  not  clear  cut. 

From  the  same  horizons,  Arkhangelskaya  (1974)  has  also  recorded  these  two  species  of 
Hymenozonotriletes  as  well  as  taxonomic  equivalents  of  Rhabdosporites  langii,  Densosporites 
devonicus,  Grandispora  protea,  and  Samarisporites  spp.  These  were  compared  with  the  Weatherall 


306 


PALAEONTOLOGY,  VOLUME  25 


table  1.  Comparative  table  of  selected  Soviet  and  western  taxa  occurring  on  Fair  Isle.  Records  include  likely 
synonymies,  part  synonymies,  and  close  comparisons.  Sources  include  Ozolin'a  1955,  1960a,  19606,  1961,  1963, 
Kedo  1955,  1960,  1966,  1967,  Chibrikova  1959,  1962,  1972,  1977,  Chibrikova  and  Naumova  1974, 
Arkhangelskya  1963,  1974,  Naumova  1953,  Andreeva  1962,  Raskatova  1969,  1974,  and  Pokrovskaya  1966. 


WESTERN  TAXA 

SOVIET  COMPARISONS 

EIFELIAN 

GIVETIAN 

FRASNIAN 

Archaeotri letes  sincerus  Kedo 

• 

• 

Ancyrospora  ancyrea 

Archaeotriletes  splendidus  Kedo 

• 

Archaeotri letes  hamulus  Naumova 

• 

• 

Retusotri letes  verrucosus  (Naumova  in  litt.)  Kedo 

• 

Aneurospora  greggsii 

Re tusotri letes  punctatus  Chibrikova 

• 

• 

Archaeozonotriletes  nalivkinii  Naumova 

• 

• 

Cirratriradites  avius 

Hymenozonotri letes  punctomonogrammos  Arkhangelskaya 

• 

Cirratriradites  sp.  A 

Hymenozonotri letes  monogranmos  Arkhangelskaya 

• 

Densosporites  devonicus 

Hymenozonotri letes  meonacanthus  Naumova 

• 

• 

• 

Archaeozonotriletes  meonacanthus  Nomen  nudum 

• 

Geminospora  tuberculata 

Archaeozonotriletes  tuberculatus  Kedo 

• 

Archaeozonotriletes  plicata  (Naumova  in  litt.)  Kedo 

• 

Retusotri letes  parvimammatus  Kedo 

• 

• 

Geminospora  svalbardiae 

Archaeozonotriletes  notatus  Naumova 

• 

Archaeozonotriletes  rugosus  Naumova 

• 

• 

Archaeozonotriletes  visendus  Chibrikova 

• 

• 

Archaeozonotriletes  egregius  Naumova 

• 

• 

Archaeozonotriletes  tuberculatus  var. aculeatus  Raskatova 

• 

Geminospora  sp.  A 

Archaeozonotri letes  acutus  Raskatova 

• 

Archaeozonotriletes  visendus  var.  echinatus  Chibrikova 

• 

Archaeozonotriletes  pensus  Kedo 

• 

Archaeozonotriletes  lasius  var  minor  Naumova 

# 

Archaeozonotriletes  comptus  Naumova 

# 

Geminospora  sp.  B 

Archaeozonotriletes  comptus  var.  expletivus  Chibrikova 

• 

Archaeozonotriletes  tuberculatus  var.  minor  Kedo 

• 

Archaeozonotriletes  tuberculatus  var.  triangulatus  Kedo 

• 

Grandispora  naumovii 

Archaeozonotriletes  naumovii  Kedo 

• 

• 

Hymenozonotri letes  curticonicus  Kedo 
Hymenozonotri letes  ventosus  Kedo 

• 

Grandispcra  protea 

Hymenozonotri letes  proteus  Naumova 
Hymenozonotri letes  verus  Naumova 

• 

• 

Hymenozonotri letes  endemicus  Chibrikova 

• 

Hymenozonotri letes  echini formis  Naumova 

• 

• 

Hymenozonotri letes  longus  Arkhangelskaya 

• 

• 

Grandispora  velata 

Hymenozonotri letes  tener  var.  concinna  Chibrikova 

• 

Hymenozonotri letes  proteus  Naumova 

• 

• 

Grandispora  velata  (Richardson)  Playford 

• 

Hymenozonotri letes  facetus  Arkhangelskaya 

• 

• 

Hymenozonotri letes  polymorphus  (Naumova)  Kedo 

• 

• 

Rhabdosporites  langii 

Archaeozonotriletes  micromanifestus  Naumova 

• 

• 

Hymenozonotri letes  varius  Naumova 

• 

• 

Rhabdosporites  parvulus  Richardson 

• 

• 

Formation  from  Arctic  Canada  (Givetian),  the  Thurso  and  Eday  Beds  from  the  Orcadian  Basin 
(Givetian),  and  the  Achanarras  Fish  Bed  (Eifelian-Givetian).  The  recent  zonal  scheme  produced  by 
Chibrikova  and  Naumova  (1974)  is  not  very  detailed  with  regard  to  specific  taxa,  and  is  largely 
restricted  to  generic  distribution.  The  appearance  of  Archaeozonotriletes  is  considered  significant, 
and  is  used  as  a marker  for  the  Givetian  stage.  The  zonal  scheme  produced  by  Naumova  (1953)  is 
difficult  to  apply,  as  the  taxa  are  not  often  placed  in  a detailed  zonal  scheme,  but  in  broad  categories 


MARSHALL  AND  ALLEN:  DEVONIAN  MIOSPORES 


307 


such  as  Middle  or  Upper  Devonian.  The  Fair  Isle  assemblages  fit  best  into  the  Givetian  and  Frasnian 
groups. 

A compilation  of  these  occurrences  has  been  presented  by  Andreeva  (in  Pokrovskaya  1966)  and, 
as  indicated,  uses  Chibrikova’s  (1962)  and  Naumova’s  (1953)  work.  The  significant  occurrence  of 
Archaeozonotriletes  ( = Geminospora)  with  subordinate  Hymenozonotriletes  is  used  as  a marker  for 
the  Givetian  stage.  The  spores  recovered  by  Chibrikova  (1962)  from  the  Vorobyevkian  of  Western- 
Bashkiria  contain  several  species  which  are  similar  to  the  Fair  Isle  assemblages,  with  examples  such  as 
Archaeozonotriletes  comptus  var.  expletivus,  A.  meoncanthus,  A.  visendus,  and  A.  egregius. 


00 

3 

3 

*2. 

N 

o 

3 

3 

$ 

rt 

c 

D. 

3 $ 

a § 
^ S' 

,3 

Other 

fT 

CD 

O 

S' 

2 

z 

o 

sS 

o 

o 

3 

^ £ 

2 ^ 
a 

3 

O 

"3 

o 

2 

ft 

St 

a *p 

Fair 

24 

6 

14 

39 

2 

20 

19 

Bu  Ness  Group  Fair 

28 

4 

39 

3 

0-5 

30 

23-5 

Fair 

33 

2 

19 

15 

3 

53 

8 

Observatory  Group  Fair 

36 

8 

47 

11 

2 

13 

19 

Fair 

67 

5 

37 

2 

27 

27 

2 

Ward  Hill  Group  Fair 

66 

6 

33 

36 

- 

21 

1 

Fair 

64 

19 

35 

2 

5 

38 

1 

table  2.  Relative  proportions  of  major  miospore  types  from  Fair  Isle.  The  figures 
are  percentages  of  numbers  of  each  group  from  150  total  counts. 

Ecological  considerations 

An  ecological  control  on  the  distribution  of  some  Devonian  miospores  has  been  speculated  by  several 
authors  (e.g.  Streel  1967;  Richardson  1965,  1967,  1969),  and  the  example  cited  is  usually  that  of 
Geminospora  spp.  {Archaeozonotriletes  of  Soviet  palynologists).  These  miospores  are  thought  to  be 
dispersed  from  plants  growing  on  and  around  the  lower  flood  plain  and  marginal  marine  areas,  whilst 
miospores  from  the  bifurcate  and  pseudosaccate  zonate  groups  are  more  continental  in  aspect  and 
related  to  upper  flood  plain  and  possibly  lacustrine  deposits.  The  assemblages  from  Fair  Isle  are 
relevant  to  this  discussion  as  they  often  contain  abundant  Geminospora  spp.  (fluctuating  from  39%  to 
2%,  see  Table  2),  in  what  appears  to  be  a true  internal  basin  facies.  The  presence  of  Svalbardia  scotica 
(Chaloner  1972)  in  the  Bu  Ness  Group  is  also  significant,  because  other  species  are  known  to  contain 
in  situ  spores  of  Geminospora  and  Aneurospora  (see  Allen  1980),  suggesting  that  their  origin  is  local. 
The  occurrences  are  also  not  restricted  in  the  Fair  Isle  succession  (see  text-fig.  6),  being  found  in  the 
minor  argillite  horizons  (?lacustrine  deposits)  of  the  Ward  Hill  Group  (alluvial  fan  deposits)  as  well 
as  the  overbank  deposits  of  the  Bu  Ness  Group.  This  distribution  contrasts  strongly  with  that  of 
Richardson  ( 1 965, 1 967,  1 969),  who  noted  an  absence  of  Geminospora  from  the  lacustrine  and  fluvial 
deposits  from  the  Orcadian  basin  lying  to  the  south.  A possible  explanation  is  that  the  time  difference 
between  the  Orcadian  and  Fair  Isle  deposits  is  sufficient  to  allow  the  migration  of  an  ‘ Archaeozono- 
triletes ’ microflora  from  the  Russian  Platform,  where  it  occurs  in  lower  Givetian  deposits  (Raskatova 
1974)  contemporaneous  with  the  Orcadian  Basin.  A less  appealing  hypothesis  is  that  the  Fair  Isle 
lacustrine  and  fluvial  deposits  were  sustained  by  permanent  water  flow,  giving  more  equable 
conditions  than  those  in  the  Orcadian  Basin,  and  similar  to  conditions  seen  on  delta  tops  or  marine 
margins,  where  Geminospora  is  known  to  thrive  (see  Becker  et  al.  1974).  These  authors  describe  in 
detail  the  distribution  of  Aneurospora  greggsii,  and  suggest  it  is  not  restricted  by  ecological  factors  in 
their  environment  studied,  which  unfortunately  do  not  include  lacustrine  or  alluvial  fan  facies.  It 


308 


PALAEONTOLOGY,  VOLUME  25 


seems  that  this  conflict  between  an  edaphic  or  stratigraphic  control  will  not  be  resolved  without 
further  detailed  palaeoecological  studies  of  this  type. 

Acknowledgements.  J.  E.  A.  Marshall  is  indebted  to  the  Natural  Environment  Research  Council  for  the  receipt  of 
one  of  their  C.A.S.E.  studentships  and  also  to  the  Institute  of  Geological  Sciences  (Leeds)  for  the  use  of  their 
facilities.  The  support  of  Newcastle  University  is  also  acknowledged  for  the  latter  part  of  the  work.  Both  K.  C. 
Allen  and  J.  E.  A.  Marshall  acknowledge  the  use  of  research  facilities  and  support  of  the  Botany  Department, 
Bristol  University.  Useful  and  informative  discussions  have  also  been  of  great  assistance  from  the  following:  Dr. 
B.  Owens  (I.G.S.  Leeds),  Professor  T.  S.  Westoll,  F.R.S.  (Newcastle  University),  and  Dr.  W.  Mykura  (I.G.S. 
Edinburgh).  The  receipt  of  some  preliminary  samples  from  the  I.G.S.  Edinburgh  (courtesy  of  Mr.  P.  J.  Brand)  is 
also  gratefully  acknowledged.  The  help  of  the  warden  and  staff  at  the  Fair  Isle  Bird  Observatory  during  the  field- 
work is  also  acknowledged. 


REFERENCES 

AGASSIZ,  L.  1844-5.  Monographie  des  Poissons  fossiles  du  Vieux  Gres  Rouge  ou  Systeme  Devonien  ( Old  Red 
Sandstone)  des  lies  Britanniques  et  de  Russie.  Neuchatel. 
allen,  k.  c.  1965.  Lower  and  Middle  Devonian  spores  of  North  and  Central  Vestspitsbergen.  Palaeontology , 8, 
687-748. 

— 1967.  Spore  assemblages  and  their  stratigraphical  application  in  the  Lower  and  Middle  Devonian  of  North 
and  Central  Vestspitsbergen.  Ibid.  10,  280-297. 

— 1980.  A review  of  in  situ  late  Silurian  and  Devonian  spores.  Rev.  Palaeobot.  Palynol.  29,  253-270. 
andreeva,  e.  m.  1962.  Important  Middle  Devonian  spores  from  the  Kuznetz  Basin.  Contr.  Phytostratig.  Dev. 

Sed.  Altae-Sayan  Mountain  Region.  Trudy.  VSEGEI,  n.s.  70,  191-213.  [In  Russian.] 

Arkhangelskaya,  a.  d.  1963.  New  species  of  spores  from  Devonian  deposits  of  the  Russian  Platform.  Trudy 
Vnigri , 37,  18-30.  [In  Russian.] 

— 1974.  Spore  based  zonation  of  the  lower  part  of  the  Middle  Devonian  and  its  inter-regional  correlation  in 
central  and  eastern  European  regions  of  the  U.S.S.R.  In  Palynology  of  Proterophy te  and  Palaeophyte.  Proc. 
3rd  Int.  Palynol.  Conf.,  Novosibirsk,  1971.  56-59.  [In  Russian.] 

balme,  b.  e.  1962.  Upper  Devonian  (Frasnian)  spores  from  the  Carnarvon  Basin,  Western  Australia. 
Palaeobotanist,  9,  1-10. 

beck,  c.  b.  1957.  Tetraxylopteris  schmidtii  gen.  et  sp.  nov.,  a probable  pteridosperm  precursor  from  the 
Devonian  of  New  York.  Am  J.  Bot.  44,  350-367. 

becker,  G.,  bless,  m.  h.  m.,  streel,  m.  and  thorez,  j.  1974.  Palynology  and  ostracode  distribution  in  the  Upper 
Devonian  and  basal  Dinantian  of  Belgium  and  their  dependance  on  sedimentary  facies.  Meded.  Rijks.  geol. 
Dienst,  n.s.  25,  9-99. 

beju,  d.  1972.  Zoning  and  correlation  of  the  Palaeozoic  of  the  Moesian  Platform  on  the  basis  of 
palynoprotistological  assemblages.  Petrol,  si  Gaze,  23,  714-722.  [In  Romanian.] 
berry,  w.  1937.  Spores  from  the  Pennington  Coal,  Rhea  County,  Tennessee.  Amer.  Midi.  Nat.  18,  155-160. 
bharadwaj,  d.  c.  and  venkatachala,  b.  s.  1962.  Spore  assemblages  out  of  a Lower  Carboniferous  shale  from 
Spitzbergen.  Palaeobotanist,  10,  18-47. 

birks,  h.  j.  b.  1970.  Inwashed  pollen  spectra  at  Loch  Fada,  Isle  of  Skye.  New  Phytol.  69,  807-820. 
bonamo,  p.  m.  and  banks,  h.  p.  1967.  Tetraxylopteris  schmidtii.  its  fertile  parts  and  its  relationships  within  the 
Aneurophytales.  Am.  J.  Bot.  54,  755-768. 

brooks,  c.,  hart,  s.  R.  and  wendt,  I.  1972.  Realistic  use  of  two  error  regression  treatments  as  applied  to 
Rubidium-Strontium  data.  Rev.  Geophys.  Space  Phys.  10,  551-557. 
burmann,  g.  1969.  Inkchlungundmechanishe  deformation.  Zeitzh.  fur  angewandte  Geologic,  Bd.  15, 7,  355-363. 
butter  worth,  m.  a.  and  collaborators.  1964.  Densosporites  (Berry)  Potonie  and  Kremp  and  related  genera.  C.r. 
5iemeC0ng.  Strat.  Geol.  carbonif.  Paris  (1963),  1,  1049  1057. 

— and  williams,  r.  w.  1958.  The  small  spore  floras  of  coals  in  the  Limestone  Coal  Group  and  Upper 
Limestone  Group  of  the  Lower  Carboniferous  of  Scotland.  Trans.  R.  Soc.  Edinb.  63,  353-392. 

chaloner,  w.  G.  1967.  Spores  and  land  plant  evolution.  Rev.  Palaeobot.  Palynol.  1,  83-93. 

1972.  Devonian  plants  from  Fair  Isle,  Scotland.  Ibid.  14,  49-61. 

chi,  b.  and  hills,  l.  v.  1976.  Biostratigraphy  and  taxonomy  of  Devonian  megaspores,  Arctic  Canada.  Bull.  Can. 
Petrol.  Geol.  24,  640-818. 

chibrikova,  E.  v.  1959.  Spores  from  Devonian  and  older  deposits  of  Bashkir.  Akad.  Nauk.  S.S.S.R.,  Materials 
on  Palaeont.  and  Strat.  of  Devonian  and  older  deposits  of  Baskir,  3-116.  [In  Russian.] 


MARSHALL  AND  ALLEN:  DEVONIAN  MIOSPORES 


309 


CfflBRiKOVA,  e.  v.  1962.  Spores  of  Devonian  terrigenous  deposits  of  western  Bashkiria  and  the  western  slopes  of 
the  southern  Urals.  Acad.  Sci.  U.S.S.R.,  Bashkir.  Branch,  Inst.  Mining  Geol.,  Brachiopods,  Ostracods  and 
Spores  Middle  and  Upper  Dev.  Bashkiria.  353-476.  [In  Russian.] 

1972.  Plant  microfossils  from  the  southern  Urals  and  Cisuralia.  Acad.  Sci.  U.S.S.R.,  Bashkir.  Branch,  Inst. 

Geol.  1-222.  [In  Russian.] 

— 1977.  The  stratigraphy  of  the  Devonian  and  the  older  Palaeozoic  from  the  Ural  and  Preural  territories  of  the 
U.S.S.R.  Ibid.  1-160.  [In  Russian.] 

— and  naumova,  s.  n.  1974.  Zonal  complexes  of  Devonian  spores  and  pollen  in  the  European  part  of  the 
U.S.S.R.  and  their  analogues  abroad.  In  Palynology  of  Propterophyte  and  Palaeophyte,  Proc.  3rd.  Int. 
Palynol.  Conf.,  Novosibirsk,  1971,  39-47.  [In  Russian.] 

clayton,  G.  and  graham,  j.  r.  1974.  Miospore  assemblages  from  the  Devonian  Sherkin  Formation  of  south- 
west County  Cork,  Republic  of  Ireland.  Pollen  Spores,  16,  565-588. 

— higgs,  K.  t.  and  keegan,  j.  b.  1977.  Late  Devonian  and  early  Carboniferous  occurrences  of  the  miospore 
genus  Emphanisporites  McGregor  in  southern  Ireland.  Ibid.  19,  415-426. 

couper,  R.  A.  1953.  Upper  Mesozoic  and  Cainozoic  spores  and  pollen  grains  from  New  Zealand.  N.Z.  Geol. 
Surv.  Palaeont.  Bull.  22,  1-73. 

dawson,  j.  w.  1859.  On  fossil  plants  from  the  Devonian  rocks  of  Canada.  Q.  Jl  geol.  Soc.  London,  15, 477-488. 
donovan,  R.  n.,  archer,  r.,  turner,  p.  and  tarling,  d.  h.  1976.  Devonian  palaeogeography  of  the  Orcadian 
Basin  and  the  Great  Glen  Fault.  Nature,  Lond.  259,  550-551. 

— Collins,  A.,  Rowlands,  M.  a.  and  archer,  R.  1978.  The  age  of  sediments  on  Foula,  Shetland.  Scott.  J.  Geol. 
14,  87-88. 

— foster,  r.  j.  and  westoll,  t.  s.  1974.  A stratigraphical  revision  of  the  Old  Red  Sandstone  of  North-eastern 
Caithness.  Trans.  R.  Soc.  Edinb.  69,  171-205. 

eisenack,  a.  1944.  Uber  einige  pflanzliche  Funde  in  Geschieben,  nebst  Bemerkungen  zum  Hystricho- 
sphaerideen-Problem.  Z.  Geschiebeforsch.  19,  103-124. 

evans,  p.  h.  1970.  Revision  of  the  miospore  genera  Perotrilites  Erdtm.  ex  Couper  1953  and  Diaphanospora  Balme 
and  Hassell  1962.  Bull.  Bur.  Miner.  Resour.  Geol.  Geophys.  Aust.  116,  65-74. 
fannin,  N.  G.  T.  1970.  The  sedimentary  environment  of  the  Old  Red  Sandstone  of  western  Orkney.  Ph.D.  thesis, 
University  of  Reading. 

filiminova,  a.  b.  and  Arkhangelskaya,  A.  d.  1963.  Spore  assemblage  of  the  Vorbyev  horizon  of  the  Middle 
Devonian  of  the  Russian  Platform.  Trans.  Vnigri,  37,  31-36. 
fletcher,  a.  j.  1976.  Studies  on  Devonian  megaspores.  Ph.D.  thesis.  University  of  Bristol. 
flinn,  d.  1961.  Continuation  of  the  Great  Glen  Fault  beyond  the  Moray  Firth.  Nature,  Lond.  191,  589-591. 
guennel,  G.  k.  1963.  Devonian  spores  in  a Middle  Silurian  reef.  Grana  palynol.  4,  245-261. 
hamid,  m.  e.  p.  1974.  Sporenvergesellschaftungen  aus  dem  unteren  Mitteldevon  (Eifel-Stufe)  des  siidlichen 
Bergischen  Landes  (Rheinisches  Schiefergebirge).  Neues  Jahrb.  Geol.  Palaeont.,  Abh.  147,  163-217. 
hemer,  d.  o.  and  nygreen,  p.  w.  1967.  Devonian  palynology  of  Saudi  Arabia.  Rev.  Palaeobot.  Palynol.  5,  51-61. 
hoffmeister,  w.  s.,  staplin,  F.  L.  and  malloy,  r.  e.  1955.  Mississippian  plant  spores  from  the  Hardinsburg 
Formation  of  Illinois  and  Kentucky.  J.  Paleont.  29,  372-399. 
kedo,  G.  i.  1955.  Spores  of  the  Middle  Devonian  of  the  northeastern  Byelorussian  S.S.R.  Acad.  Sci.  B.S.S.R., 
Inst.  Geol.  Sci.,  Palaeont.  Stratigr.  B. S.S.R.  1,  5-59.  [In  Russian.] 

— 1960.  Spore  pollen  complexes  of  the  Paleozoic  deposits  of  Byelorussia.  Inter.  Geol.  Congr.,  21st  Sess., 
Rept.  Sov.  Geol.,  Prob.  6,  196-200.  [In  Russian.] 

— 1966.  Spore  of  Middle  Devonian  deposits  from  the  western  part  of  the  Russian  Platform.  Inst.  Geol. 
Riga,  Palynol.  in  Geol.  Research  in  the  Baltic  Sov.  Republ.  7,  7-14.  [In  Russian.] 

— 1976.  New  species  of  spores  from  the  Lower  and  Middle  Devonian  of  the  western  Russian  Platform.  Acad. 
Nauk.  B. S.S.R.  1976,  108-171.  [In  Russian.] 

kermack,  k.  a.  and  haldane,  J.  b.  s.  1950.  Organic  correlation  and  allometry.  Biometrika,  37,  30-41. 
lang,  w.  h.  1925.  Contributions  to  the  study  of  the  Old  Red  Sandstone  flora  of  Scotland.  1.  On  plant  remains 
from  the  fish  beds  of  Cromarty.  2.  On  a sporangium  bearing  branch  system  from  the  Stromness  Beds.  Trans. 
R.  Soc.  Edinb.  54,  253-280. 

leclercq,  s.  and  bonamo,  p.  m.  1971.  A study  of  the  fructification  of  Milleria  ( Protopteridium ) thomsonii  Lang 
from  the  Middle  Devonian  of  Belgium.  Palaeontographica,  Abt.  B,  136,  83-1 14. 

— 1973.  Rellimia  thomsonii  a new  name  for  Milleria  ( Protopteridium ) thomsonii  Lang  1926  emend. 
Leclercq  and  Bonamo  1971.  Taxon,  22,  435-437. 

lele,  k.  m.  and  streel,  m.  1969.  Middle  Devonian  (Givetian)  plant  microfossils  from  Goe  (Belgium).  Soc.  Geol. 
Belg.  Ann.  92,  89-121. 


310 


PALAEONTOLOGY,  VOLUME  25 


loboziak,  s.  and  streel,  m.  1980.  Miospores  from  Givetian  to  Lower  Frasnian  sediments  dated  with  conodonts 
in  the  Boulonnais  (France).  Rev.  Palaeobot.  Palynol.  29,  285-299. 
mcgregor,  d.  c.  1960.  Devonian  spores  from  Melville  Island,  Canadian  Arctic  Archipelago.  Palaeontology,  3, 
26-44. 

— 1961.  Spores  with  proximal  radial  pattern  from  the  Devonian  of  Canada.  Bull.  geol.  Surv.  Can.  76,  1-11. 

— 1964.  Devonian  miospores  from  the  Ghost  River  Formation,  Alberta.  Ibid.  109,  1-31. 

— 1973.  Lower  and  Middle  Devonian  spores  of  eastern  Gaspe,  Canada.  I.  Systematics.  Palaeontographica, 
Abt.  B,  144,  1-77. 

— 1977.  Lower  and  Middle  Devonian  spores  of  eastern  Gaspe,  Canada.  II.  Biostratigraphy.  Ibid.  163, 
111-142. 

— 1979a.  Spores  in  Devonian  Stratigraphical  Correlation.  In  house,  m.  r.,  scrutton,  c.  t.  and  bassett,  m.  g. 
(eds.).  The  Devonian  System.  Spec.  Pap.  Palaeont.  23,  163-184. 

— 19796.  Devonian  miospores  of  North  America.  Palynology,  3,  31-52. 

— and  camfield,  M.  1976.  Upper  Silurian?  to  Middle  Devonian  spores  of  the  Moose  River  Basin,  Ontario. 
Bull.  geol.  Surv.  Can.  263,  1-63. 

— and  uyeno,  T.  T.  1972.  Devonian  spores  and  conodonts  of  Melville  and  Bathurst  Islands,  District  of 
Franklin.  Geol.  Surv.  Pap.  Can.  71-13,  1 -37. 

marshall,  j.  e.  A.  1980.  A method  for  the  successful  oxidation  and  subsequent  stabilisation  of  high  rank,  poorly 
preserved  spore  assemblages.  Rev.  Palaeobot.  Palynol.  29,  313-319. 
massa,  D.  and  moreau-benoit,  a.  1976.  Essai  de  synthese  stratigraphique  et  palynologie  du  systeme  Devonien  en 
Libye  occidentale.  Revue  Inst.  fr.  Petrole,  31,  287-333. 

moreau-benoit,  a.  1979.  Les  spores  du  Devonien  de  Libye,  Premiere  partie.  Cahiers  de  Micropaleontologie,  1979 
(4),  1-58. 

— 1980.  Les  spores  du  Devonien  de  Libye.  Deuxieme  partie.  Ibid.  1980  (1),  1-53. 

Mortimer,  m.  G.  and  chaloner,  w.  G.  1972.  The  palynology  of  concealed  Devonian  rocks  in  southern  England. 
Bull.  geol.  Surv.  Gr.  Br.  39,  1-56. 

mykura,  w.  1972a.  The  Old  Red  Sandstone  sediments  of  Fair  Isle,  Shetland  Islands.  Ibid.  41,  1-31. 

19726.  Igneous  intrusions  and  mineralisation  in  Fair  Isle,  Shetland  Islands.  Ibid.  33-53. 

1976.  Orkney  and  Shetland.  British  Regional  Geology.  Institute  Geological  Sciences,  Edinburgh. 

— and  young,  b.  r.  1969.  Sodic  Scapolite  (Dipyre)  in  the  Shetland  Islands.  Rep.  Inst.  geol.  Sci.,  No. 
69  (4),  1-8. 

nathorst,  A.  G.  1915.  Zur  Devonflora  des  westlichen  Norwegens.  Bergens  Mus.  Aarbok,  1914-15,  No.  9,  12-34. 
naumova,  s.  N.  1953.  Spore-pollen  assembages  of  the  Upper  Devonian  of  the  Russian  Platform  and  their 
statigraphic  significance.  Trans.  Inst.  Geol.  Sci.,  Acad.  Sci.  U.S.S.R.  143  (Geol.  Ser.  60),  1-204.  [In  Russian.] 
neves,  R.  and  dale,  b.  1963.  Modified  filtration  system  for  palynological  preparations.  Nature,  Lond.  198, 
775-776. 

and  Owens,  b.  1966.  Some  Namurian  camerate  miospores  from  the  English  Pennines.  Pollen  Spores,  8, 

337-360. 

owens,  b.  1971.  Miospores  from  the  Middle  and  early  Upper  Devonian  rocks  of  the  western  Queen  Elizabeth 
Islands,  Arctic  Archipelago.  Geol.  Surv.  Pap.  Can.  70-38,  1-157. 

— and  richardson,  j.  b.  1972.  Some  recent  advances  in  Devonian  palynolygy— a review.  Report  of  C.I.M.P. 
Working  Group  No.  13B.  7e  Cong.  Inter.  Stratig.  Geol.  Carbonifere,  1971,  1,  325-343. 

ozolin'a,  v.  1955.  Upper  Devonian  (Dze)  spore-pollen  complexes  from  the  Latvian  S.S.R.  Latvijas  P.S.R. 
Zinatnu  Akademijas  izdevums,  Riga.  2 (91),  53-61.  [In  Latvian.] 

— 1960a.  Data  on  investigation  of  spores  and  pollen  of  the  Frasnian  stage  of  the  Upper  Devonian  of  the 
Latvian  S.S.R.  Latv.  Acad.  Sci.,  Geologijas  un  Derigo  Izraktenu  Instituta  Raksti,  5,  171-187.  [In  Latvian.] 

— 19606.  Middle  Devonian  spores  from  the  Latvian  S.S.R.  Ibid.  189-205.  [In  Latvian.] 

— 1961.  Spore  pollen  spectra  of  the  Frasnian  Stage  of  the  Upper  Devonian  of  the  Latvian  S.S.R.  Ibid.  7, 
129-139.  [In  Russian.] 

— 1963.  Spore  pollen  spectra  of  the  Frasnian  stage  in  the  Alanda  Well.  Ibid.  10,  129-139.  [In  Russian.] 
pacht,  r.  1849.  Der  devonische  kalk  in  Livland,  Ein  Beitrag  zur  Geognoise  de  Ostee  provinzen.  Dorpat. 
peppers,  R.  a.  and  damberger,  h.  h.  1969.  Palynology  and  Petrography  of  a Middle  Devonian  coal  in  Illinois. 

Circ.  III.  St.  geol.  Surv.  445,  1-36. 

pettitt,  j.  m.  1965.  Two  heterosporous  plants  from  the  Upper  Devonian  of  North  America.  Bull.  Br.  Mus.  Nat. 
Hist.  10,  81-92. 

playford,  g.  1971.  Lower  Carboniferous  spores  from  the  Bonaparte  Gulf  Basin,  Western  Australia  and 
Northern  Territory.  Bull.  Bur.  Miner.  Resour.  Geol.  Geophys.  Aust.  115,  1-104. 


MARSHALL  AND  ALLEN:  DEVONIAN  MIOSPORES 


311 


— 1976.  Plant  microfossils  from  the  Upper  Devonian  and  Lower  Carboniferous  of  the  Canning  Basin, 
Western  Australia.  Palaeontographica,  Abt.  B,  158,  1-71. 

Pokrovskaya,  i.  m.  (ed.)  1966.  Palaeopalynology.  Proc.  all  Union  Sci.  Res.  Geol.  Inst.  n.s.  141.  3 vols.  [In 
Russian.] 

potonie,  r.  1956.  Synopsis  der  Gattungen  der  Sporae  dispersae.  I.  Teil:  Sporites.  Beih.  Geol.  Jahrb.  23,  1-103. 

— and  kremp,  G.  1 954.  Die  Gattungen  der  palaozoischen  Sporae  dispersaeund  ihre  Stratigraphie.  Geol.  Jahrb. 
69,  111-194. 

raskatova,  L.  G.  1969.  Spore  complexes  from  the  Middle  and  Upper  Devonian  from  the  south  eastern  part  of 
the  Central  Devonian  field.  Voronezh  University  Press , Voronezh,  1969,  1-168.  [In  Russian.] 

— 1974.  Detailed  stratigraphic  subdivision  of  the  Devonian  of  the  Central  Devonian  field  in  the  light  of 
palynological  evidence.  In  Palynology  of  Proterophyte  and  Palaeophyte,  Proc.  3rd  Int.  Palynol.  Conf., 
Novosibirsk,  1971,  67-71.  [In  Russian.] 

richardson,  j.  b.  1960.  Spores  from  the  Middle  Old  Red  Sandstone  of  Cromarty,  Scotland.  Palaeontology,  3, 
45-63. 

— 1962.  Spores  with  bifurcate  processes  from  the  Middle  Old  Red  Sandstone  of  Scotland.  Ibid.  5,  171- 
194. 

— 1965.  Middle  Old  Red  Sandstone  spore  assemblages  from  the  Orcadian  Basin,  north-east  Scotland.  Ibid.  7, 
559-605. 

— 1967.  A reconnaissance  of  some  Upper  Devonian  and  Lower  Carboniferous  spores  from  New  York  State 
and  Pennsylvania  (U.S.A.).  Rev.  Palaeobot.  Palynol.  1,  63-64.  (Abs.  only.) 

— 1969.  Devonian  spores.  In  tschudy,  r.  h.  and  scott,  r.  a.  (eds.),  Aspects  of  Palynology,  193-222.  Wiley- 
Interscience,  New  Y ork. 

— 1974.  The  stratigraphic  utilization  of  some  Silurian  and  Devonian  miospore  species  in  the  northern 
hemisphere:  an  attempt  at  a synthesis.  Int.  Symp.  Belgian  Micropalaeont.  Limits,  Namur,  1974.  Publ. 
No.  9,  1-13. 

riegel,  w.  1968.  Die  Mitteldevon-Flora  von  Lindlar  (Rheinland).  2.  Sporae  dispersae.  Palaeontographica,  Abt. 
B, 123,  76-96. 

— 1973.  Sporenformen  aus  den  Heisdorf-,  Lauch-  und  Nohn-Schichten  (Emsium  und  Eifelium)  der  Eifel, 
Rheinland.  Ibid.  142,  78-104. 

— 1974.  Spore  floras  across  the  Lower/Middle  Devonian  boundary  in  the  Rhineland  (G.F.R.).  Palynology  of 
Proterophyte  and  Palaeophyte,  Proc.  3rd  Int.  Palynol.  Conf.,  Novosibirsk,  1973,  47-53. 

Sanders,  r.  B.  1968.  Devonian  spores  of  the  Cedar  Valley  coal  of  Iowa,  U.S.A.  J.  Palynology  2 and  3,  17-32. 
schopf,  J.  M.,  wilson,  L.  R.  and  bentall,  r.  1944.  An  annotated  synopsis  of  Palaeozoic  fossil  spores  and  the 
definition  of  generic  groups.  Rep.  Inv.  III.  State  Geol.  Surv.  91,  1-72. 
smith,  a.  h.  v.  and  butter  worth,  m.  a.  1967.  Miospores  in  the  coal  seams  of  the  Carboniferous  of  Great  Britain. 
Spec.  Pap.  Palaeont.  1,  1-324. 

smith,  D.  i.  1977.  The  Great  Glen  Fault.  In  The  Moray  Firth  Area  Geological  Studies,  46-59.  Inverness  Field 
Club. 

staplin,  F.  l.  1960.  Upper  Mississippian  plant  spores  from  the  Golata  Formation,  Alberta,  Canada. 
Palaeontographica,  Abt.  B 107,  1-40. 

and  jansonius,  j.  1964.  Elucidation  of  some  Palaeozoic  Densospores.  Ibid.  114,  95-1 1 7. 

streel,  M.  1964.  Une  association  de  spores  du  Givetian  inferieur  de  la  Vesdre,  a Goe  (Belgique).  Ann.  Soc.  Geol. 
Belg.  87,  233-262. 

— 1967.  Associations  de  spores  du  Devonien  inferieur  Beige  et  leur  signification  stratigraphique.  Ibid.  90, 
11-53. 

— 1972.  Dispersed  spores  associated  with  Leclercqia  complexa  Banks,  Bonamo  and  Grierson  from  the  late 
Middle  Devonian  of  New  York  State.  (U.S.A.).  Rev.  Palaeobot.  Palynol.  14,  205-215. 

taugourdeau-lantz,  j.  1967.  Spores  nouvelles  du  Frasnien  du  Bas  Boulonnais  (France).  Revue  micropaleont. 
10,  48-60. 

till,  R.  1974.  Statistical  methods  for  the  earth  scientist,  pp.  1-154.  Macmillan. 

urban,  j.  b.  1970.  Ancyrospora  fallacia,  a new  sporomorph  exhibiting  deceptive  variations  in  preservation. 
Micropaleontology,  16,  221-226. 

van  der  zwan,  c.  j.  1979.  Aspects  of  Late  Devonian  and  Early  Carboniferous  palynology  of  southern  Ireland.  1 . 
The  Cyrtospora  cristifer  morphon.  Rev.  Palaeobot.  Palynol.  28,  1 -20. 

— 1980.  Aspects  of  Late  Devonian  and  Early  Carboniferous  palynology  of  southern  Ireland.  2.  The 
Auroraspora  macra  morphon.  Ibid.  30,  133-155. 

vigran,  j.  o.  1964.  Spores  from  Devonian  deposits,  Mimerdale,  Spitsbergen.  Skr.  nor.  Polarinst.  132,  1-32. 


312 


PALAEONTOLOGY,  VOLUME  25 


westoll,  t.  s.  1979.  Devonian  fish  biostratigraphy.  In  house,  m.  r.,  scrutton,  c.  t.  and  bassett,  m.  g.  (eds.).  The 
Devonian  System.  Spec.  Pap.  Palaeont.  23,  341-353. 

Williamson,  j.  h.  1968.  Least-squares  fitting  of  a straight  line.  Can.  J.  Phys.  46,  1845-1847. 
wilson,  l.  r.  and  coe,  e.  z.  1 940.  Descriptions  of  some  unassigned  plant  microfossils  from  the  Des  Moines  Series 
of  Iowa.  Amer.  Midi.  Nat.  23,  182-186. 

york,  d.  1966.  Least-squares  fitting  of  a straight  line.  Can.  J.  Phys.  44,  1079-1086. 

J.  E.  A.  MARSHALL 

Gearhart  Geodata  Services  Ltd 
Howe  Moss  Drive 
Kirkhill  Industrial  Estate 
Dyce,  Aberdeen  AB2  OGL 

k.  c.  ALLEN 

Department  of  Botany 
The  University 
Bristol 
BS8  1UG 


Typescript  received  10  June  1980 

Revised  typescript  received  12  December  1980 


CONODONTS,  GONIATITES  AND  THE 
BIOSTRATIGRAPHY  OF  THE  EARLIER 
CARBONIFEROUS  FROM  THE  CANTABRIAN 
MOUNTAINS,  SPAIN 

by  A.  C.  HIGGINS  and  C.  H.  t.  wagner-gentis 


Abstract.  Six  conodont  zones,  including  one  new  one,  the  Paragnathodus  multinodosus  Zone,  are  represented  in 
the  Lower  Carboniferous  of  the  Cantabrian  Mountains.  The  absence  of  conodont  zones  known  to  occur  in 
north-west  Europe  points  to  four  major  stratigraphic  gaps  in  the  Spanish  sequence,  three  in  the  Tournaisian  and 
one  in  the  Visean.  High  conodont  abundances,  indicating  a low  rate  of  sedimentation  in  the  Tournaisian  rocks, 
and  the  numerous  gaps  in  the  stratigraphic  record  point  to  a slow  initial  Carboniferous  transgression.  The 
patchy  distribution  of  earliest  Tournaisian  sediments  contrasts  with  the  widespread  distribution  of  the  latest 
(anchor alis  Zone)  Tournaisian  sediments  when  the  transgression  reached  its  culmination.  Lower  conodont 
abundances  and  thicker  successions  coincide  with  a more  complete  Visean/early  Namurian  sequence  in  which 
only  one  major  break  is  detectable.  A goniatite  fauna  from  the  lower  Visean  is  described  including  one  new 
genus,  Pseudogirtyoceras  and  two  new  species  P.  villabellacoi  and  Winchelloceras palentinus.  This  fauna  confirms 
the  early  Visean  age  of  the  typicus  Zone.  One  new  species  of  conodont,  Scaliognathus  angustilateralis  sp.  nov.,  is 
also  named. 


The  Dinantian  and  early  Namurian  sedimentary  sequence  in  the  Cantabrian  Mountains  of  north- 
west Spain  consists  of  less  than  50  m of  condensed  deposits  formed  on  the  Cantabrian  Block.  This 
was  a stable  platform  uplifted  during  the  late  Famennian  and  the  recipient  of  renewed  sedimentation 
from  latest  Famennian  onwards.  The  age  and  nature  of  these  sediments  is  not  always  the  same  all 
over  the  area  and  this  is  true  particularly  of  the  deposits  formed  during  the  latest  Famennian  and 
Tournaisian.  Several  formations  are  involved  (compare  Wagner,  Winkler  Prins,  and  Riding  1971). 
The  earliest  deposits  belong  to  the  Ermita  Formation  (Gres  de  l’Ermitage  of  Comte  1959)  which  was 
laid  down  after  the  Famennian  uplift  and  which  appears  to  have  a diachronous  base,  older  at  the 
flanks  of  the  uplift  and  younger  in  its  central  part.  A calcareous  sandstone  lithology  is  the  most 
common  for  the  Ermita  Formation  but  limestone  has  been  recorded  in  the  top  part.  Samples  in  the 
highest  Ermita  Formation  yielded  conodonts  attributed  to  the  costatus  Zone  (Higgins,  Wagner- 
Gentis,  and  Wagner  1964),  but  a few  localities  have  now  shown  the  presence  of  a Protognathodus 
fauna  of  earliest  Carboniferous  age.  The  Ermita  Formation  thus  appears  to  span  the  latest 
Famennian  and  the  earliest  Tournaisian. 

In  the  south-eastern  part  of  the  Cantabrian  Mountains,  corresponding  to  the  northern  part  of  the 
province  of  Palencia,  there  is  an  area  which  shows  many  resemblances  to  the  western  Pyrenees  and 
which  is  characterized  by  a nodular  limestone,  the  Vidrieros  Formation  of  Van  Veen  (1965), 
spanning  the  Famennian -Tournaisian  boundary  (Van  Adrichem  Boogaert  1967).  This  region  does 
not  show  the  effects  of  Famennian  uplift  and  contains  a different  development  of  Devonian  strata,  i.e. 
the  Palentine  facies  of  Brouwer  (1964).  It  occupies  an  area  between  two  major  faults,  viz.  the  Ruesga 
Fault  in  the  south  and  the  southern  boundary  fault  of  the  Picos  de  Europa  to  the  north.  Appreciable 
tectonic  shortening  is  a feature  of  both  faults  and  the  abrupt  contacts  between  areas  of  different  facies 
development  is  apparently  due  to  extreme  telescoping  as  a result  of  the  tightening  of  the  arcuate  fold 
belt  in  north-west  Spain  (R.  H.  Wagner,  pers.  comm.). 

The  Ruesga  Fault  is  apparently  continued  westwards  by  the  Leon  Fault  of  Marcos  (1967) 


IPalaeontology,  Vol.  25,  Part  2,  1982,  pp.  313-350,  pis.  34-36-1 


314 


PALAEONTOLOGY,  VOLUME  25 


(originally  described  as  the  Leon  Line  by  De  Sitter  1962).  Although  this  important  fracture  zone  has 
been  widely  credited  as  a syn-sedimentary  control,  it  seems  to  be  due  to  post-sedimentary  tectonics 
with  a telescoping  effect  which  gradually  diminishes  westwards  (R.  H.  Wagner,  pers.  comm.).  The 
sediments  investigated  for  the  present  paper  were  all  laid  down  south  of  the  Leon-Ruesga  tectonic 
fault  line  (text-fig.  1)  even  though  some  of  the  localities  are  presently  north  of  this  fault. 


text-fig.  1 . Structural  palaeogeographic  areas.  Palaeozoic,  Cantabrian  Mountains. 


In  this  area  the  Ermita  Formation  is  generally  present.  It  is  followed  by  either  the  Vegamian 
Formation  of  black  shales  or  the  Baleas  Formation  of  crinoidal  limestone.  These  two  formations 
(Wagner  et  al.  1971)  seem  to  be  mutually  exclusive.  Since  they  both  seem  to  correspond  to  the  same 
time  interval  within  the  Tournaisian  (cooperi-communis  Zone  of  the  present  paper)  it  may  be  that  the 
Baleas  Limestone  was  formed  on  ridges  in  the  basin  which  received  the  Vegamian  Shales  (compare 
Higgins  et  al.  1964).  The  Baleas  Formation,  described  by  Wagner  et  al.  (1971)  is  only  a few  metres 
thick.  The  Vegamian  Formation  (Comte  1959;  Wagner  etal.  1971)  is  generally  less  than  10mthick.lt 
is  characterized  by  phosphatic  and  chert  nodules  and  black  shales  which  are  often  cherty.  Its 
macrofauna  includes  the  goniatite  Muensteroceras  arkansanum  Gordon,  as  recorded  by  Wagner- 
Gentis  (in  Wagner  et  al.  1971).  This  species  was  described  originally  from  the  late  Kinderhookian  of 
Arkansas,  U.S.A. 

The  most  widespread  formation  in  the  Carboniferous  of  north-west  Spain  is  the  succeeding 
Genicera  Formation  of  Wagner  et  al.  (1971)  which  is  the  ‘Marbre  Griotte’  of  Barrois  (1882)  (also 
called  Villabellaco  Formation— Wagner  and  Wagner-Gentis  1963— and  Alba  Formation— Van 
Ginkel  1965).  It  is  characterized  by  nodular  and  wavy-bedded  limestones  and  up  to  25  m thick.  The 
basal  unit  (Gorgera  Member  of  Wagner  et  al.  1 97 1 ) is  marly  and  usually  a vivid  red  colour,  although  a 
slightly  reddish-grey  colour  has  also  been  found.  Over  a large  area  of  northern  Leon,  including  the 
Genicera  type  section,  an  interval  of  red  shales  and  cherts  (Lavandera  Member)  separates  the  basal, 


HIGGINS  AND  WAGNER-GENTIS:  CONODONTS  AND  GONI ATITES 


315 


marly  unit  of  nodular  limestones  from  the  main  part  of  generally  wavy-bedded  to  nodular  limestones 
(Canalon  Member  of  Wagner  et  al.  1971).  The  chert  unit  is  absent  in  the  southernmost  exposures  of 
northern  Leon  and  in  the  Revilla  Nappe  structure  of  northern  Palencia.  Conodont  work  reported  in 
the  present  paper  shows  that  the  chert  unit  corresponds  to  a sizable  time  gap  which  is  equally 
apparent  where  the  cherts  are  absent.  The  basal  unit,  which  is  only  a few  metres  thick,  has  yielded 
goniatites  of  the  late  Pericyclus  (II  j 8/y)  Zone  corresponding  to  late  Tournaisian.  Its  conodont  fauna 
corresponds  to  the  anchoralis  Zone.  The  Gorgera  Member  may  be  occasionally  absent  and  although 
the  rare  occurrences  where  this  seems  to  be  the  case  ought  to  be  checked  for  thrusting  (shearing  in  the 


Age 

SSW  of  Genicera  (Leon) 

Olleros  de  Alba 
(Leon) 

Formation 

Member 

Formation 

Namurian  A/B 

Barcaliente 

Barcaliente 

Olleros 

Genicera 

Canalon 

Olaja  Beds 
Genicera 

Visean 

Lavandera 

Gorgera 

Tournaisian 

Vegamian 

Vegamian 

Ermita 

Ermita 

Famennian 

table  1.  Formations  in  the  Porma-Bernesga  area,  based  on  Wagner  et  al.  (1971). 


flanks  of  isoclinal  synclines  is  of  common  occurrence  in  the  Cantabrian  Mountains),  it  appears  likely 
that  the  base  of  the  Genicera  Formation  may  not  always  be  of  exactly  the  same  age. 

The  top  of  the  formation  reaches  into  the  lower  Namurian  (E2b  Zone)  and  is  undoubtedly 
diachronous.  Wagner  et  al.  (1971)  described  the  Olaja  Beds,  a condensed  mudstone  sequence  of  only 
a few  metres,  as  the  lateral  equivalent  of  the  highest  part  of  the  Canalon  Member.  The  lower 
Namurian  Olaja  Beds  occur  only  in  the  southernmost  exposures  of  the  Cantabrian  Mountains  in 
northern  Leon,  where  they  form  the  base  of  a thick  terrigenous  sequence  with  turbidites  (Olleros 
Formation  of  Wagner  et  al.  1971).  In  the  more  northerly  exposures,  the  more  complete  Canalon 
Member  is  followed  with  gradual  transition  by  the  Barcaliente  Limestone  Formation.  The  latter  was 
formed  on  a carbonate  platform  (the  same  Cantabrian  Block  as  accumulated  the  Genicera  Limestone 
and  Chert  Formation)  which  was  sufficiently  shallow  to  give  rise  to  evaporitic  sediments.  The 
presence  of  a turbidite  basin  to  the  south  indicates  the  diminished  area  of  the  Cantabrian  Block  in 
mid-Namurian  times,  whilst  the  terrigenous  facies  of  the  Olaja  Beds  provides  an  early  indication  of 
the  northwards  withdrawal  of  the  southern  margin  of  the  block  in  early  Namurian  times. 


316 


PALAEONTOLOGY,  VOLUME  25 


text-fig.  2.  Location  of  the  sampled  sections  in  the  southern  part  of  the  Cantabrian  Mountains. 


CONODONT  ZONES  AND  LOWER  VISEAN  GONIATITE  OCCURRENCE 

Six  conodont  zones  and  a Protognathodus  fauna  are  represented  in  the  sequence  but  the  boundaries  of  the  fauna 
and  the  lower  three  zones  are  non-sequences  and  the  sequence  is  probably  incomplete.  Five  of  these  concurrent 
range  zones  were  defined  by  Higgins  (1974)  but  are  here  modified  and  one  zone  is  new. 

Protognathodus  fauna.  The  base  of  the  Ermita  Formation  is  a transgressive  horizon  of  late  Devonian  age, 
belonging  to  the  costatus  Zone  of  Ziegler  (1962).  In  only  two  of  the  sections,  the  Aviados  and  the  Villabellaco 
sections  in  the  Porma-Bernesga  area,  does  the  base  of  the  Carboniferous  rest  on  pr e-costatus  rocks.  On  the  other 
hand,  only  at  Santiago  de  las  Villas  and  Olleros  de  Alba  is  there  a transition  between  the  Devonian  and 
Carboniferous  where  a thin  limestone  at  the  top  of  the  Ermita  Formation  yields  a conodont  fauna  belonging  to 
the  Protognathodus  fauna.  This  fauna  includes  Polygnathus  communis  communis  Branson  and  Mehl, 
Polygnathus  inornatus  sensu  Branson  and  Mehl,  Protognathodus  kockeli  (Bischoff ),  Protognathodus  meischneri 
Ziegler,  Protognathodus  kuehni  Ziegler  and  Leuteritz,  and  one  specimen  doubtfully  referred  to  Pseudo- 
polygnathus  dentilineatus.  This  fauna  was  referred  to  the  kockeli-dentilineatus  Zone  by  Higgins  et  al.  (1964)  as 
was  a similar  fauna  described  by  Adrichem  Boogaert  (1967)  from  the  Triollo  area.  The  presence  of 
Protognathodus  kuehni  suggests  that  this  fauna  belongs  to  the  younger  Protognathodus  fauna  of  Ziegler  (1969). 
The  presence  of  a newly  discovered  siphonodellid  in  this  fauna  also  clearly  indicates  closer  affinity  with  the 
Upper  rather  than  the  Lower  Protognathodus  fauna.  This  siphonodellid  was  referred  by  Higgins  (1974)  to 
Siphonodella  sulcata  (Huddle)  but  it  is  poorly  preserved  and  of  very  doubtful  determination.  The  assignment  of 
these  samples  to  the  Upper  Protognathodus  fauna  places  them  in  the  Carboniferous  rather  than  at  the  top  of  the 
Devonian  (Sandberg,  Ziegler,  Leuteritz,  and  Brill  1978). 

Siphonodella  cooperi-polygnathus  communis  Zone.  This  zone,  defined  by  Higgins  (1974),  is  widespread  in  the 
Porma-Bernesga  area  where  it  occurs  in  the  Baleas  and  Vegamian  Formations.  Key  species  in  the  rich  and 
abundant  faunas  include  Siphonodella  obsoleta  Hass,  S.  duplicata  Branson  and  Mehl  morphotype  2,  and  S. 
cooperi  Hass  morphotype  2 of  Sandberg  et  al.  (1978).  These  range  from  the  Upper  Duplicata  Zone  to  the  Upper 
Crenulata  Zone  according  to  Sandberg  et  al.  (1978).  However,  Gnathodus  delicatus  Branson  and  Mehl,  G. 
punctatus  (Cooper),  and  G.  semiglaber  Bischoff,  which  are  also  present  in  the  zone,  overlap  the  range  of  species  of 
Siphonodella  in  the  Belgian  Tournaisian  (Groessens  \911b)  in  late  Tn2.  Also  present  in  the  zone  is  Polygnathus 
communis  carina  Hass  which  first  appears  in  Tn3  in  the  Belgian  sequence.  A late  Tn2/early  Tn3  age  indicates  the 
span  of  the  cooperi-communis  Zone.  This  age  probably  indicates  at  least  partial  equivalence  to  the  German 
crenulata  Zone  (see  text-fig.  7). 

Anchoralis  Zone.  The  anchoralis  Zone  is  easily  recognized  and  is  a widespread  zone  in  the  Cantabrian 
Mountains.  Scaliognathus  anchoralis  Branson  and  Mehl,  Doliognathus  latus  Branson  and  Mehl,  Pseudo- 
polygnathus  triangulus  pinnatus,  Gnathodus  texanus  (Roundy)  pseudo  semiglaber  Thompson,  and  G.  delicatus  are 


HIGGINS  AND  WAGNER-GENTIS:  CONODONTS  AND  GONI ATITES  317 


text-fig.  3.  Ranges  of  the  stratigraphically  important  conodont  species  in  the  Cantabrian  Mountains. 


318 


PALAEONTOLOGY,  VOLUME  25 


table  2.  Distribution  and  abundance  of  conodonts  from  the  Cantabrian  Mountains. 


among  the  key  indices  in  a rich  and  varied  fauna.  D.  latus  is  used  as  a basal  subzonal  index  in  the  Missouri 
(Thompson  and  Fellows  1970)  and  Belgian  (Groessens  \911b)  sections.  In  the  Porma-Bernesga  area,  D.  latus  is 
present  throughout  the  zone  and  in  most  samples  of  this  age,  which  implies  that  the  upper  part  of  the  zone  is 
missing  in  the  southern  part  of  the  Cantabrian  Mountains,  and  that  the  zone  is  of  early  Tn3c  age. 

Gnathodus  typicus  Zone.  This  zone  was  defined  by  Marks  and  Wensink  from  the  Pyrenees  (1970).  The 
characteristic  species  G.  typicus  Cooper,  G.  antetexanus,  and  G.  texanus  pseudosemiglaber  mark  the  beginning  of 


HIGGINS  AND  WAGNER-GENTIS:  CONODONTS  AND  GONI ATITES 


319 


\ species 
sample  \ 

Pinornotus 
Bi.  stabilis 
S.  obsoleta 
G de/icatus 
G semig/ober 

Ps  triangu/us  pinna t us 
D lotus 
Sc  anchora/is 
Sc.  angusti/atera/is 
Pbischoffi 
G ant et exanus 

G.  texanus  pseudosemig/aber 
G typicus 
G symmutatus 
G homopunctatus 

Pa  commutatus 
G bi/ineafus 
Sp.  Campbell i 
G girtyi 
Pa  nodosus 
Pa.  mu/tinodosus 

total  per  kilo 

1197B 

2 

3 14 

19 

Robledo 

1 

3 2 

6 

de  Caldas  1194 

8 

9 

1193 

4 

i ■’ 

18 

Viadangos  1^58 

10  20 

30 

de  Arbos  1357 

5 1 

0 |0 

48 

3520 

2 

i 

2 

9 

N 

20  i 

15 

3 

40 

14 

4 

25 

25  - 

2 

41 

1 

1 

12 

i 

2 

47 

F 

6 30  ! 

20 

83 

E 

2 K)  1 

D 

2 40  ; 

44 

c 

2 

0 15 

2 1 

40 

B 

1 2 

38 

A 

6 2 

2 

3077 

1069T 

i 

S 

2 

R 

4 

18 

L 

4 

5 

C t'  K 

2 

2 

Getino  j 

6 16 

2! 

5 25 

4 3 

H 

10  50  < 

82 

G 

F 

4 

E 

2 10 

D 

7 25  ; 

C 

6 20 

B 

7 21 

28 

A 

10  60  ( 

9 7 

1072 N 

l 

1 

M 

1 12 

13 

L 

2 

2 

ii 

4 

Venta  de  G 

3 

c 

8 

5 

4 

1340  1 

> 

11698 

1 

II 

3 

Genicera  ,,66c  - 

1 

i 

20 

5 

27 

23 

4 10 

91  1 

158 

1338 

4 

\ species 
sample  \ 

Pr  kocke/i 
Pr  meischneri 
Ps.  primus 

P communis  communis 

Pinornotus 
Bi  stabilis 
Ps.  mu/tistriafus 
G.  de/icatus 

Ps.  triangu/us  pinna t us 
D.  lotus 
S anchora/is 
G.  antetexanus 
G texanus  semig/ober 
G typicus 
G symmutatus 
G.  homopunctatus 
Pa  commutatus 
G bi/ineatus 
Sp.  campbelli 

Pa.  nodosus 
Pa.  mu/tinodosus 
M bipluti 

total  per  kilo 

1120 

1119 

1118 

1116 

1115 

Santiago  1114 

de  las  Villas  1113 
1112 

1110 

1314 

1311 

1310 

1 

1 

1 

1 

2 

in 

i. 

1 

1 

2 

40 

36 

20 

1 

8 

• 1 

1 42 

3 

58 

1149T 

Entrago  h 

0 

B 

1 1 

2 

1 

1 

1 

14 

1 ; 

9 

127 

1 

II 

■ 2 

2 2 ; 

12  144 

9 

165 

14  < 

1 2 15 

3 

38 

10  12 

1 

23 

4 25 

29 

1 2 

' 

2916 

2915 

2913 

2912 

2911 

2910 

2909 

Villabellaco  ““ 

2905 

2904 

2903 

2902 

2901 

138C 

1 2 

l 

4 

6 

8 

4 4 

8 

:■  18  21 

3 

48 

14  36 

53 

1 21 

1 

..  1 

53 

Revilla  V| 

25 

25 

133 

31 

31 

\ 9 51 

2 

1 28 

5 

5 55 

3 3 25 

i' 

4o 

5 23153  f 

18 

7 30 

3 ' 

1 

12 

7 

Sta.Olajad.  Varga  372 

_ _ , 1103m 

S.E.  of  lv 

Genicera  y 

6 

6 

90  - 

10 

104 

70 

table  3.  Distribution  and  abundance  of  conodonts  from  the  Cantabrian  Mountains. 


the  dominance  of  species  of  Gnathodus  which  characterizes  much  of  the  Visean  and  early  Namurian.  G. 
homopunctatus  Bischoflf  appears  at  the  base  of  the  zone  and  this  occurrence  together  with  the  other  species  would 
suggest  an  equivalence  of  this  zone  to  the  G.  homopunctatus  Subzone  of  Vla  age  in  Belgium  (Groessens  19776). 

The  goniatites  from  the  basal  marly  unit  ( typicus  Zone)  of  the  Villabellaco  Limestone  in  northern  Palencia 
have  been  found  in  the  following  sequence: 

138  D,  5 0 m above  the  base  Pseudogirtyoceras  villabellacoi  sp.  nov. 

Merocanites  marshallensis  (Winchell) 

Merocanites  subhenslowi  Wagner-Gentis 
Ammonellipsites  kayseri  (Schmidt) 


320 


PALAEONTOLOGY,  VOLUME  25 


138  c,  2-3  m above  the  base 


1 - 1 - 2 m above  the  base 

138  b,  0-5-0-6  m above  the  base 
136  B,  east  of  locality  138,  at 
the  same  level  as  138  c 


Merocanites  marshallensis  (Winchell) 
Merocanites  subhenslowi  Wagner-Gentis 
Ammonellipsites  kayseri  (Schmidt) 

N autellipsites  hispanicus  (Foord  and  Crick) 
Merocanites  marshallensis  (Winchell) 
Muensteroceras  cf.  crassum  Foord 
Winchelloceras  palentinus  sp.  nov. 
Muensteroceras  parallelum  (Hall) 


Merocanites  marshallensis  and  Winchelloceras  are  known  from  the  Marshall  Sandstone  in  Michigan,  which  is 
considered  to  belong  to  the  Osagean  (Manger  1979,  pp.  214,  215).  Popov  (1968)  recorded  Winchelloceras  from 
the  Fascipericyclus  3 Zone  (Lower  Visean)  of  the  Tien  Shan  in  Central  Asia  and  (1975)  from  the  CjV,  unit  in  the 
Urals. 

Ammonellipsites  kayseri  has  been  described  from  Liebstein  (Erdbach  and  Breitscheid  section).  West  Germany 
in  the  Pey  Zone,  in  Aragon  (Spain),  Schmidt  (1931)  mentioned  it  with  a Lower  Visean  fauna  and  in  south-west 
England  it  occurs  at  Tawstock  and  Coddon  Hill  (Prentice  and  Thomas  1960,  p.  6). 

Muensteroceras  parallelum  was  described  from  the  Rockford  Limestone  of  Indiana.  A conodont  fauna  below 
the  ammonoid  layer  in  the  type  locality  belongs  to  the  Osagean  (Rexroad  and  Scott  1964)  and  Manger  (1979, 
p.  213)  attributes  the  goniatite  bed  also  to  the  Osagean.  The  species  is  also  known  from  the  Sj  unit  of  the  Hassi 
Sguilma  Stage  of  Pareyn  (1961,  p.  50  and  table  IV)  which  contains  a basal  Visean  fauna.  Librovitch  (1927,  p.  42) 
mentioned  this  species  from  the  Tien  Shan,  where  it  occurs  with  a fauna  of  the  middle  and  upper  1 Pericyclus 
Stufe’  (Schmidt  1925). 


Revilla 


Conodont  Zones 


Villabellaco 


mu/tinodosus 


nodosus 


bi/ineatus 


typ/cus 


2916 
2915 
2914 
2913  H 
2912 
2911 
2910 
2909  H 
2908- 
2907- 
2906- 
2905- 
2904- 
2903. 
2902- 
2901- 
138D- 
138C- 
138B- 
2899- 
2898  H 


134X1- 

X- 

IX- 

VIII- 

VII- 

— VI- 
V- 

IV- 

— Ill- 


Goniatites 

Pseudogirtyoceras  villabellacoi 
Nauteiiipsites  hispanicus 
Muensteroceras  cf.  crassum 
Merocanites  marshallensis 
Merocanites  subhenslowi 
Ammonellipsites  kayseri 

Winchelloceras  palentinus 


Grey,  nodular 
and  wavy  bedded 
limestone 


metres 


text-fig.  4.  Conodont  zones  and  lower  Visean  goniatite  occurrence  in  the  Villabellaco  area,  Revilla  Nappe. 


HIGGINS  AND  WAGNER-GENTIS:  CONODONTS  AND  GONIATITES 


321 


Muensteroceras  cf.  crassum  is  known  from  the  Lower  Carboniferous  Limestone  of  Ballinacarriga,  Co. 
Limerick,  Eire. 

Merocanites  subhenslowi  and  Nautillepsites  hispanicus  have  previously  been  described  from  Olleros  de  Alba 
(Leon)  and  Villabellaco  (Palencia)  (Higgins  et  al.  1964;  Wagner-Gentis  1960).  In  both  papers  it  was  suggested 
that  they  occurred  in  an  equivalent  to  the  lower  B-zone,  but  this  is  now  seen  to  be  incorrect.  An  earlier  horizon  is 
involved.  The  lower  part  of  the  Villabellaco  Limestone  yielded  conodonts  of  the  typicus  Zone  (Higgins,  this 
paper)  which  corresponds  to  Vla  of  the  Belgian  sequence.  An  early  Visean  age  is  also  indicated  by  the  majority  of 
the  goniatite  species  found. 

The  goniatite  fauna  from  the  lower  part  of  the  Villabellaco  Limestone  has  Muensteroceras  parallelum  in 
common  with  the  Rockford  Limestone  of  Indiana,  and  Merocanites  marshallensis  and  Winchelloceras  (a 
different  but  comparable  species,  Winchelloceras  allei)  with  the  Marshall  Formation  of  Michigan.  The  latter  has 
been  assigned  an  early  Visean  age  by  Miller  and  Garner  (1955,  pp.  118-119)  (see  also  Brenckle,  Lane,  and 
Collinson  1974)  but  a late  Tournaisian  age  by  Manger  (1979,  p.  221,  fig.  3).  The  lower  part  of  the  Villabellaco 
Limestone  containing  the  goniatites  comparable  to  those  of  the  Marshall  Formation,  has  also  yielded  conodonts 
of  the  typicus  Zone  which  belongs  to  Vla  of  the  Belgian  sequence.  It  thus  appears  that  the  Marshall  Formation, 
may  be  lower  Visean  rather  than  upper  Tournaisian.  On  the  other  hand,  Kazakhstania,  another  element  of  the 
Marshall  Fauna,  is  regarded  as  a late  Tournaisian  index  (Librovitch  1940,  pp.  324-325;  Manger  1971).  This 
goniatite  does  not  occur  in  the  Villabellaco  Limestone. 

Gnathodus  bilineatus  Zone.  The  marked  reduction  of  the  texanus  group  and  its  disappearance  a little  way  into 
the  zone  marks  the  beginning  of  the  Gnathodus  bilineatus  Zone.  This  change  coincides  with  the  appearance  of  the 
important  G.  bilineatus  (Roundy)  and  Paragnathodus  commutatus  (Branson  and  Mehl).  This  abrupt  change 
is  due  to  the  absence  of  several  faunas,  notably  the  late  appearance  of  P.  commutatus  which  appears  before 
G.  bilineatus  in  northern  Europe,  and  the  absence  of  the  G.  texanus  fauna  which  is  important  in  Missouri 
(Thompson  and  Fellows  1970). 

Paragnathodus  nodosus  Zone.  The  increase  in  the  complexity  of  P.  commutatus  by  the  addition  of  platform 
nodes  marks  the  appearance  of  P.  nodosus  (Bischoff).  This  widespread  species  appears  in  the  zone  of 
Neoglyphioceras  spirale,  Go  in  Germany  (Meischner  1970)  and  in  V3b  in  Belgium  (Groessens  19776).  G. 
bilineatus  is  still  a very  important  zonal  constituent. 

Paragnathodus  multinodosus  Zone.  The  increasing  complexity  of  platform  ornamentation  marks  the 
appearance  of  P.  multinodosus  (Wirth).  Again  G.  bilineatus  is  an  important  element  of  the  fauna.  Budinger  and 
Kullmann  (1964)  related  the  appearance  of  P.  multinodosus  to  the  granosus  Zone  (Go  y).  The  consistent  later 
appearance  of  P.  multinodosus  relative  to  P.  nodosus,  often  only  1 m higher,  is  only  detected  in  closely  sampled 
sections.  The  P.  multinodosus  Zone  continues  into  the  overlying  Namurian  where  in  the  absence  of  the  later  form 
of  G.  bilineatus,  G.  bilineatus  bollandensis  and  the  rarity  of  G.  girtyi  Hass  it  has  not  proved  possible  to  subdivide 
the  early  Namurian. 

Consideration  of  the  Faunas.  The  faunas  are  dominated  by  platformed  conodonts  with  wide  basal 
cavities  and  flaring  platforms  typified  by  species  of  Gnathodus  and  Paragnathodus.  In  the  Visean  these 
are  characteristic  of  deep-water  cephalopod  limestones  in  many  parts  of  the  world  and  the  deep- 
water nature  of  the  Visean  griotte  limestones  and  the  presence  of  goniatites  confirms  this  as  a general 
observation.  Nevertheless,  both  in  the  Tournaisian  and  the  Visean  there  are  peculiarities  in  the 
faunas  which  are  worthy  of  note. 

The  Tournaisian  faunas  are  of  two  types:  one,  occurring  in  the  crystalline  limestone  of  the  Baleas 
Formation  and  the  basal  unit  of  the  Genicera  Formation,  is  dominated  by  gnathodids  whereas  the 
other,  occurring  in  the  fine  grained  limestone  of  the  Ermita  Formation  and  in  the  conglomerates  and 
shales  of  the  Vegamian  Formation,  is  dominated  by  siphonodellids  and  polygnathids.  Meischner 
(1970)  suggested  that  the  gnathodid  faunas  of  the  German  anchor alis  Zone  were  characteristic  of 
schwellen  environments  whereas  the  polygnathids  dominate  in  the  basin  environments.  Matthews, 
Sadler,  and  Sellwood  (1972)  suggested  that  the  contrast  is  between  faunas  with  large  basal  cavities 
and  flaring  platforms,  the  gnathodids,  and  those  with  small  basal  cavities,  the  polygnathids.  The 
distribution  of  the  Baleas  Formation  in  the  Porma-Bernesga  region  is  in  the  form  of  an  east-west 
trending  area  within  the  Vegamian  shale  outcrop  and  it  was  suggested  by  Higgins  et  al.  (1964)  that  it 
was  formed  on  a ridge,  in  a deeper-water  sea.  The  conodont  faunas  would  confirm  this  view,  because 
the  Vegamian  faunas,  although  few,  are  dominated  by  polygnathid  conodonts.  Whether  the 


322 


PALAEONTOLOGY,  VOLUME  25 


text-fig.  5.  Conodont  zones  in  the  sections  at  Entrago  and  Robledo  de  Caldas. 


HIGGINS  AND  WAGNER-GENTIS:  CONODONTS  AND  GON I ATITES 


323 


text-fig.  6.  Generalized  geological  map  of  part  of  northern  Leon  after  Wagner  et  al.  (1971)  with  structural  units  after  de  Sitter  (1962). 


324 


PALAEONTOLOGY,  VOLUME  25 


crystalline  limestone  faunas  of  the  Baleas  were  formed  in  the  same  depth  of  water  as  the  griotte 
limestone  is  unknown  but  it  seems  likely  in  view  of  the  similarity  of  conodont  type. 

The  other  anomaly  occurs  in  the  Visean  where  the  rarity  of  G.  girtyi  contrasts  with  its  abundance  in 
the  basin  limestones  and  shales  of  northern  Europe.  This  species  is  occasionally  found  in  the  G. 
bilineatus  Zone  but  principally  occurs  in  the  P.  nodosus  Zone.  In  the  Pyrenees  (Marks  and  Wensink 
1970)  it  appears  to  occur  only  in  the  P.  nodosus  Zone.  There  is  no  record  of  G.  girtyi  in  the  Namurian 
of  Spain.  Similarly,  the  absence  of  G.  bilineatus  bollandensis,  the  high  abundance  of  Spathognathodus 
campbelli  Rexroad,  and  the  almost  total  restriction  of  P.  multinodosus  to  the  Cantabrian  Mountains 
and  the  Pyrenees,  point  to  the  partial  geographic  isolation  of  this  area  during  the  early  Namurian 

(Ej,  E2). 

IMPLICATIONS  OF  THE  CONODONT  DATINGS  FOR  THE  HISTORY  OF 

THE  AREA 

Comparison  of  the  conodont  sequence  with  more  complete  ones  in  Belgium,  Germany,  and  Great  Britain 
indicates  the  presence  of  considerable  breaks  in  the  Lower  Carboniferous  succession  in  the  Cantabrian 
Mountains.  However,  there  is  little  evidence  of  extensive  erosion  and  the  lack  of  faunal  reworking  and  the  high 
faunal  abundance  implies  slow  but  continuous  deposition  separated  by  periods  of  non-deposition.  The  sequence 
of  events  is  as  follows: 


NAMURIAN  (El5  E2) 


VISEAN 


multinodosus  Zone 
nodosus  Zone 
bilineatus  Zone 


typicus  Zone 


anchoralis  Zone 


tournaisian  cooperi-communis  Zone 

Upper  Protognathodus  f. 


DEVONIAN 


costatus  Zone 


Average  number  of 
specimensjKg 


Continuous  deposition  ] 39 

l 27 

Widespread  non-deposition 

Widespread  deposition  19 

Local  non-deposition 

Major  transgression  97 

Widespread  non-deposition, 

some  erosion.  Continuation 

of  swell  topography  during  early 

part  of  the  zone 

Development  of  swells  89 

Widespread  non-deposition 

Patchy  distribution 

Erosion  surface  in  most  areas 

Widespread  transgression 


Late  Devonian  history.  Adrichem  Boogaert  (1967)  clearly  demonstrated  the  transgressive  nature  of 
the  late  Devonian  rocks  which,  in  the  form  of  the  Ermita  Formation,  commonly  rest  unconformably 
on  the  underlying  rocks. 

Tournaisian  history.  There  are  undoubtedly  sections  in  the  Cantabrian  Mountains  which  show  a 
transition  from  the  Devonian  into  the  Lower  Carboniferous.  The  sections  at  Santiago  de  las  Villas 
and  Olleros  de  Alba  in  the  Porma-Bernesga  area  yield  the  earliest  Carboniferous,  Upper 
Protognathodus,  fauna  in  a sequence  which  is  uninterrupted  from  the  Devonian.  Unfortunately,  the 
latest  Devonian  rocks  do  not  yield  conodont  faunas.  A transition  probably  also  exists  in  the  Vidrieros 
Formation  in  the  northern  part  of  the  Palentine  Basin  (Adrichem  Boogaert  1967)  where  the  costatus 
Zone  is  followed  by  the  Protognathodus  fauna,  but  a full  sequence  of  zones  has  not  yet  been 
demonstrated.  Similarly,  there  may  well  be  a transition  in  the  Candamo  Formation  in  Loredo,  but 
again  a full  sequence  of  zones  has  not  been  proved  (Del  Rio  and  Menendez  Alvarez  1978).  These  are 


HIGGINS  AND  WAGNER-GENTIS:  CONODONTS  AND  GONI ATITES  325 


text-fig.  7.  Correlation  of  the  conodont  sequence  in  the  Cantabrian  Mountains  with  those  of  Belgium  and  Germany.  The  suggested  faunal  and 
stratigraphic  breaks  in  the  Spanish  sequence  are  indicated  by  diagonal  lines. 


CANTABRIAN  MOUNTAINS 

(after  Groessens  1975, 
Higgins  & Bouckaert  1968) 

GERMANY 

(after  Meischner  1970, 
Conil  & Paproth  1968) 

SIPHONODELLA 

ZONATION 

(after  Sandberg  et  al.  1978) 

Pa.  multinodosus 
Pa.  nodo8u8 

e2 

G.  bilineatus  bollandensis 

S2 

G.  bilineatus  schmidti 

isosticha-U . crenulata 

\\\\\\\\\\\\\N 

E! 

Pa.  nodosus 

v3c 

v3bY 

Pa.  nodoeu8 

Go 

Gos 

G.  bilineatus  bilineatue 
G.  typicus 

V3ba 

G.  bilineatus  bilineatus 

% 

G.  bilineatus  bilineatus 

VV3« 

- “ 

- 

anchoralis-bilineatus 

Vlb 

Via 

M.  beckmanni 

Pa.  cormrutatus 
G.  homopunctatus 

mm 

S.  anchoralis 

-3 

S.  anchoralis 

S.  anchoralis 
E.  burling tonensis 
D.  lotus 

S.  anchoralis 
S.  crenulata 

mms. 

S.  oooperi-P.  communis 

mm 

P.  communis  carina 

Do.  bouckaerti 
Sp.  bultyncki 

Siphonodella 

G, 

Si.  triangulus  triangulue 

sandbergi 

Si.  tTiawulus  ivaequalis 

u.  duplicata 
L.  duplicata 

kocke li-denti lineatua 

sulcata 

u.  Pro tognathodu8  fauna 

U.  Protognathodus  fauna 

ccstatu* 

Pro togna thodu 8 fauna 

L.  Protognathodus  fauna 
M.  & u.  costatus 

praesulcata 

text-fig.  7.  Correlation  of  the  conodont  sequence  in  the  Cantabrian  Mountains  with  those  of  Belgium  and  Germany.  The  suggested  faunal  and 
stratigraphic  breaks  in  the  Spanish  sequence  are  indicated  by  diagonal  lines. 


326 


PALAEONTOLOGY,  VOLUME  25 


the  exceptions,  however,  for  the  Devonian-Carboniferous  boundary  is  usually  an  erosion  surface 
representing  a regression  of  early  Carboniferous  age.  In  the  Porma-Bernesga  area  the  gap  at  the  base 
of  the  Carboniferous  increases  in  size  in  a northerly  direction  from  Santiago  de  las  Villas  to 
Viadangos  de  Arbas  where  the  Visean  rests  directly  on  the  late  Devonian  (see  text-fig.  8).  This  trend  is 
also  seen  in  the  pre-Ermita  unconformity  in  this  area  (Adrichem  Boogaert  1967)  where  it  was  inter- 
preted as  part  of  a general  trend  of  increasing  hiatus  in  a northerly  and  easterly  direction  towards  the 
‘Asturian  Geanticline’  which  occupies  the  central  part  of  the  Cantabrians.  The  same  trend  towards 
this  platform  area  is  evident  in  the  Lower  Carboniferous,  but  is  one  of  increasing  disconformity 
rather  than  unconformity  and  is  an  expression  of  the  Bretonic  Movements. 

These  early  Tournaisian  (Tnlb)  deposits  represent  the  termination  of  the  late  Devonian 
transgression  and  were  followed  by  a long  period  of  regression.  Evidence  of  the  upper  part  of  the 
sulcata  Zone  and  the  following  zones  up  to  the  upper  part  of  the  sandbergi  Zone  of  Sandberg  et  al. 
(1978)  is  lacking.  Comparison  with  Belgium  (Groessens  19776)  reveals  the  absence  of  all  but  the 
highest  part  of  the  Siphonodella  Zone  or  lower  part  of  the  Polygnathus  communis  carina  Zone  which 
implies  a gap  ranging  from  Tnlb  to  Tn3a.  This  regression  is  represented  in  all  sections  of  the 
Porma-Bernesga  area  by  an  erosion  surface.  The  subsequent  transgression  resulted  in  a different 
pattern  of  sedimentation  due  in  part  to  the  emergence  of  positive  areas  in  a generally  subsiding  basin. 
Dominating  the  facies  pattern  is  the  Vegamian  Formation,  a thin  sequence  of  black,  often  cherty, 
shales  with  thin  conglomeratic  and  phosphatic  nodule  horizons.  In  the  Porma-Bernesga  area  a 
conglomeratic  sandstone  at  the  base  of  the  Vegamian  Formation  at  Santiago  de  las  Villas  and 
Genicera  yields  a fauna  of  polygnathids,  siphonodellids,  and  pseudopolygnathids  belonging  to  the 
cooperi-communis  Zone.  The  upper  limit  of  the  Formation  is  more  difficult  to  define.  The  highest  beds 
of  the  Genicera  section  yield  Pseudopoly gnathus  triangulus  pinnata  Voges,  which  ranges  through  Tn3 
in  Belgium  and  in  other  sections  it  is  overlain  by  beds  of  anchoralis  (Tn3c)  age.  A pr e-anchoralis  age 
for  the  major  part  of  the  Vegamian  seems  likely,  although  an  early  anchoralis  age  for  the  highest  part 
cannot  be  ruled  out.  To  the  east  and  north-east  of  the  Porma-Bernesga  area,  the  base  of  the 
Vegamian  appears  to  be  of  cooperi-communis  age  (Adrichem  Boogaert  1967)  where  the  faunas  were 
referred  to  the  lower  anchoralis  zone.  The  upper  surface  appears  to  be  diachronous  since  Budinger 
and  Kullmann  (1964)  recovered  late  Visean  (GoIII  a/j3)  goniatites  from  the  highest  part  of  the 
Formation,  although  Wagner  et  al.  (1971)  question  the  attribution  of  these  faunas  to  the  Vegamian 
Formation. 

In  the  Porma-Bernesga  area,  along  a zone  extending  from  Aviados  to  Pola  de  Gordon,  the 
Vegamian  black  shales  are  absent  and  their  place  is  taken  by  the  crystalline  limestones  of  the  Baleas 
Formation.  This  thin  unit,  12  m thick,  is  rich  in  conodonts  with  an  average  of  eighty-nine  conodonts 
per  kilogram,  indicating  slow  sedimentation.  The  faunas  are  of  the  swell  type  and  indicate  the 
presence  of  a swell  extending  east-west  through  the  Porma-Bernesga  area,  which  is  here  named  the 
Pola  de  Gordon  Swell.  In  the  area  to  the  east,  north  of  Sabero,  a similar  crystalline  limestone  again 
replaces  the  Vegamian  Shales  and  is  also  of  cooperi-communis  age  (Adrichem  Boogaert  1967).  It  may 
represent  the  easterly  continuation  of  the  Pola  de  Gordon  Swell.  At  Entrago,  Budinger  and 
Kullmann  ( 1 964)  recorded  from  the  lower  part  of  the  sequence  a crystalline  limestone  of  ‘Baleas’  type 
also  yielding  a mid-late  Tournaisian  conodont  fauna.  As  in  the  area  to  the  south,  the  Vegamian  is 
absent  and  it  seems  probable  that  this  marks  the  position  of  another  swell.  These  swells  are  close  to, 
and  parallel  with,  the  early  Carboniferous  coastline  (see  Adrichem  Boogaert  1967)  and  represent 
crinoidal  banks  forming  on  ridges  close  to  the  shore-line. 

The  anchoralis  Zone,  although  extremely  thin  (c.  2 m)  is  the  most  widespread  horizon  in  the 
Tournaisian  of  the  Cantabrian  Mountains  and  represents  a considerable  extension  of  the  sea.  In  the 
majority  of  sections  the  base  of  the  zone  is  an  erosion  surface,  but  one  section,  at  Genicera,  exhibits  a 
transition  from  the  Vegamian  Formation  into  the  overlying  Genicera  Formation.  The  earliest 
deposits  of  the  zone  included  the  top  1 m of  the  Baleas  Formation  in  the  Pola  de  Gordon  area  where 
erosion  surfaces  at  the  base  and  top  separate  the  unit  from  cooperi-communis  Zone  below  and  the 
later  part  of  the  anchoralis  Zone  above.  The  most  widespread  part  of  the  zone  occurs  at  the  base  of  the 
Genicera  Formation  where  it  occupies  the  basal  metre  of  the  red  and  grey  nodular  limestones 


HIGGINS  AND  WAGNER-GENTIS:  CONODONTS  AND  GONIATITES 


327 


text-fig.  8.  Geographical  location  and  conodont  sequences  of  the  sections  on  the  Porma-Bernesga  area.  The  units 
are  structural  ones,  usually  isoclinal  synclines  separated  by  thrusts.  The  inset  ribbon  diagram  illustrates  the 

distribution  of  the  zones. 


328 


PALAEONTOLOGY,  VOLUME  25 


(griotte).  The  fauna  commonly  occurs  in  a fine  grained  non-nodular  limestone  and  is  of  the  ‘swell’ 
type  being  composed  of  gnathodids  together  with  some  locally  abundant  polygnathids  and 
pseudopolygnathids.  However,  there  are  no  swells  identifiable  in  this  zone,  merely  a general 
shallowing  over  the  whole  region  except  in  the  extreme  east  where  the  Vegamian  Formation  persists. 
In  the  south-east,  the  Revilla  Nappe,  the  Vegamian  Formation  is  absent  and  the  oldest 
Carboniferous  is  a thin  sandy  limestone  of  anchor alis  age  representing  the  southerly  extension  of  the 
Lower  Carboniferous  sea. 

Visean  history.  As  already  stated,  it  is  believed  that  the  anchoralis  Zone  is  incomplete,  only  the  lower 
part  being  present.  There  is,  however,  no  obvious  physical  expression  of  this  break  in  the  rock 
successions  except  the  occasional  absence  of  the  anchoralis  faunas  as  at  Viadangos  de  Arbas  and 
possibly  at  Gildar-Monto  where  Budinger  and  Kullmann  (1964)  date  the  base  of  the  Carboniferous 
as  late  Visean. 

The  Visean  begins  with  the  typicus  Zone,  less  than  5 m thick,  and  with  a much-reduced  conodont 
abundance  of  nineteen  specimens  per  kilogram  compared  to  ninety-seven  per  kilogram  in  the 
anchoralis  Zone,  indicating  more  rapid  deposition.  At  the  top  of  the  zone  the  nodular  limestones  give 
way  to  horizons  and  nodules  of  radiolarian  cherts  in  red  shales,  which  are  extensively  developed  in 
the  Porma-Bernesga-Esla  area. 

Immediately  above  the  chert  the  fauna  changes  markedly  with  the  incoming  of  Gnathodus 
bilineatus  and  the  change  reflects  the  absence  of  at  least  one  fauna.  In  comparison  with  Belgium  and 
Ireland  Paragnathodus  commutatus  appears  later  in  Spain,  both  in  the  Cantabrian  Mountains  and  in 
the  Pyrenees  (Marks  and  Wensink  1970),  by  at  least  one  zone.  Also  absent  is  the  G.  bulbosus  fauna  of 
mid-Osage  age  (Thompson  and  Fellows  1970).  It  is  possible  that  these  zones  are  represented  by  the 
chert  horizon  but  in  chert-free  sections  the  faunal  break  is  still  present.  This  non-sequence  is  clearly  of 
more  than  local  significance. 

Above  the  chert  the  sequence  appears  to  be  unbroken,  although  still  condensed,  and  both  the 
goniatite  and  the  conodont  records  are  complete. 

SYSTEMATIC  PALAEONTOLOGY 
(A.  C.  Higgins) 

CONODONTS 

Where  specimens  can  be  related  to  natural  assemblages  they  are  referred  to  multi-element  species  such  as 
Gnathodus  bilineatus  and  Idioprioniodus  conjunctus.  Where  this  is  not  possible  specimens  are  referred  to  disjunct 
species.  The  non-platform  elements  in  the  Spanish  sections  are  too  disproportionately  represented,  either  due  to 
breakage  or  selective  sorting,  to  allow  the  construction  of  multi-element  species  by  means  of  similarity  of  ratios 
and  ranges. 

Type  and  figured  specimens  are  housed  in  the  micropalaeontology  collection,  Department  of  Geology, 
University  of  Sheffield. 

Multi-element  species 

Genus  gnathodus  Pander,  1856 
Type  species.  Gnathodus  mosquensis  Pander  1856. 

Gnathodus  bilineatus  bilineatus  (Roundy)  1926 

1926  Polygnathus  bilineata  Roundy,  p.  13,  pi.  3,  fig.  10 a-c. 

1964  Gnathodus  bilineatus  (Roundy),  Schmidt  and  Muller  1964,  p.  1 14,  figs.  7,  8. 

P element 
Plate  34,  figs.  1,  3 

For  a recent  synonymy  see  Higgins  1975,  p.  28. 


HIGGINS  AND  WAGNER-GENTIS:  CONODONTS  AND  GONI ATITES 


329 


Remarks.  In  the  Gnathodus  bilineatus  Zone  two  morphotypes  of  this  species  occur.  One  is  the  typical 
form  of  G.  bilineatus  with  an  inner  platform  consisting  of  a transversely  ribbed  ridge  extending  to  the 
posterior  end  of  the  platform  whereas  the  other  has  a short  inner  platform  which  does  not  extend  to 
the  posterior  end.  The  latter  morphotype  was  referred  to  G.  delicatus  by  Adrichem  Boogaert  (1967) 
and  to  G.  bilineatus  subsp.  nov.  by  Higgins  (1974).  Although  it  bears  some  similarities  to  G.  delicatus 
it  would  considerably  extend  the  range  of  that  species,  and  its  occurrence  is  separated  from  that 
species  by  a stratigraphic  gap.  It  is  probably  better  regarded  as  a variant  of  G.  bilineatus  rather  than  a 
relative  of  G.  delicatus. 


O Element 
Plate  34,  fig.  19 

1932  Bryantodus  delicatus  Stauffer  and  Plummer,  p.  29,  pi.  2,  fig.  27. 

1941  Ozarkodina  delicatula  (Stauffer  and  Plummer),  Ellison,  p.  120,  pi.  20,  figs.  40-42,  47. 

For  a recent  synonymy  see  Higgins  1975,  p.  69. 

Remarks.  The  denticles  of  the  O element  and  the  other  non-platform  elements  of  Gnathodus  bilineatus 
are  ornamented  with  marked  striations  similar  to  those  of  the  Oz  element  of  Idiognathodus  delicatus 
illustrated  by  von  Bitter  (1972).  They  consist  of  striations,  1 to  2 microns  thick,  rounded  and 
sometimes  continuous  but  sinuous  and  in  other  instances  thickening  and  thinning  down  the  length  of 
the  denticle.  This  pattern  has  been  described  by  Norby  (1975). 


N Element 
Plate  34,  fig.  25 

1933  Synprioniodina  sp.  Gunnell,  p.  296,  pi.  31,  fig.  6. 

1941  Synprioniodina  microdenta  Ellison,  pp.  108-111,  119,  pi.  120,  figs.  43-46. 
For  a recent  synonymy  see  Higgins  1975,  pp.  38,  39. 


Ala+b  Elements 
Plate  34,  figs.  20,  22 

1957  Hindeodella  ibergensis  Bischoff,  p.  28,  pi.  8,  figs.  33,  37,  39. 

For  a recent  synonymy  see  Higgins  1975,  pp.  38,  39. 

Remarks.  Baesemann  (1973)  provided  a complete  description  of  these  two  elements.  The  Ala 
element,  typically  with  forwardly  inclined  denticles  on  the  anterior  process  is  illustrated  extensively 
by  Higgins  (1975,  p.  39,  fig.  8a,  c,  d,f).  The  Alb  element,  with  strongly  inwardly  flexed  denticles  on  the 
anterior  process  is  illustrated  by  Higgins  (1975,  p.  39,  fig.  8 b,  e). 


Alc  Element 

1959  Hindeodina  uncata  Hass,  p.  383,  pi.  47,  fig.  6. 

A recent  synonymy  is  given  in  Higgins  (1975,  p.  44). 

Remarks.  Although  this  element  was  not  included  by  Baesemann  (1973)  or  von  Bitter  (1972)  in 
Idiognathodus  delicatus  it  is  commonly  present  in  Lower  and  Upper  Carboniferous  faunas  in 
association  with  gnathodid  and  idiognathodid  elements  and  it  was  recorded  from  such  assemblages 
by  Schmidt  and  Muller  (1964,  p.  1 17,  fig.  9 (6)). 


330 


PALAEONTOLOGY,  VOLUME  25 


A 2 Element 
Plate  34,  fig.  24 

1957  Angulodus  walrathi  Bischoff,  p.  17,  pi.  5,  figs.  44,  45. 

1975  Hindeodella  simplex  (Higgins  and  Bouckaert),  Higgins  1975,  pi.  5,  figs.  10,  12,  13. 

A recent  synonymy  is  given  in  Higgins  1975,  pi.  42. 

Remarks.  A complete  description  is  given  in  Higgins  and  Bouckaert  (1968,  pp.  28, 29)  and  Baesemann 
(1973,  p.  704). 

A 3 Element 
Plate  34,  fig.  26 

1965  Hibbardella  acuta  Murray  and  Chronic,  p.  598,  pi.  73,  figs.  3-5. 

A recent  synonymy  is  given  in  Higgins  1975,  p.  34. 


EXPLANATION  OF  PLATE  34 
All  specimens  x 40 

Gnathodus  bilineatus  bilineatus  (Roundy) 

Figs.  1,  3.  P Element,  Villabellaco  Section,  sample  2902,  oral  views  of  2902  (1  and  2). 

Fig.  19.  O Element,  Olleros  de  Alba  Section,  sample  OLI,  inner  lateral  views  of  OLI  (1). 

Fig.  20.  Aib  Element,  Entrago  Section,  sample  1149C,  inner  lateral  view  of  1 149C  (1). 

Fig.  22.  Ala  Element,  Entrago  Section,  sample  1 149K,  inner  lateral  view  of  1 149K  (1). 

Fig.  24.  A 2 Element,  Matallana  Section,  sample  1060A,  inner  lateral  view  of  1060A  (1). 

Fig.  25.  N Element,  Entrago  Section,  sample  1 149C,  inner  lateral  view  of  1 149C  (2). 

Fig.  26.  A3  Element,  Entrago  Section,  sample  1 149K,  inner  lateral  view  of  1 149K  (2). 

Fig.  2.  Gnathodus  delicatus  Branson  and  Mehl,  Pola  de  Gordon  Section,  sample  3085,  oral  view  of  3085  (1). 
Figs.  4,  14.  Gnathodus  cuneiformis  Mehl  and  Thomas.  4,  Baleas  Quarry,  Sample  1264,  oral  view  of  1264  (18). 

14,  Olleros  de  Alba  Section,  sample  OLV,  oral  view  of  OLV  (1). 

Fig.  5.  Gnathodus  typicus  Cooper.  Olleros  de  Alba  Section,  sample  OLIV,  oral  view  of  OLIV  (1). 

Fig.  6.  Gnathodus  semiglaber  Bischoff.  Baleas  Quarry,  sample  1264,  oral  view  of  1264  (19). 

Fig.  7.  Gnathodus  homopunctatus  Bischoff.  Revilla  Section,  sample  134V  oral  view  of  134V  (1). 

Fig.  8.  Protognathodus  collinsoni  Ziegler.  Santiago  de  las  Villas  Section,  sample  1310,  oral  view  of  1310  (1). 

Fig.  9.  Gnathodus  girtyi  girtyi  Hass.  Revilla  Section,  sample  134IX,  oral  view  of  134X  (1). 

Fig.  10.  Protognathodus  meischneri  Ziegler  1969.  Santiago  de  las  Villas  Section,  sample  1310,  oral  view  of  1310. 
Fig.  11.  Gnathodus  texanus  pseudosemiglaber  Thompson  and  Fellows  1970,  Matallana  Section,  sample  1060A, 
oral  view  of  1060A  (2). 

Figs.  12,  13,  15.  Paragnathodus  multinodosus  (Wirth).  Revilla  Section,  sample  134IX,  oral  view  of  134IX  (2-4). 
Figs.  16,  17.  Mestognathus  beckmanni  Bischoff.  Villabellaco  Section,  sample  2910,  oral  and  aboral  views  of 
2910(1). 

Idioprioniodus  conjunctus  (Gunnell) 

Fig.  18.  B3a  Element,  Entrago  Section,  sample  1 149K,  posterior  view  of  1 149K  (3). 

Fig.  21.  Nj  Element,  Olleros  de  Alba  Section,  sample  OLI,  outer  lateral  view  of  OLI  (2). 

Fig.  23.  Blb  Element,  Entrago  Section,  sample  1 149C,  inner  lateral  view  of  1 149C  (3). 

Fig.  27.  N2  Element,  Matallana  Section,  sample  1060A,  outer  lateral  view  of  1060A  (3). 

Fig.  28.  Bla  Element,  Matallana  Section,  sample  1060A,  inner  lateral  view  of  1060A  (4). 

Fig.  29.  B3b  Element,  Matallana  Section,  sample  1060A,  outer  lateral  view  of  1060A  (5). 

Fig.  30.  Scaliognathus  sp.  nov.  Olleros  de  Alba  section,  sample  1340,  oral  view  of  specimen  1340  (1),  x60. 

Fig.  31.  Scaliognathus  sp.  nov.  Olleros  de  Alba  section,  sample  OLI,  oral  view  of  specimen  OLI  (5),  x 60. 

Fig.  32.  Scaliognathus  sp.  nov.  Olleros  de  Alba  section,  sample  OLI,  oral  view  of  specimen  OLI  (4),  x 60. 


PLATE  34 


HIGGINS  and  WAGNER-GENTIS,  Earlier  Carboniferous  conodonts 


332 


PALAEONTOLOGY,  VOLUME  25 
Genus  idioprioniodus  Gunnell  1933 


1933  Idioprioniodus  Gunnell,  p.  265. 

1952  Duboisella  Rhodes,  p.  895. 

1972  Neoprioniodus  Rhodes  and  Muller,  von  Bitter,  p.  68. 

1973  Idioprioniodus  Gunnell,  Baesemann,  p.  703. 

1974  Idioprioniodus  Gunnell,  Merrill  and  Merrill,  pp.  1 19-130. 

Type  species.  Idioprioniodus  typus  Gunnell,  1933. 

Idioprioniodus  conjunctus  (Gunnell)  1931 

1931  Prioniodus  conjunctus  Gunnell,  p.  247,  pi.  29,  fig.  7. 

1974  Idioprioniodus  conjunctus  (Gunnell),  Merrill  and  Merrill,  p.  120. 

Remarks.  Merrill  and  Merrill  (1974)  proposed  this  species  for  pre-Missourian  representatives  of 
Idioprioniodus  which  would  be  assigned  to  Idioprioniodus  typus  (Gunnell)  except  for  the  presence  of 
an  N2  (metalonchodinid)  element. 


Nx  Element 
Plate  34,  fig.  21 

1931  Prioniodus  conjunctus  Gunnell,  p.  247,  pi.  29,  fig.  7. 

A complete  synonymy  is  given  in  Higgins  1975,  p.  66. 

N2  Element 
Plate  34,  fig.  27 

1931  Prioniodus  bidentatus  Gunnell,  p.  247,  pi.  29,  fig.  6. 

1941  Metalonchodina  bidentata  (Gunnell),  Branson  and  Mehl,  p.  106,  pi.  19,  fig.  34. 
A complete  synonymy  is  given  in  Higgins  1975,  p.  63. 


Bla  Element 
Plate  34,  fig.  28 

1933  Idioprioniodus  typus  Gunnell,  p.  265,  pi.  31,  fig.  47. 

1941  Ligonodina  typa  (Gunnell),  Ellison,  p.  1 14,  pi.  20,  figs.  8-11. 

1953  Ligonodina  roundyi  Hass,  p.  82,  pi.  15,  figs.  5-9. 

1972  Neoprioniodus  conjunctus  (Gunnell),  Von  Bitter,  p.  69,  pi.  12,  fig.  3,  Hi  element. 

1973  Idioprioniodus  lexingtonensis  (Gunnell),  Baesemann,  p.  703,  pi.  3,  fig.  1. 

Remarks.  Higgins  (1975)  regarded  Ligonodina  roundyi  and  L.  typa  as  separate  species  because  the 
former  species  has  discrete  anterior  process  denticles.  However,  both  species  have  the  same  range  and 
there  are  transitional  specimens  throughout  this  range. 

Blt)  Element 
Plate  34,  fig.  23 

1931  Prioniodus  clarki  Gunnell,  p.  247,  pi.  29,  fig.  8. 

1941  Lonchodina  clarki  (Gunnell),  Ellison,  p.  116,  pi.  20,  figs.  21,  27,  30,  31. 

1957  Lonchodina  cf.  projecta  Ulrich  and  Bassler,  Bischoff,  p.  34,  pi.  1,  fig.  20. 

1961  Lonchodina  cf.  projecta  Ulrich  and  Bassler,  Higgins,  pi.  11,  fig.  10. 

1968  Lonchodina  bischoffi  Higgins  and  Bouckaert,  p.  43. 

1972  Neoprioniodus  conjunctus  (Gunnell),  Von  Bitter,  p.  69,  pi.  12,  fig.  4 a-c,  PI.  element. 


HIGGINS  AND  WAGNER-GENTIS:  CONODONTS  AND  GONIATITES 


333 


1973  Idioprioniodus  lexingtonensis  (Gunnell),  Baesemann,  p.  704,  pi.  3,  fig.  2,  BU)  element. 

1975  Lonchodina  bischoffi  Higgins  and  Bouckaert,  Higgins,  p.  59,  pi.  2,  figs.  1-4,  8. 

For  a complete  description  see  Higgins  and  Bouckaert  (1968),  p.  43. 

Remarks.  None  of  the  illustrated  specimens  of  Lonchodina  clarki  from  the  Pennsylvanian  are  as 
complete  as  the  specimens  assigned  to  L.  bischoffi  which  have  been  described  from  the  Visean  and 
Namurian.  Both  the  anterior  process  and  the  posterior  process  of  the  former  species  are  incomplete. 
None  the  less,  in  all  other  respects  the  two  species  are  identical  and  there  seems  little  justification  in 
retaining  them  as  separate  species. 


B2  Element 

1931  Prioniodus  lexingtonensis  Gunnell,  p.  246,  pi.  29,  fig.  4. 

1941  Ligonodina  lexingtonensis  (Gunnell),  Ellison,  p.  115,  pi.  20,  figs.  13-15. 
1973  Idioprioniodus  lexingtonensis  (Gunnell),  Baesemann,  p.  704,  pi.  3,  figs.  3,  8. 

For  a more  complete  synonymy  and  a description  see  Higgins  (1975),  p.  58. 


B3a  Element 
Plate  34,  fig.  18 

1931  Prioniodus  subacodus  Gunnell,  p.  246,  pi.  29,  fig.  5. 

1941  Hibbardella  subacoda  (Gunnell),  Ellison,  p.  118,  pi.  20,  figs.  22,  26. 

1953  Roundya  barnettana  Hass,  p.  89,  pi.  16,  figs.  8,  9. 

1958  Roundya  costata  Rexroad,  p.  26,  pi.  2,  figs.  5-8. 

1972  Neoprioniodus  conjunctus  (Gunnell),  Von  Bitter,  p.  70,  pi.  16,  figs.  2a,  b,  Tr  element. 

1973  Idioprioniodus  lexingtonensis  (Gunnell),  Baesemann,  p.  704,  pi.  3,  fig.  9,B3a  element. 

Remarks.  Specimens  from  the  Missourian  illustrated  by  Baesemann  (1973)  and  the  late  Pennsyl- 
vanian ilustrated  by  Von  Bitter  (1972)  and  referred  to  Roundya  subacoda  are  indistinguishable  from 
early  Carboniferous  specimens  of  R.  barnettana.  The  distinction  between  the  two  species  is  the 
massivity  of  the  unit  of  R.  subacoda  compared  to  R.  barnettana.  Comparison  between  them  is  made 
difficult  by  the  incomplete  nature  of  the  later  Pennsylvanian  specimens  but  the  distinction  would 
seem  to  be  of  superficial  importance. 


B3b  Element 
Plate  34,  fig.  29 

1941  Lonchodina  ?ponderosa  Ellison,  p.  116,  pi.  20,  figs.  37-39. 

1958  Lonchodina  paraclaviger  Rexroad,  p.  22,  pi.  4,  figs.  7-10. 

1973  Idioprioniodus  lexingtonensis  (Gunnell),  Baesemann,  p.  704,  pi.  3,  figs.  4,  5,  B3b  element. 

1975  Lonchodina  paraclaviger  Rexroad,  Higgins,  p.  60,  pi.  2,  fig.  9. 

Remarks.  The  similarity  between  Lonchodina  ponderosa  and  L.  paraclaviger  is  illustrated  by  Higgins 
1975,  pi.  2,  figs.  9, 1 1 . Both  species  have  processes  almost  in  the  same  plane,  subequal  denticles  on  the 
anterior  process  which  are  approximately  equal  in  size  to  the  cusp,  and  a similar  basal  cavity.  The 
subsymmetrical  nature  of  the  unit  suggests  that  this  is  a B3  element  as  suggested  by  Baesemann  rather 
than  a B2  element. 

Disjunct  elements 

Genus  doliognathus  Branson  and  Mehl,  1941 
Type  species.  Doliognathus  latus  Branson  and  Mehl,  1941. 


334  PALAEONTOLOGY,  VOLUME  25 

Doliognathus  latus  Branson  and  Mehl,  1941 
For  synonymy  up  to  1967,  see  Thompson  (1967). 

1970  Doliognathus  latus  Branson  and  Mehl,  Thompson  and  Fellows,  p.  45. 

1971  Doliognathus  latus  Branson  and  Mehl,  Higgins,  pi.  2,  figs.  1,  3-6,  8. 

1971  Doliognathus  cf.  latus  Branson  and  Mehl,  Higgins,  pi.  2,  figs.  2,  7. 

Remarks.  The  considerable  variation  in  this  species  was  illustrated  by  Voges  (1959,  p.  274)  and 
Higgins  (1971,  pi.  2).  The  typical  form,  illustrated  by  fig.  1 of  pi.  2 (Higgins  1971)  has  a transversely 
ribbed  or  noded  platform,  smoothly  curved  margins,  and  the  outer  lateral  process  is  approximately  at 
right  angles  to  the  main  axis  of  the  unit.  The  main  variant  (pi.  2,  figs.  2 and  7)  has  more  pronounced 
platform  ribs,  strongly  irregular  platform  margins,  and  the  outer  lateral  process  is  more  strongly 
curved  or  directed  posteriorly.  This  latter  form  was  referred  to  Doliognathus  cf.  latus  by  Higgins 
(1971). 


Genus  gnathodus  Pander,  1956 
Type  species.  Gnathodus  mosquensis  Pander,  1956. 

Gnathodus  cuneiformis  Mehl  and  Thomas 
Plate  34,  figs.  4,  14 

1947  Gnathodus  cuneiformis  Mehl  and  Thomas,  p.  10,  pi.  1,  fig.  2. 

1971  Gnathodus  cf.  girtyi  Hass,  Higgins,  pi.  5,  fig.  3. 

A more  complete  synonymy  is  given  in  Thompson  and  Fellows  1970,  pp.  45,  46. 

Remarks.  Thompson  and  Fellows  (1970)  pointed  out  that  Gnathodus  cuneiformis  is  homeomorphic 
with  G.  girtyi  in  the  United  States  where  the  former  species  occurs  in  the  Osagean  and  the  latter  in  the 
Chesterian  Stage.  The  same  pattern  can  be  observed  in  Spain  where  G.  cuneiformis  occurs  in  the 
anchoralis  zone  of  late  Tournaisian  age  and  G.  girtyi  occurs  in  the  bilineatus  Zone  of  late  Visean  age. 
The  range  of  variation  of  G.  girtyi  is  much  greater  than  that  of  G.  cuneiformis  but  symmetrically 
platformed  specimens  of  both  species  are  indistinguishable. 

Gnathodus  delicatus  Branson  and  Mehl 
Plate  34,  fig.  2 

1938  Gnathodus  delicatus  Branson  and  Mehl,  p.  145,  pi.  34,  figs.  25-27. 

1971  Gnathodus  delicatus  Branson  and  Mehl,  Higgins,  pi.  5,  figs.  5,  7,  8,  11,  13. 

Recent  synonymies  have  been  given  by  Butler  (1973,  p.  497)  and  Matthews  et  al.  (1972,  p.  559). 

Remarks.  Butler  (1973)  commented  that  specimens  from  the  upper  part  of  the  range  of  this  species 
show  the  development  of  a distinct  parapet  at  the  anterior  end  of  the  inner  side  of  the  platform.  This 
form  (pi.  5,  figs.  5,  7 of  Higgins  1971)  also  occurs  in  the  Spanish  sections  where  it  overlaps  the  range  of 
Gnathodus  antetexanus,  a species  which,  as  pointed  out  by  Butler,  has  a similar  feature. 

Gnathodus  girtyi  girtyi  Hass,  1953 
Plate  34,  fig.  9 

1953  Gnathodus  girtyi  Hass,  p.  80,  pi.  14,  figs.  22-24. 

1975  Gnathodus  girtyi  girtyi  Hass,  Higgins,  p.  31,  pi.  10,  figs.  5,  6. 

A more  complete  synonymy  is  given  in  Higgins  (1975),  p.  31. 

Remarks.  Only  the  weakly  ornamented  subspecies  of  Gnathodus  girtyi  is  represented  in  the  Spanish 
succession  and  it  is  a rare  species. 


HIGGINS  AND  WAGNER-GENTIS:  CONODONTS  AND  GONI ATITES  335 

Gnathodus  typicus  Cooper,  1939 
Plate  34,  fig.  5 

1939  Gnathodus  typicus  Cooper,  p.  388,  pi.  42,  figs.  77,  78. 

1964  Gnathodus  typicus  Cooper,  Rexroad  and  Scott,  p.  31,  pi.  2,  fig.  3. 

1970  Gnathodus  typicus  Cooper,  Thompson  and  Fellows,  pp.  89,  90,  pi.  3,  figs.  3,  13. 

Remarks.  Specimens  with  a short  anteriorly  pointing  inner  platform  and  a weakly  ornamented  outer, 
wide,  and  rounded  platform  are  referred  to  this  species.  Expansion  of  the  carina  would  allow  it  to  be 
placed  in  Gnathodus  semiglaber  with  which  species  it  has  much  in  common. 

Genus  paragnathodus  Higgins,  1975 
Type  species.  Spathognathodus  commutatus  Branson  and  Mehl,  1941. 

Remarks.  Paragnathodus  commutatus  and  P.  nodosus  (Bischoff)  1957,  have  recently  been  described 
by  Higgins  1975,  pp.  70-72.  The  composition  of  the  multi-element  genus  is  unknown  but  it  is  likely  to 
correspond  to  the  natural  assemblage  Lochreia  of  Scott  1942. 

Paragnathodus  multinodosus  (Wirth),  1967 
Plate  34,  figs.  12,  13,  15;  text-fig.  106 

1962  Gnathodus  commutatus  var.  multinodosus  Higgins,  pp.  8,  9,  pi.  2,  figs.  13-18. 

1967  Gnathodus  commutatus  multinodosus  n.ssp.  Wirth,  p.  208,  pi.  19,  figs.  19,  20. 

1974  Gnathodus  commutatus  multinodosus  Higgins,  Austin  and  Husri,  pi.  2,  fig.  13. 

Discussion.  Variation  in  this  species  is  mainly  seen  in  the  shape  of  the  cup  and  its  ornamentation.  The 
cup  shape  can  vary  from  being  subsymmetrical  to  asymmetrical  where  the  inner  side  is  strongly 
folded  both  anteriorly  and  posteriorly  and  does  not  gradually  taper  to  the  posterior.  The  nodes 
typically  evenly  cover  the  platform  surface  where  they  are  situated  on  a slightly  raised  shelf.  PI.  34,  fig. 
12,  illustrates  a specimen  in  which  the  nodes  are  raised  on  two  anteriorly  directed  ridges  identical  to 
those  found  in  Paragnathodus  nodosus  and  it  may  be  a variant  of  this  species.  Scanning  photographs 
reveal  the  presence  of  micronodes  on  the  major  nodes  of  P.  multinodosus  (PI.  35,  fig.  2). 

Remarks.  The  distribution  of  this  species  is  very  restricted.  Apart  from  its  widespread  presence  in  the 
Cantabrian  Mountains  it  occurs  in  the  Pyrenees  (Wirth  1967;  Marks  and  Wensink  1970;  Perret  1974) 
and  possibly  in  Ireland  (Austin  and  Husri  1974)  but  is  unknown  elsewhere  despite  the  large 
numbers  of  faunas  of  this  age  which  have  been  described.  The  Irish  occurrence  is  abnormal  because  it 
apparently  occurs  before  P.  nodosus , whereas  in  Spain  it  appears  slightly  later. 

Genus  scaliognathus  Branson  and  Mehl,  1941 
Type  species.  Scaliognathus  anchoralis  Branson  and  Mehl,  1941. 

Discussion.  Despite  the  widespread  nature  and  stratigraphical  importance  of  this  distinctive  anchor- 
shaped genus  it  remains  monospecific.  The  Spanish  successions,  being  condensed  and  broken  by  non- 
sequences, do  not  give  a clear  pattern  of  evolution  of  the  genus.  Nevertheless,  the  material  is  well 
preserved  and  abundant,  and  it  is  possible  to  identify  the  main  variations  in  the  genus.  Three  main 
forms  are  recognized: 

1.  Forms  with  plate-like  lateral  and  anterior  processes.  These  are  referred  to  Scaliognathus 
anchoralis. 

2.  Forms  with  a plate-like  anterior  process,  but  blade-like  subequal  lateral  processes  which  are 
curved  anteriorly.  These  are  referred  to  S.  angustilateralis  sp.  nov. 

3.  Forms  with  straight  unequal  lateral  limbs  which  project  at  right  angles  to  the  anterior  process. 
The  posterior  limbs  are  blade-like.  These  are  referred  to  Scaliognathus  sp.  nov. 


336 


PALAEONTOLOGY,  VOLUME  25 


Groessens  (1977a)  suggested  an  origin  for  Scaliognathus  from  Dollymae  bouckaerti  (Groessens)  in 
which  the  first  scaliognathid  was  a slender,  highly  arched  form  with  poor  plate  development.  This 
form  would  resemble  form  2 above. 

Scaliognathus  anchoralis  Branson  and  Mehl,  1941 
Text-fig.  9a,  b , d,  g 

1941  Scaliognathus  anchoralis  Branson  and  Mehl,  p.  102,  pi.  19,  figs.  29-32. 

1964  Scaliognathus  anchoralis  Branson  and  Mehl,  Higgins  et  al.  pi.  iv,  fig.  17. 

1967  Scaliognathus  anchoralis  Branson  and  Mehl,  Adrichem  Boogaert,  p.  50,  pi.  5,  figs.  2-4,  8,  9. 

1969a  Scaliognathus  anchoralis  Branson  and  Mehl,  Matthews,  pp.  272,  273,  pi.  49,  figs.  2,  4,  8,  and  9 

only. 

19696  Scaliognathus  anchoralis  Branson  and  Mehl,  Matthews,  pi.  51,  figs.  1,  2. 

1971  Scaliognathus  anchoralis  Branson  and  Mehl,  Groessens,  pi.  1,  fig.  10  only. 

1971  Scaliognathus  anchoralis  Branson  and  Mehl,  Higgins,  pi.  3,  figs.  3,  5-7,  9;  pi.  4,  fig.  2. 

1974  Scaliognathus  anchoralis  Branson  and  Mehl,  Matthews  and  Thomas,  pi.  50,  figs.  8,  9. 

1974  Scaliognathus  anchoralis  Branson  and  Mehl,  Jenkins,  pi.  1 19,  figs.  6-9. 


text-fig.  9.  a,  Scaliognathus  anchoralis  Branson  and  Mehl.  Olleros  de  Alba  section,  sample  1340,  oral  view  of 
specimen  1340  (2),  x 60.  b , Scaliognathus  anchoralis  Branson  and  Mehl.  Baleas  Quarry,  sample  1264,  oral  view 
of  specimen  1 264  (23),  x 60.  c,/,  Scaliognathus  angustilateralis  sp.  nov.  Baleas  Quarry,  sample  1264;  fig.  c aboral 
view  of  cusp,  x 1080;  fig.  / aboral  view  of  specimen  1264  (20),  x 60.  d,  g,  Scaliognathus  anchoralis  Branson  and 
Mehl.  Olleros  de  Alba  section,  sample  OLI;  fig.  5 oral  view  of  specimen  OLI  (3),  x 60;  fig.  g detail  of  main 
denticle,  x 360.  e,  Scaliognathus  angustilateralis  sp.  nov.  Baleas  Quarry,  sample  1264,  oral  view  of  holotype, 

specimen  1264(22),  x60. 


HIGGINS  AND  WAGNER-GENTIS:  CONODONTS  AND  GONI ATITES 


337 


Diagnosis.  Paired  conodonts  with  an  anchor-like  shape  consisting  of  three  processes.  The  anterior 
process  is  wide,  tapering  to  its  anterior  extremity,  the  two  posterior  lateral  processes  are  also  wide  and 
plate-like  with  a row  of  posteriorly  inclined  denticles  on  a flat  and  horizontal  upper  surface,  which  is 
in  the  same  plane  as  the  surface  of  the  lateral  process. 

Description.  The  anterior  limb  is  broad  at  the  posterior  tapering  anteriorly  with  a convex  outer  and  a convex, 
straight  or  slightly  concave,  inner  margin.  There  is  a prominent  median  carina  consisting  of  large  discrete 
denticles  at  the  anterior  becoming  fused  into  a low  ridge  towards  the  posterior.  At  the  posterior  the  carina  is 
continued  beyond  the  extremity  of  the  platform  as  a horn-like  denticle  which  is  commonly  large  and  may  have 
a triangular  cross-section  where  the  carinal  ridge  continues  up  its  oral  face,  but  there  is  no  sign  of  this  in  the 
holotype  which  has  a small  posterior  denticle.  There  is  commonly  a sulcus  on  either  side  of  the  carina  and 
bordering  the  sulci  is  a row  of  nodes  which  may  develop  into  transverse  ridges.  The  lateral  extension  of  the 
anterior  limb  is  marked,  giving  rise  to  a plate-like  structure  with  sharp  margins  and  a flat  to  concave  upper 
surface. 

The  lateral  processes  are  as  long  or  slightly  shorter  than  the  anterior  limb  either  in  length,  denticulation,  or 
curvatures.  They  are  projected  or  curved  anteriorly  at  approximately  70°  to  the  anterior  limb  and,  if  curved, 
curvature  is  stronger  on  the  outer  limb.  They  are  as  wide  or  wider  than  the  anterior  limb  in  the  median  area  of  the 
unit  but  taper  to  a point.  The  oral  surface  is  flat  and  in  the  same  plane  as  the  oral  surface  of  the  anterior  limb  with 
which  it  is  continuous.  There  is  commonly  a row  of  low  nodes  along  its  anterior  margin.  Near,  or  at  the  posterior 
margin,  is  a row  of  posteriorly  inclined  denticles  which  may  be  subequal  in  size  or  increase  slightly  in  height 
towards  the  extremity  of  the  limbs.  The  holotype  and  para  type  have  a prominent  shelf,  with  a row  of  the 
transverse  ridges  posterior  to  the  denticle  row,  but  this  is  atypical  for  the  species  in  general  and  more  commonly 
the  shelf,  although  often  present,  is  insignificant.  The  denticles  may  be  discrete,  but  are  more  commonly  in 
contact  in  the  lower  third  of  their  length. 

The  aboral  surface  of  both  the  anterior  and  the  lateral  limbs  is  convex  and  smooth  except  for  prominent  split 
keels  which  are  grooved  along  their  length.  The  keels  meet  at  the  centre  of  the  lateral  limbs  and  a triangular  open 
pit  which  is  open  in  small  specimens  but  becomes  more  closed  in  adult  specimens.  The  anterior  face  of  the  lateral 
process  is  almost  at  right  angles  to  the  oral  face,  whereas  the  posterior  face  is  at  an  angle  of  approximately 
45  degrees. 

Discussion.  The  holotype  of  Branson  and  Mehl  (1941,  pi.  19,  figs.  30,  32)  has  very  broad  plate-like 
lateral  processes  and  a denticular  row  which  originates  near  the  midline  of  these  processes.  The 
specimen  figured  by  Jenkins  (1974,  pi.  1 19,  fig.  6)  is  close  to  the  holotype  but  few  others,  mainly  from 
Europe,  are  exactly  of  this  type.  The  description  and  diagnosis  given  above  retains  the  plate-like 
nature  of  the  processes  but  the  typical  specimen  has  a denticular  row  originating  from  the  posterior 
edge  of  the  processes  which  typifies  the  majority  of  figured  specimens.  This  redefinition  of  the  species 
would  include  form  1 and  possibly  form  2 of  Matthews  (1969a). 

Range.  Scaliognathus  anchoralis  Zone. 

No.  of  specimens  fifty. 


Scaliognathus  angustilateralis  sp.  nov. 

Text-fig.  9c,  e,f 

1967  Scaliognathus  anchoralis  Branson  and  Mehl,  Adrichem  Boogaert,  p.  185,  pi.  3,  fig.  11. 

1969  Scaliognathus  anchoralis  Branson  and  Mehl,  Matthews,  pp.  272,  273,  pi.  49,  figs.  1,  6 only. 

1971  Scaliognathus  anchoralis  Branson  and  Mehl,  Groessens,  pi.  1,  fig.  9 only. 

1971  Scaliognathus  anchoralis  Branson  and  Mehl,  Higgins,  pi.  3,  figs.  1,  2,  4,  8,  10  only. 

1973  Scaliognathus  anchoralis  Branson  and  Mehl,  Butler,  p.  510,  pi.  58,  figs.  6 and  7 only. 

Derivation  of  name.  Refers  to  the  narrowness  of  the  lateral  process. 

Holotype.  Text-fig.  9 d,  from  the  Baleas  Formation,  Baleas  Quarry,  Pola  de  Gordon.  Slide  1264(22). 

Diagnosis.  A species  of  Scaliognathus  with  narrow  lateral  processes  which  are  posteriorly  in- 
clined, curved  anteriorly,  and  in  the  same  plane  as  the  row  of  denticles  which  are  developed  on  its 
surface. 


338 


PALAEONTOLOGY,  VOLUME  25 


Description.  The  anterior  limb  is  slender,  plate-like,  with  a convex  outer  and  convex  to  concave  inner  margin 
which  are  approximately  parallel  in  the  posterior  half  but  taper  sharply  in  the  anterior  half  usually  terminating 
before  the  end  of  the  process.  The  oral  surface  of  the  limb  is  bisected  by  a median  carina  consisting  of  partially 
fused,  large  denticles  in  the  anterior  half  which  becomes  a fused,  low,  nodular  ridge  in  the  posterior  half 
extending  up  the  oral  face  of  the  horn-like  terminal  denticle.  The  margins  of  the  limb  are  slightly  crenulate 
because  of  the  development  of  a row  of  nodes  and  transverse  ridges  on  each  side.  These  are  separated  from  the 
carina  by  a shallow  sulcus. 

The  lateral  limbs  are  subequal  and  are  curved  anteriorly,  often  strongly,  and  in  extreme  variants  the  anterior 
half  of  the  outer  limb  may  be  parallel  to  the  anterior  limb.  There  is  no  development  of  a plate  on  the  limbs  and 
they  consist  of  narrow,  strongly  posteriorly  inclined  and  often  curved,  blade-like  processes.  There  may  be  a row 
of  small  nodes  along  the  anterior  margin  or  the  oral  face  may  be  smooth.  The  denticles  originate  from  the 
posterior  margin  and  are  in  the  same  plane  as  the  processes.  They  are  long,  discrete  for  more  than  half  their 
length,  and  laterally  compressed. 

The  aboral  side  is  convex  and  has  a raised  split  keel  which  meets  in  a triangular  pit. 

The  oral  surface  of  the  platform  adjacent  to  the  main  denticle  and  the  carina  is  covered  by  an  irregular  to 
hexagonal  pattern  of  furrows.  These  cross  and  interrupt  a pattern  of  branching  fine  striae.  The  oral  surface  of  the 
main  denticle  also  has  a pattern  of  coarse  striations  but  raised  rather  than  sunken  and  always  semiregular.  These 
run  approximately  parallel  to  the  margin  of  the  denticle  and  meet  at  the  ridge  which  bisects  it.  There  are  no  fine 
striae  on  the  oral  surface.  The  aboral  surface  has  an  irregular  pattern  of  coarse  striations,  again  ridges,  overlying 
a finer  one,  but  in  this  instance  the  finer  one  may  branch  off  the  coarser  ones.  These  patterns  are  also  found  in 
Scaliognathus  anchoralis  and  Scaliognathus  sp.  nov. 

Range.  Lower  part  of  the  anchoralis  Zone. 

No.  of  specimens  sixty. 


Scaliognathus  sp.  nov. 

Plate  34,  figs.  30-32;  text-fig.  10a 

1969  Scaliognathus  anchoralis  Branson  and  Mehl,  Matthews,  pp.  272,  273,  pi.  49,  figs.  5 and  7 only. 

Description.  Anterior  limb  is  slender  and  strongly  arched  with  convex  margins  to  a weakly  developed  plate.  The 
oral  surface  of  the  plate  is  unornamented  and  smooth  except  for  the  median  row  of  discrete  pointed  denticles 
which  are  inclined  posteriorly. 

The  lateral  limbs  are  unequal,  the  outer  being  up  to  twice  the  length  of  the  inner.  Both  limbs  are  slender,  being 
only  slightly  thickened  adjacent  to  the  junction  with  the  anterior  limb.  They  are  steeply  inclined,  almost  vertical, 
with  a slight  inclination  towards  the  posterior.  Their  oral  edge  is  ornamented  with  discrete,  long  denticles 
which  are  laterally  flattened,  with  up  to  five  on  the  outer  and  two  or  three  on  the  inner  side.  The  denticles  are  in 
the  same  plane  as  the  limbs.  The  largest  denticle  is  at  the  end  of  the  carina  and  is  triangular  being  bisected  by 
a low  ridge. 


text-fig.  10.  a,  Scaliognathus  sp.  nov.  Detail  of  platform  area  adjacent  to  main  denticle  of  specimen  1340  (1), 
xl080.  b,  Paragnathodus  multinodosus  (Wirth).  Detail  of  carinal  node  of  specimen  1341X,  x 550.  c, 
Protognathodus  meischneri  Ziegler  1969.  Detail  of  carinal  node  of  specimen  1310  (2),  x 550. 


HIGGINS  AND  WAGNER-GENTIS:  CONODONTS  AND  GONI ATITES 


339 


The  aboral  side  of  the  anterior  limb  is  convex  but  that  of  the  lateral  limb  is  sharp  edged  except  adjacent  to  their 
junction  with  the  anterior  limb.  All  the  limbs  have  split  keels  which  meet  in  a triangular  cavity  beneath  the  base  of 
the  triangular  denticle. 

Range.  Upper  part  of  the  anchoralis  Zone. 


GONIATITES 

(C.  H.  T.  Wagner-Gentis) 

Family  prolecanitidae  Hyatt,  1884 
Genus  merocanites  Schindewolf,  1922 

Type  species.  Ellipsolithes  compressus  Sowerby,  1813. 

Merocanites  marshallensis  (Winchell) 

Plate  35,  figs.  1,  2,  6;  text-fig.  11a,  b 
1862  Goniatites  Marshallensis  Winchell,  pp.  362,  363. 

1955  Merocanites  marshallensis  Winchell,  Miller  and  Garner,  pp.  154-157,  pi.  VII,  figs.  5-9,  text-figs. 
13c  and  16. 

Material.  Two  specimens  and  a large  number  of  whorl-sections  from  localities  138,  138d,  1548,  from  the 
Villabellaco  Limestone  in  Palencia. 

Description.  The  shell  is  a serpenticone  with  a rectangular  cross-section  of  the  whorl.  The  venter  and  ventro- 
lateral edge  are  both  rounded,  the  umbilical  wall  is  flat  and  perpendicular  to  the  lateral  side  (see  text-fig.  1 1 A). 
Neither  ornament  nor  constrictions  have  been  observed.  The  suture  consists  of  an  inflated  ventral  lobe  of  which 
the  siphonal  point  can  be  rather  long.  The  ventro-lateral  saddle  is  rounded,  constricted,  and  low.  The  first  lateral 
lobe  is  pointed,  constricted,  and  not  as  deep  as  the  ventral  lobe,  but  it  is  wider  than  the  ventral  or  second  lateral 
lobe.  The  following  lateral  saddle  is  rounded,  constricted,  and  higher  than  the  ventro-lateral  saddle.  The  second 
lateral  lobe  is  pointed,  constricted,  and  longer  than  the  first  lateral  lobe.  The  last  lateral  saddle  is  rounded, 
constricted,  and  smaller  than  the  preceding  saddles.  The  third  lateral  lobe  is  pointed,  considerably  smaller  than 
the  previous  lateral  lobes,  and  slightly  asymmetrical.  The  suture  crosses  the  umbilical  wall  in  a straight  line, 
sloping  downwards.  It  forms  a lobe  on  the  dorsal  side,  which  is  followed  by  a narrow,  rounded  saddle.  The 
median,  dorsal  lobe  is  narrow,  long,  and  rounded  (see  text-fig.  1 In). 

The  dimensions  of  one  of  the  shells  (loc.  138d)  are:  diameter  38  mm;  width  10  mm;  umbilicus  14  mm;  height  of 
whorl  15  mm;  opening  13  mm. 

Remarks.  Merocanites  marshallensis  europaeus  Kullmann,  1963,  pp.  276-278,  from  the  Esla  area  and 
Puente  de  Alba,  province  of  Leon,  is  quite  different  from  M.  marshallensis  (Winchell),  in  that  its 
whorl  cross-section  is  too  circular,  its  ventral  lobe,  although  inflated,  is  considerably  slimmer,  and  its 
third  lateral  lobe  is  too  developed. 

Comparisons.  The  specimens  described  here  differ  from  M.  applanatus  bicarinatus  Pareyn  in  that  the 
first  lateral  lobe  is  wider;  the  whorls  are  slightly  more  indented  and,  above  all,  no  remnants  are  found 
of  any  ventro-lateral  ridges.  One  specimen  from  Olleros  de  Alba  (Leon),  which  is  identical  to 
bicarinatus,  shows  very  clearly  the  differences  mentioned  above. 

Occurrence.  M.  marshallensis  (Winchell)  is  known  from  the  Marshall  Sandstone  in  Michigan,  U.S.A.,  where  it 
occurs  together  with  Winchelloceras  allei,  Muensteroceras  oweni,  Kazakhstania  karagandaensis,  Gatlendorfia 
stummi,  and  Imitoceras  romingeri  which  indicate  a basal  Osagean  age  (Furnish  and  Manger  1973,  p.  3).  In  the 
Villabellaco  Limestone  (Palencia),  it  occurs  together  with  Merocanites  subhenslowi,  Nautellipsites  hispanicus, 
Ammonellipsites  kayseri,  and  Pseudogirtyoceras  villabellacoi  at  1-5  to  2-5  m above  the  base  of  the  limestone, 
which  indicate  a lowest  Visean  age. 

Family  muensteroceratidae  Librovitch,  1957 
Genus  muensteroceras  Hyatt,  1 883 

Type  species.  Goniatites  oweni  var.  parallela  Hall,  1860. 


340 


PALAEONTOLOGY,  VOLUME  25 


Merocanites  morshollensis  (Winchell) 


a x2  b xl 

text-fig.  1 1.  a,  Suture  of  specimen  BM(NH)  C.82316.  b,  Cross-section  of  same 
specimen. 


Muensteroceras  parallelum  (Hall) 

Plate  35,  figs.  4,  5;  text-fig.  12 
1860  Goniatites  oweni  var.  parallela  Hall,  p.  100,  figs.  13-14. 

1903  Muensteroceras  parallelum  Hall,  J.  P.  Smith,  pp.  121,  122,  pi.  XVI,  fig.  3;  pi.  XIX,  figs.  1,  2. 

1927  Muensteroceras  aff.  parallelum  Hall,  Librovitch,  pp.  32,  33,  pi.  V,  figs.  8,  9;  pi.  VI,  fig.  1. 

1951  Munsteroceras  parallelum  Hall,  Miller  and  Collinson,  p.  471,  fig.  7. 

1961  Munsteroceras  parallelum  Hall,  Pareyn,  pp.  96,  97,  pi.  VII,  figs.  1-3. 

Material.  One  solid  specimen  from  locality  136b  from  the  Villabellaco  Limestone  (Palencia). 

Description.  The  shell  is  a very  flat,  involute  platycone,  with  rounded  venter  and  flat  sides.  At  a diameter  of 
60  mm  the  width  measures  1 5 mm.  The  umbilicus  is  poorly  preserved,  but  is  probably  one-sixth  of  the  diameter. 
The  previous  whorl,  however,  at  a diameter  of  22  mm  has  a width  of  10  mm,  which  is  twice  as  wide  as  the  last 
whorl.  It  also  has  more  rounded  lateral  sides  than  those  of  the  last  whorl.  Neither  constrictions  nor  ornament 
are  preserved.  The  sutures  consist  of  ventral  lobes  with  parallel  sides  which  touch  each  other,  and  thus  form 
two  parallel  lines  on  the  venter.  The  ventro-lateral  saddles  are  fairly  narrow  and  rounded;  the  lateral  lobes  are 
V-shaped  and  their  points  reach  lower  than  the  ventral  lobes;  they  do  not,  however,  touch  the  preceding  ventro- 
lateral saddles.  The  lateral  lobes  are  wide  and  rounded.  There  is  a small,  narrow,  pointed  umbilical  lobe  on  the 
umbilical  edge.  The  dimensions  of  the  shells  are:  diameter  60  mm;  width  15  mm;  umbilicus  10-12  mm;  height 
of  whorl  25  mm. 

Comparisons.  Muensteroceras  parallelum  is  very  similar  to  M.  rotella  de  Koninck  but  the  latter 
appears  to  differ  by  having  the  ventral  lobes  encasing  each  other,  whereas  in  M.  parallelum  they  only 
touch  each  other. 


EXPLANATION  OF  PLATE  35 

Fig.  1.  Merocanites  marshallensis  (Winchell).  BM(NH)  C.82315.  Lateral  view,  showing  the  rather  wide  ventro- 
lateral lobe,  x 3. 

Fig.  2.  Merocanites  marshallensis  (Winchell).  BM(NH)  C.82316.  Lateral  view,  showing  shape  of  shell,  x 3. 

Fig.  3.  Muensteroceras  cf.  crassum  (Foord).  BM(NH)  C.82318.  Ventro-lateral  view,  showing  sutures,  x 3. 
Fig.  4.  Muensteroceras  parallelum  (Hall).  BM(NH)  C.82317.  Showing  lateral  and  part  of  ventral  sutures,  x 1. 
Fig.  5.  Muensteroceras  parallelum  (Hall).  Same  specimen  as  fig.  4.  Lateral  view,  showing  the  different  shapes  of 
the  last  and  penultimate  whorl,  x 1 . 

Fig.  6.  Merocanites  marshallensis  (Winchell).  BM(NH)  C.82315.  Ventral  view,  showing  inflated  ventral  lobe,  x 1. 


PLATE  35 


HIGGINS  and  WAGNER-GENTIS,  Earlier  Carboniferous  goniatites 


342 


PALAEONTOLOGY,  VOLUME  25 


Occurrence.  Villabellaco  Limestone,  locality  1 36b,  at  approximately  the  same  horizon  as  138c,  but  further  east  in 
the  outcrop.  M.  parallelum  was  found  originally  in  the  Rockford  Limestone  of  Indiana,  U.S.A.  It  is  also  known 
from  Hassi  Sguilma  in  Algeria,  in  the  S!  unit  of  Pareyn  1961,  and  from  the  Tien  Shan  in  Central  Asia  (Librovitch 
1927). 


Muensteroceros  parallelum  (Hall) 


text-fig.  12.  Suture  of  specimen  BM(NH) 
C. 82317. 


Muensteroceras  cf.  crassum  Foord 
Plate  35,  fig.  3;  text-fig.  13 

1903  Glyphioceras  ( Muensteroceras ) crassum  Foord,  pp.  193-194,  pi.  XLIII,  fig.  10 a-c. 

1927  Muensteroceras  crassum  Foord;  Librovitch,  pp.  34-35,  pi.  VI,  fig.  6 a-c. 

1941  Muensteroceras  crassum  Foord;  Delepine,  p.  58,  pi.  II,  figs.  4-6. 

1961  Muensteroceras  crassum  Foord;  Pareyn,  pp.  100-101,  pi.  VIII,  figs.  11-16. 

1964  Muensteroceras  cf.  crassum  Foord;  Wagner-Gentis  in  Higgins  et  al. 

Material.  One  solid  specimen  from  the  Villabellaco  Limestone,  found  between  localities  138b  and  c. 

Description.  The  shell  is  an  involute  ellipsocone.  Width  about  half  the  diameter.  The  whorls  have  a rounded 
venter  and  rounded  sides,  with  the  greatest  width  near  the  umbilicus.  The  umbilicus  is  approximately  one- 
quarter  of  the  diameter.  Umbilical  edges  are  rounded  and  the  umbilical  wall  is  almost  perpendicular  and  sloping 
towards  the  centre  of  the  umbilicus  (see  text-fig.  1 3 b). 

No  ornament  has  been  preserved.  There  is  a faint,  shallow  impression  of  a constriction  which  crosses  three- 
quarters  of  the  sides  more  or  less  in  a straight  line,  and  then  forms  a rounded  sinus  across  the  venter. 

The  suture  line  consists  of  a fairly  narrow,  parallel-sided  ventral  lobe  with  a low  median  saddle.  The  ventro- 
lateral saddles  are  rounded  and  narrow;  the  lateral  lobes  are  pointed  and  as  deep  as  the  ventral  lobe.  The  ventral 
side  of  the  lateral  lobe  is  straight,  whereas  the  umbilical  side  is  curved.  The  second  lateral  saddle  is  low,  wide,  and 
rounded.  The  umbilical  lobe  is  small  and  pointed,  and  situated  just  past  the  umbilical  edge,  on  the  umbilical  wall 
(see  text-fig.  \2>a).  The  suture  lines  do  not  encase  each  other. 

Dimensions.  Diameter  23  mm;  width  12  mm;  umbilicus  6 mm;  height  of  whorl  9 mm. 

Comparisons.  The  specimen  compares  with  Muensteroceras  subglobosum  Librovitch,  1927, 
pp.  35-36,  text-fig.  17;  pi.  VI,  fig.  7;  pi.  VII,  figs.  1,  2,  in  having  a similar  suture  line,  but  differs  in 
having  a larger  umbilicus. 

Occurrence.  Villabellaco  Limestone  (Palencia)  between  the  localities  1 38b  and  c.  It  has  also  been  described  from 
Olleros  de  Alba  (Leon).  M.  crassum  was  first  recorded  from  the  Lower  Carboniferous  Limestone  of 
Ballinacarriga,  Co.  Limerick,  Eire.  It  is  also  known  from  Hassi  Sguilma,  in  Algeria,  in  the  unit  of  Pareyn 
(1961),  and  from  the  Tien  Shan  in  Central  Asia  (Librovitch  1927). 

Family  pericyclidae  hyatt,  1900 
Genus  ammonellipsites  Parkinson,  1822 


Type  species.  Ellipsolithes  funatus  Sowerby,  1814. 


HIGGINS  AND  WAGNER-GENTIS:  CONODONTS  AND  GONI ATITES 


343 


Ammonellipsites  kayseri  (Schmidt) 

Plate  36,  figs.  2,  3,  5-7;  text-fig.  14a,  b 

1889  Pericyclus  virgatus  Holzapfel,  p.  34,  Taf.  Ill,  figs.  8,  9. 

1925  Pericyclus  kayseri  Schmidt,  pp.  554,  555,  Taf.  20,  fig.  10. 

Material.  Ten  specimens  from  the  Villabellaco  Limestone  (Palencia)  and  one  specimen  from  Olleros  de  Alba 
(Leon).  All  show  sutures  and  a number  show  the  ornament  fairly  clearly  or  at  least  traces  of  the  ornament.  A 
plaster  cast  of  Holzapfel’s  specimen  (1889,  III,  figs.  8,  9)  has  been  used  in  the  description. 

Description.  The  shell  is  discoidal,  involute  with  a small  umbilicus.  The  venter  and  the  sides  are  rounded.  The 
early  whorls  are  wider  than  high,  but  gradually  become  higher  than  wide.  The  greatest  width  is  half-way  down 
the  lateral  side  of  the  whorl.  The  umbilicus,  about  one-fifth  of  the  diameter,  has  rounded  shoulders  and  rounded 
sides,  which  are  perpendicular  to  the  lateral  sides  (see  text-fig.  14 b). 

The  ornamentation  consists  of  fairly  fine  undivided  ribs,  which  lean  slightly  forward  over  the  sides,  and  form  a 
very  shallow  sinus  over  the  venter.  The  ribs  are  sharp  edged  and  on  the  venter  the  distance  between  the  ribs  is 
more  than  the  width  of  the  rib  itself.  A specimen  with  a diameter  of  about  20  mm  has  six  ribs  per  5 mm,  on  the 
venter.  No  constrictions  are  observed. 

The  suture  consists  of  a medium  saddle,  which  reaches  to  one-third  of  the  ventro-lateral  saddles.  These  are 
slightly  spatulate  with  rounded  points  and  have  a tendency  to  diverge.  The  rounded  points  are  due  to  the  easy 
erosion  of  the  sharp  point.  The  lateral  lobe  is  spatulate  and  pointed.  The  second  lateral  saddle  is  spatulate  with  a 
rather  blunted  point  and  reaches  four-fifths  of  the  height  of  the  first  lateral  saddle.  The  umbilical  lobe  is  situated 
on  the  umbilical  wall  and  is  pointed  and  wide. 


Villabellaco  Limestone Plastercastofholotype 

B.M.  (N.H.)  numbers  C82324  C82321  C82322  Holzapfel’s  specimen 


Diameter 

20  mm 

19  mm 

— 

38-5  mm 

Width 

10  mm 

1 1 mm 

17  mm 

19 

mm 

Umbilicus 

— 

4 mm 

— 

8 

mm 

Height  of  whorl 

8 mm 

8 mm 

20  mm 

17 

mm 

Height  of  opening 

— 

6 mm 

12  mm 

12 

mm 

Comparison.  The  specimens  look  very  similar  to  Pericyclus  virgatus  de  Koninck,  and  Holzapfel 
identified  his  specimens  from  Liebstein  with  this  species.  Unfortunately,  Pericyclus  virgatus  does  not 
show  any  sutures  and  it  is  therefore  impossible  to  decide  whether  kayseri  is  the  same  as  virgatus. 

Occurrence.  In  the  Erdbach  and  Breitscheid  cephalopod  limestone  at  Liebstein,  Germany  (Holzapfel  1889,  pp. 
34  and  35;  Schmidt  1925,  p.  494),  which  according  to  Schmidt  (1925)  belongs  to  the  Ily  zone  of  the  Visean,  which 


344 


PALAEONTOLOGY,  VOLUME  25 


Ammonellipsites  kayseri  (Schmidt) 


t 


specimen  BM(NH)  C.82322. 


is  also  known  as  the  Pey  of  the  Erdbachium,  the  very  base  of  the  Visean.  From  Spain,  Schmidt  (1931,  p.  1035) 
recorded  it  together  with  P.  kochi,  Imiteroceras  sp.,  Merocanites  applanatus,  Muensteroceras  aff.  inconstans,  M. 
cf.  spheroidale,  in  Palencia  at  1 -3  to  2-8  m above  the  base  of  the  Villabellaco  Limestone,  and  in  Leon  at  Olleros  de 
Alba.  In  Great  Britain  it  is  mentioned  by  Prentice  and  Thomas  (1960,  p.  6)  from  Tawstock  and  Codden  Hill 
where  it  occurs  with  Prolecanites  aff.  similis. 

Family  girtyoceratidae  Wedekind,  1918 
Genus  winchelloceras  Ruzhencev,  1965 

Type  species.  Beyrichoceras  allei  Miller  and  Garner,  1955. 

Winchelloceras  palentinus  sp.  nov. 

Plate  36,  fig.  1;  text-fig.  15a-c 

Type  material.  One  solid  specimen  preserved  in  a marly  limestone,  showing  part  of  the  living  chamber,  sutures, 
and  constrictions. 

Repository  of  holotype.  British  Museum  (Nat.  Hist.),  No.  C82317. 

Diagnosis.  Shell  involute,  platyconiform  with  an  umbilicus  less  than  one-sixth  of  the  diameter. 
Constrictions  strongly  marked,  with  a deep,  narrowly  rounded  sinus  on  the  venter,  ventro-lateral 


EXPLANATION  OF  PLATE  36 

Fig.  1.  Winchelloceras  palentinus  sp.  nov.  BM(NH)  C.82319.  Ventro-lateral  view,  showing  sutures  and 
constriction,  x 3. 

Fig.  2.  Ammonellipsites  kayseri  (Schmidt).  BM(NH)  C. 82324.  Ventro-lateral  view,  showing  ornament,  x 3. 

Fig.  3.  Ammonellipsites  kayseri  (Schmidt).  BM(NH)  C.82320.  Ventro-lateral  view,  showing  ornament  and 
suture,  x 3. 

Fig.  4.  Pseudogirtyoceras  villabellacoi  sp.  nov.  BM(NH)  C.82323.  Ventro-lateral  view,  showing  sutures  and 
keeled  venter,  x 3. 

Fig.  5.  Ammonellipsites  kayseri  (Schmidt).  BM(NH)  C.82322.  Lateral  view,  showing  traces  of  ornament  and 
sutures,  x 3. 

Fig.  6.  Ammonellipsites  kayseri  (Schmidt).  BM(NH)  C.82321.  Ventral  view  showing  sutures,  x 2. 

Fig.  7.  Ammonellipsites  kayseri  (Schmidt).  Plastercast  of  holotype.  Lateral  view,  x 1-5. 


PLATE  36 


HIGGINS  and  WAGNER-GENTIS,  Earlier  Carboniferous  goniatites 


346 


PALAEONTOLOGY,  VOLUME  25 


salient,  shallow  sinus,  and  low  salient  on  the  lateral  sides.  Suture  with  a ventral  lobe,  of  which  the 
spikes  are  pointed  outwards  and  the  cheeks  strongly  diverging  apically.  At  a diameter  of  38  mm  the 
median  saddle  is  still  at  the  very  bottom  of  the  ventral  lobe.  Ventro-lateral  saddle  rounded,  lateral 
lobe  pointed,  second  lateral  saddle  wide  and  rounded.  A small  V-shaped  umbilical  lobe  is  positioned 
on  the  umbilical  wall. 

Description.  The  shell  is  a nearly  flat,  involute  platycone.  The  fairly  flat,  only  slightly  rounded,  venter  is 
perpendicular  to  the  almost  flat  lateral  sides.  The  ventro-lateral  edge  is  rounded  and  there  is  a suggestion  of  a 
groove  just  below  the  edge.  From  there,  the  lateral  side  gently  curves  to  become  flat  near  the  umbilical  region, 
where  the  shell  has  its  greatest  width.  The  umbilicus  is  small,  and  stepped  with  rounded  edges  and  narrow  walls, 
which  are  perpendicular  to  the  sides  (see  text-fig.  156). 


text-fig.  15.  a , Suture  of  specimen  BM(NH)  C. 82319.  b.  Cross-section, 
same  specimen,  c.  Constriction,  same  specimen. 


No  ornament  is  preserved.  The  shell  shows  four  deep,  narrow  constrictions  per  whorl.  They  form  a deep, 
narrowly  rounded  sinus  over  the  venter  and  a ventro-lateral  salient,  which  may  create  the  impression  of  a ventro- 
lateral groove.  On  the  lateral  side  the  constrictions  form  a shallow  sinus  and  low  salient.  They  then  fade  out  into 
the  umbilicus  (see  text-fig.  15c). 

The  sutures  have  a ventral  lobe  with  an  extremely  low  median  saddle,  flanked  by  spikes  that  are  pointing 
outwards.  The  cheeks  of  the  ventral  lobe  are  sinuous,  first  bending  outwards  and  then  straightening,  in  a manner 
which  is  opposite  to  the  sinuous  cheeks  of  the  ventral  lobe  in  the  beyrichoceratids.  The  ventro-lateral  saddles  are 
rounded.  The  pointed  lateral  lobes  are  wide  and  slightly  inflated  and  the  second  lateral  saddles  are  wide  and 
rounded,  ending  in  a small  V-shaped  lobe  on  the  umbilical  wall  (see  text-fig.  1 5a).  There  are  twelve  sutures  visible 
on  the  last  whorl,  half  of  the  whorl  forming  the  living  chamber. 

The  dimensions  of  the  shell  are:  diameter  38  mm;  width  10  mm;  umbilicus  6 mm;  height  of  whorl  19  mm;  height 
of  opening  7 mm. 

Comparisons.  Differs  from  Winchelloceras  allei  in  having  a lower  median  saddle,  a flatter  venter,  and 
generally  a slimmer  outline. 

Occurrence.  Found  in  the  basal  50  cm  of  the  Villabellaco  Limestone  (Palencia)  section.  The  genus  is  known  from 
the  U.S.A.,  where  it  occurs  in  the  Coldwater  shale/Marshall  sandstone  of  Michigan,  which  is  of  basal  Osagean 
age  (Furnish  and  Manger  1973).  It  is  also  recorded  from  Tien  Shan  (Popov  1968)  and  the  Urals  (Popov  1975)  in 
rocks  of  CjVj  age. 


pseudogirtyoceras  gen.  nov. 
Type  species.  Pseudogirtyoceras  villabellacoi  sp.  nov. 


HIGGINS  AND  WAGNER-GENTIS:  CONODONTS  AND  GONI ATITES 


347 


Diagnosis.  P seudogirtyoceras  closely  fits  the  description  of  Girtyoceras,  but  is  distinguished  by  having 
a narrow  ventral  lobe. 

Repository  of  type  species.  British  Museum  (Nat.  Hist.),  No.  C82323. 

P seudogirtyoceras  villabellacoi  sp.  nov. 

Plate  36,  fig.  4;  text-fig.  16a,  b 

Description.  The  shell  is  an  involute  oxycone  at  a diameter  of  20  mm.  At  a diameter  of  approximately  1 5 mm  the 
venter  is  still  rounded  but  at  20  mm  it  starts  to  form  a keel.  The  greatest  width  (10  mm)  is  at  the  umbilicus  from 
where  the  sides  curve  gently  to  the  keel.  The  umbilicus  is  two-sevenths  of  the  diameter,  its  walls  being  nearly 
perpendicular  to  the  sides  with  which  it  makes  a rounded  edge  (see  text-fig.  16 b).  Two  constrictions  are  visible, 
and  these  form  a wide,  shallow  sinus  on  the  lateral  sides  and  a sinus  on  the  ventral  side.  No  ornament  is 
preserved. 

The  suture  consists  of  an  extremely  narrow  ventral  lobe,  with  its  cheeks  diverging  apically.  The  two  secondary 
lobes  on  both  sides  of  the  median  saddle  are  narrow  and  pointed.  The  median  saddle  reaches  over  half  the  height 
of  the  ventral  lobe.  The  siphonal  notch  consists  of  a deep  loop.  The  first  lateral  saddles  are  rounded  and  directed 
towards  the  umbilicus.  The  first  lateral  lobes  are  V-shaped  and  pointed,  whereby  the  ventral  cheek  is  convex  and 
the  umbilical  cheek  is  concave.  The  second  lateral  saddle  is  wide  and  rounded  and  at  the  umbilical  edge  there  is  a 
small,  pointed  umbilical  lobe  (see  text-fig.  16a). 


text-fig.  16.  a,  Suture  of  specimen  BM(NH)  C. 82323.  b , Cross-section  of  same 
specimen. 


Comparison.  This  specimen  is  similar  to  Girtyoceras  form  H of  Moore  (1946,  p.  403)  but  differs  in 
having  a much  narrower  ventral  lobe. 

Occurrence.  It  occurs  in  the  Villabellaco  Limestone  (Palencia)  at  locality  138d,  3 m above  the  base,  together  with 
Nautellipsites  hispanicus,  Ammonellipsites  kayseri,  Merocanites  marshallensis,  and  M.  subhenslowi. 

Merocanites  subhenslowi  Wagner-Gentis 

See  Higgins  et  al.  1964,  pp.  238-245,  pi.  Ill,  figs.  1 1-13;  pi.  IV,  fig.  14;  text-figs.  A-c,  for  the  description  of  this 
species. 


Nautellipsites  hispanicus  (Foord  and  Crick) 

See  Wagner-Gentis  1960,  pp.  43-51,  figs.  1-3,  for  the  description  of  this  species. 

Acknowledgements.  We  wish  to  thank  Dr.  R.  H.  Wagner  for  collecting  many  of  the  samples,  and  for  considerable 
discussion  during  the  writing  of  the  paper. 


348 


PALAEONTOLOGY,  VOLUME  25 


REFERENCES 

adrichem  boogaert,  h.  a.  van.  1967.  Devonian  and  Lower  Carboniferous  conodonts  of  the  Cantabrian 
Mountains  (Spain)  and  their  stratigraphic  application.  Leid.  geol.  Meded.  39,  129-192. 

Austin,  R.  l.  1974.  The  biostratigraphic  distribution  of  conodonts  in  Great  Britain  and  the  Republic  of  Ireland. 
Belgian  Geol.  Survey  Intern.  Symposium  on  Belgian  Micropalaeontological  Limits,  Namur  1974,  publ.  No.  3, 
1-17. 

— and  husri,  s.  1974.  Dinantian  conodont  faunas  of  County  Clare,  County  Limerick  and  County  Leitrim.  An 
appendix.  Ibid.  18-69. 

baesemann,  J.  f.  1973.  Missourian  (Upper  Pennsylvanian)  conodonts  of  north-eastern  Kansas.  J.  Paleont.  47, 4, 
689-710. 

barrois,  c.  1882.  Recherches  sur  les  terrains  anciens  des  Asturies  et  de  la  Galice.  Mem.  Soc.  Geol.  du  Nord,  I, 
1-630. 

bischoff,  G.  1957.  Die  Conodonten-Stratigraphie  des  rheno-herzynischen  Unterkarbons  mit  Berucksichtigung 
der  fVocklumeria-Stufe  und  der  Devon/Karbon-Grenze.  Abh.  hess  Landesamt  Bodenforsch.  19,  1-64. 
branson,  e.  b.  and  mehl,  m.  G.  1938.  Conodonts  from  the  Lower  Mississippian  of  Missouri.  Univ.  Mo.  Stud.  13, 
128-148. 

— 1941.  New  and  little  known  Carboniferous  conodont  genera.  J.  Paleont.  15,  97-106. 

brenckle,  p.,  lane,  r.  and  collinson,  c.  1974.  Progress  towards  reconciliation  of  Lower  Mississippian 
Conodont  and  Foraminiferal  Zonations.  Geology,  433-436. 
brouwer,  a.  1964.  Deux  facies  dans  le  Devonien  des  Montagnes  Cantabriques  meridionales.  Breviora  geol. 
astur.  8,  1-4,  3-10. 

budinger,  p.  and  kullmann,  j.  1964.  Zur  Frage  von  sedimentations  brechungen  im  goniatiten-und  conodonten 
fiihrenden  oberdevon  und  Karbon  de  Kantarbischen  Gebirges  (Nordspanien).  Neues  Jb.  Miner.  Geol. 
Paldont.  7,  414-429. 

butler,  m.  1973.  Lower  Carboniferous  conodont  faunas  from  the  Eastern  Mendips,  England.  Palaeontology, 
16,  3,477-517. 

comte,  p.  1959.  Recherches  sur  les  terrains  anciens  de  la  Cordillere  Cantabrique.  Mem.  Inst.  Geol.  Minero 
Espana,  60,  1 -440. 

conil,  r.  and  paproth,  e.  1968.  Mit  Foraminiferen  gegliederte  Profile  aus  dem  nordwest-deutschen  Kohlenkalk 
und  Kulm.  Decheniana,  119,  51-94. 

cooper,  c.  l.  1939.  Conodonts  from  a Bushberg-Hannibal  horizon  in  Oklahoma.  J.  Paleont.  13,  379-442. 
delepine,  g.  141.  Les  Goniatites  du  Carbonifere  du  Maroc  et  des  confins  algero-marocains  du  Sud  (Dinantien- 
Westphalien).  Not.  Mem.  Serv.  Geol.  Maroc.,  No.  56,  pp.  1-110. 
del  Rio,  p.  and  menendez  alvarez,  j.  r.  1978.  Estudio  lito  y biostratigrafico  de  la  Caliza  de  Candamo  (Loredo, 
zona  central  de  Asturias).  Trabajos  de  Geol.  379-387. 

foord,  a.  h.  1903.  Monograph  of  the  Carboniferous  Cephalopoda  of  Ireland.  Palaeontogr.  Soc.  of  London,  V, 
147-234. 

furnish,  w.  m.  and  manger,  w.  L.  1973.  Kinderhook  Ammonoids.  Proc.  Iowa  Acad.  Sci.  80,  1-10. 
ginkel,  a.  c.  van,  1965.  Spanish  Carboniferous  fusulinids  and  their  significance  for  correlation  purposes.  Leid. 
geol.  Meded.  34,  175-225. 

groessens,  e.  1977a.  Hypothese  concernant  revolution  de  conodontes  utiles  a la  biostratigraphie  du  Dinantien. 
Belgian  Geol.  Survey  Intern.  Symposium  on  Belgian  Micropaleontological  Limits,  Namur  1974,  publ.  No.  16, 
1-16. 

— 19776.  Distribution  de  conodontes  dans  le  Dinantien  de  la  Belgique.  Ibid.  publ.  no.  17,  1-193. 

hass,  w.  h.  1953.  Conodont  of  the  Barnett  Formation  of  Texas.  U.S.  Geol.  Surv.  Prof.  Paper,  243-F,  69-94. 
higgins,  a.  c.  1962.  Conodonts  from  the  ‘Griotte’  Limestone  of  North  West  Spain.  Notas  Comun.  Inst.  geol. 
min.  Esp.  65,  5-22. 

— 1971.  Conodont  biostratigraphy  of  the  late  Devonian-early  Carboniferous  rocks  of  the  south  central 
Cantabrian  Cordillera.  Trabajos  de  Geol.  3,  179-192. 

— 1974.  Conodont  zonation  of  the  Lower  Carboniferous  of  Spain  and  Portugal.  Belgian  Geol.  Survey  Intern. 
Symposium  on  Belgian  Micropaleontological  Limits,  Namur  1974,  publ.  no.  4,  1-17. 

1975.  Conodont  zonation  of  the  late  Visean-early  Westphalian  strata  of  the  south  and  central  Pennines  of 
northern  England.  Bull.  geol.  Surv.  Gr.  Br.  53,  1 -90. 

— and  bouckaert,  j.  1968.  Conodont  stratigraphy  and  palaeontology  of  the  Namurian  of  Belgium.  Mem. 
Expl.  Cartes  Geol.  et  Min.  Belgique,  10,  1-64. 

— wagner-gentis,  c.  h.  t.  and  wagner,  r.  h.  1964.  Basal  Carboniferous  strata  in  part  of  Northern  Leon, 


HIGGINS  AND  WAGNER-GENTIS:  CONODONTS  AND  GONIATITES 


349 


NW  Spain:  Stratigraphy,  Conodont  and  Goniatite  Faunas.  Bull.  Soc.  belg.  Geol.  Paleont.  Hydrol.  LXXII, 
205-248. 

holzapfel,  E.  1889.  Die  Cephalopoden-fuhrenden  Kalke  des  unteren  Carbon  von  Erdbach-Breitscheid  bei 
Herborn.  Palaont.  Abh.,  n.f.  Bd.  I,  Heft  1,  pp.  1-73. 

jenkins,  T.  b.  h.  1974.  Lower  Carboniferous  conodont  biostratigraphy  of  New  South  Wales.  Palaeontology , 
17,  909-924. 

kullmann,  J.  1963.  Die  Goniatiten  des  Unterkarbons  im  Kantabrischen  Gebirge  (Nordspanien).  II. 
Palaontologie  der  U.O.  Prolecanitina  Miller  & Furnish.  Die  Altersstellung  der  Faunen.  N.  Jb.  Geol.  Palaont. 
116,  3,  269-324. 

librovitch,  l.  s.  1927.  Lower  Carboniferous  Cephalopods  from  the  Son  Kul  Region  (Tian-Shan  Mountains). 
Com.  Geol.  Materiaux  Gener.  App/.,  Livr.  74,  1-57. 

— 1940.  Carboniferous  ammonoids  of  North  Kazakhstan.  Acad.  Sci.  USSR,  Palaeontol.  Inst.,  Palaeontology 
of  USSR,  IV,  Fasc.  1,  1-343. 

lys,  M.  and  serre,  b.  1958.  Contributions  a la  connaissance  des  microfaune  du  Paleozoique.  Etudes 
micropaleontologiques  dans  le  Carbonifere  marin  des  Asturies  (Espagne).  Revue  Inst.fr.  Petrole,  13,  879-916. 
manger,  w.  l.  1979.  Lower  Carboniferous  ammonoid  assemblages  from  North  America.  Huiteme  Congr. 
Avanc  Etud.  Stratigr.  carb.,  c.r.  3 (1975),  211-221. 

marcos,  a.  1967.  Estudio  geologica  deo  reborde  NW  de  los  Picos  de  Europa  (region  de  Onis-Cabrales, 
Cordillera  Cantabrica).  Trabajos  de  Geol.  1,  39-46. 

marks,  p.  and  wensink,  h.  1970.  Conodonts  and  the  age  of  the  ‘Griotte’  Limestone  Formation  in  the  Upper 
Aragon  Valley  (Huesca,  Spain).  Proc.  K.  ned.  Akad.  Wet.,  Ser.  B,  73,  238-275. 

Matthews,  s.  c.  1969a.  A Lower  Carboniferous  conodont  fauna  from  East  Cornwall.  Palaeontology,  12, 
262-275. 

— 1969 b.  Two  conodont  faunas  from  the  Lower  Carboniferous  of  Chudleigh,  South  Devon.  Ibid. 
276-280. 

— sadler,  p.  m.  and  selwood,  e.  b.  1972.  A Lower  Carboniferous  conodont  fauna  from  Chillaton,  south-west 
Devonshire.  Ibid.  15,  550-568. 

— and  thomas,  J.  m.  1974.  Lower  Carboniferous  conodont  faunas  from  north-east  Devonshire.  Ibid.  17, 
371-385. 

mehl,  M.  G.  and  thomas,  L.  A.  1947.  Conodonts  from  the  Fern  Glen  of  Missouri.  J.  Sclent.  Labs.  Denison  Univ. 
40,  3-19. 

meischner,  d.  1970.  Conodonten-Chronologie  des  Deutschen  Karbons.  Sixieme  Congr.  Avanc.  Etud.  Stratigr. 
Carb.  Sheffield  1967,  III,  1169-1180. 

Merrill,  G.  K.  and  Merrill,  s.  m.  1974.  Pennsylvanian  non-platform  conodonts.  Ha:  The  dimorphic  apparatus 
of  Idioprioniodus.  Geol.  Palaeontol.  8,  119-130. 

miller,  a.  K.  and  collinson,  c.  1951.  Lower  Mississippian  ammonoids  of  Missouri.  J.  Paleont.  25,  No.  4, 
454-487. 

— and  garner,  H.  F.  1955.  Lower  Mississippian  Cephalopods  of  Michigan.  Part  III.  Ammonoids  and 
Summary.  Contr.  Mus.  Palaeont.  Univ.  Mich.  XII,  8,  113-173. 

moore,  e.  w.  J.  1946.  The  Carboniferous  goniatite  genera  Girtyoceras  and  Eumorphoceras.  Proc.  Yorks,  geol. 
Soc.  25,  387-445. 

norby,  R.  d.  1975.  Micromorphology  of  natural  conodont  apparatuses  of  Chesterian  age  (abs).  Geol.  Soc.  Amer. 
Abs.  with  programmes,  7,  829. 

pareyn,  c.  1961.  Les  Massifs  Carboniferes  du  Sahara  Sud-Oranais.  Publ.  Centre  Recherches  Sahariennes  (ser. 
Geol.),  1,  tome  II,  Paleontologie,  1-244. 

perrit,  m.  F.  1974.  Biostratigraphie  par  Conodontes  du  Carbonifere  inferieur  des  Pyrenees  Bearnaises.  C.r. 
Acad.  Sci.  Paris,  279,  110,  791-794. 

POPOV,  a.  v.  1968.  Visean  ammonoids  of  the  northern  Tien  Shan  and  their  stratigraphic  significance.  Akad 
Nauk.  Kirgisian  SSR,  Geol.  Inst.  1-116. 

— 1975.  Ammonoidea.  Palaeontological  Atlas.  Carboniferous  Strata  of  the  Urals.  M-vo  geologii  SSSR. 
Vsecoyuz  nef.  naytz-issled.  geol-ravz.  in-t.  Trudi.  383,  111-130,  Tab.  34-51. 
prentice,  j.  E.  and  thomas,  J.  m.  1960.  The  Carboniferous  goniatites  of  North  Devon.  Abstr.  Proc.  Third. 
Conf.  Geol.  Geomorphol.  SW  England,  6-8. 

rexroad,  c.  B.  1958.  Conodonts  from  the  Glen  Dean  Formation  (Chester)  of  the  Illinois  Basin.  Illinois  Geol. 
Surv.  Rept.  Invest.  209,  1-27. 

and  scott,  a.  j.  1964.  Conodont  zones  of  the  Rockford  Limestones  and  New  Providence  Shale 

(Mississippian)  in  Indiana.  Bull.  Indiana  Dep.  Conserv.  Geol.  Surv.  30,  1-54. 


350  PALAEONTOLOGY,  VOLUME  25 

roundy,  p.  v.  1926.  The  microfauna  in  Mississippian  formations  on  San  Saba  County,  Texas.  U.S.  Geol.  Surv. 
Prof.  Paper , 146,  1-63. 

Sandberg,  c.  A.,  ziegler,  w.,  leuteritz,  K.  and  brill,  s.  M.  1978.  Phylogeny,  speciation  and  zonation  of 
Siphonodella  (conodonta.  Upper  Devonian  and  Lower  Carboniferous).  Newsl.  Stratigr.  7 (2),  102-120. 
schmidt,  H.  1925.  Die  carbonischen  Goniatiten  Deutschlands.  Jb.preuss.  geol.  Landesanst.  45  (1924),  489-609. 

— 1931 . Das  Palaozoikum  der  spanischen  Pyrenaen.  Abh.  Ges.  Wiss.  Gottingen,  Math.-physik.,  Kl.  (3),  5,  85. 

— and  muller,  k.  j.  1964.  Weitere  Funde  von  Conodonten-Gruppen  aus  dem  oberen  Karbon  des 
Sauerlandes.  Palaont.  Z.  38,  105-135. 

scott,  h.  w.  1942.  Conodont  assemblages  from  the  Heath  Formation,  Montana.  J.  Paleont.  16,  293-300. 
sitter,  l.  u.  de.  1962.  The  structure  of  the  southern  slope  of  the  Cantabrian  Mountains:  Explanation  of  a 
geological  map  with  sections,  scale  1 : 100,000.  Leid.  geol.  Meded.  26,  255-264. 

Thompson,  T.  L.  1967.  Conodont  zonation  of  lower  Osagean  rocks  (Lower  Mississippian)  of  southwestern 
Missouri.  Missouri  geol.  Surv.  Water  Resour.  Kept.  Inv.  39,  1-84. 

— and  fellows,  L.  d.  1970.  Stratigraphy  and  conodont  biostratigraphy  of  Kinderhookian  and  Osagean  rocks 
of  southwestern  Missouri  and  adjacent  areas.  Ibid.  Rept.  Inv.  45,  1-263. 

veen,  j.  van.  1965.  The  tectonic  and  stratigraphic  history  of  the  Cardano  area,  Cantabrian  Mountains,  north- 
west Spain,  Leid.  geol.  Meded.  35,  45-103. 

voges,  a.  1959.  Conodonten  aus  dem  Untercarbon  I und  II  (Gattendorfia  und  Pericyclus-Stufe)  des  Sauerlandes. 
Palaont.  Z.  33,  4,  266-314. 

von  bitter,  p.  h.  1972.  Environmental  control  of  conodont  distribution  in  the  Shawnee  Group  (Upper 
Pennsylvanian)  of  eastern  Kansas.  Paleont.  Contr.  Univ.  Kansas,  Art.  59,  1-105. 
wagner,  R.  h.  1963.  A general  account  of  the  Palaeozoic  rocks  between  the  Rivers  Porma  and  Bernesga  (Leon, 
NW  Spain).  Bol.  Inst.  geol.  min.  Espaiia,  LXXIV,  171-331. 

— and  wagner-gentis,  c.  h.  t.  1963.  Summary  of  the  stratigraphy  of  Upper  Palaeozoic  Rocks  in  NE 
Palencia,  Spain.  Proc.  Kon.  ned.  Akad.  Wet.,  Series  B,  66,  149-163. 

— winkler  prins,  c.  F.  and  riding,  r.  e.  1971 . Lithostratigraphic  units  of  the  lower  part  of  the  Carboniferous 
in  northern  Leon,  Spain.  Trabajos  de  Geol.  4,  603-633. 

wagner-gentis,  c.  h.  t.  1960.  On  Nautellipsites  hispanicus  (Foord  and  Crick)  Estudios  geol.  Inst.  Invest,  geol. 
Lucas  Mallada,  XVI,  43-51. 

wirth,  M.  1967.  Gliederung  des  hoheren  Palaozoikums  (Givet-Namur)  im  Gebiet  des  Quinto  Real  (West- 
pyrenaen)  mit  hilfe  von  Conodonten.  Abh.  Neues  Jahrb.  Geol.,  Pal.  127,  179-240. 
ziegler,  w.  1962.  Taxonomie  und  Phylogenie  Oberdevonischer  Conodonten  und  ihre  stratigraphische 
Bedeutung.  Abh.  Hess.  Landesamt.  Bodenf.  38,  1-166. 

— 1969.  Eine  neue  Conodontenfauna  aus  dem  hochsten  Oberdevon.  Fortschr.  Geol.  Rheinld.  u.  Westf.  17, 
343-360. 

— and  leuteritz,  k.  1970.  Palaontologischer  Anhang:  Conodonten.  In  koch,  m.,  Leuteritz,  k.  and  ziegler, 
w.,  Fortschr.  Geol.  Rheinld.  u.  Westf.  17,  712-715. 


A.  C.  HIGGINS 
Department  of  Geology 
University  of  Sheffield 
Sheffield  SI  3JD 

C.  H.  T.  WAGNER-GENTIS 

Mayfield 
Cross  Lane 

Typescript  received  19  May  1980  Calver 

Revised  typescript  received  23  October  1980  Sheffield  S30  1XS 


LIASSIC  PLESIOSAUR  EMBRYOS 
REINTERPRETED  AS  SHRIMP  BURROWS 

by  RICHARD  A.  THULBORN 


Abstract.  A peculiar  nodule  from  the  Upper  Liassic  (Toarcian)  shales  of  Whitby,  Yorkshire,  is  interpreted 
as  an  infilled  burrow  system  which  was  probably  excavated  by  a shrimp-like  crustacean  (possibly  Glyphea 
sp.).  This  interpretation  is  supported  by  comparisons  with  fossil  crustacean  burrows  in  the  ichnogenus 
Thalassinoides.  The  nodule  had  formerly  been  regarded  as  a cluster  of  fossil  embryos  from  the  aquatic  reptile 
Plesiosaurus. 


In  September  1887  H.  G.  Seeley  delivered  four  reports  on  fossil  reptiles  to  the  British  Association 
meeting  at  Manchester.  Three  of  those  reports  dealt  with  anatomy  and  systematics,  but  the  fourth 
concerned  the  more  unusual  subject  of  fossil  embryos.  Seeley  described  and  exhibited  a peculiar 
nodule  showing  supposed  embryos  of  a Jurassic  plesiosaur— an  aquatic  reptile  of  the  suborder 
Plesiosauria  (order  Sauropterygia).  His  account  of  the  embryos  was  summarized  the  following  year 
(Seeley  1888),  though  a full  description  did  not  appear  until  1896. 

Seeley’s  plesiosaur  embryos  have  attracted  little  attention:  they  have  never  been  figured,  and  it  is 
difficult  to  find  more  than  passing  mention  of  them  in  the  literature  of  vertebrate  palaeontology.  They 
were  briefly  noticed  by  Woodward  (1898,  p.  161),  Williston  (1914,  p.  94),  and  de  Saint-Seine  (1955, 
p.  422),  but  seem  never  to  have  been  re-examined  in  detail.  Abel’s  classic  work  Vorzeitliche 
Lebensspuren  (1935)  devoted  much  attention  to  embryos  and  neonates  of  ichthyosaurs  (reptile  order 
Ichthyosauria),  but  made  no  mention  at  all  of  plesiosaur  embryos.  There  has  often  been  speculation 
about  the  mode  of  reproduction  in  plesiosaurs  (see,  for  example,  Robinson  1975),  but  the  embryos 
described  by  Seeley  seem  to  have  been  overlooked. 

This  paper  provides  the  first  illustrations  of  the  supposed  plesiosaur  embryos,  and  offers  a 
reinterpretation  of  these  curious  fossils. 


MATERIAL 

The  material  described  by  Seeley  was  acquired  in  1909  by  the  British  Museum  (Natural  History),  where  it  is 
now  catalogued  as  R 3585.  It  comprises:  an  irregular  nodule  of  pyritic  mudstone  and  shale,  approxi- 
mately 11  x 8x8  cm  (text-fig.  1);  a small  fragment,  broken  from  the  main  nodule;  five  slides,  each  with  a 
thin  section;  a slide  with  twenty-three  stained  serial  sections  from  the  neck  region  of  a modern  lizard 
embryo.  The  small  fragment  was  detached  in  Seeley’s  search  for  internal  structure,  and  it  provided  material  for 
the  thin  sections;  the  serial  sections  of  the  lizard  embryo  had  been  obtained  for  comparative  purposes  (Seeley 
1896,  p.  21). 

The  nodule  almost  certainly  came  from  coastal  exposures  of  the  Upper  Lias  (Whitbian  sub-stage  of  the 
Toarcian)  near  Whitby,  Yorkshire.  Seeley  received  the  nodule  in  1887  from  J.  F.  Walker  (then  curator  of  the 
Yorkshire  Natural  History  Society),  who  had  obtained  it  from  a dealer  in  Whitby.  The  Whitby  fossil-dealers 
were  not  averse  to  importing  their  wares  from  the  Lower  Liassic  of  Lyme  Regis,  Dorset,  but  probably  did  so  only 
in  the  case  of  exceptionally  fine  and  valuable  specimens  (notably  fossil  fishes;  see  Blake  1876,  pp.  257-259).  The 
Liassic  shales  near  Whitby  are  rife  with  oddly  shaped  concretions  and  nodules  (Hallam  1962;  Howarth  1962), 
and  it  seems  unlikely  that  a dealer  would  have  imported  an  example  from  Dorset  when  so  many  were  to  hand 
locally. 


[Palaeontology,  Vol.  25,  Part  2,  1982,  pp.  351  359. | 


352 


PALAEONTOLOGY,  VOLUME  25 


The  Whitby  Lias  is  about  76  m thick,  and  may  be  summarized  as  follows  (incorporating  data  from  Hallam 
1962;  Howarth  1962,  1973;  Hemingway  1974): 


Formation  Thickness  (m)  Ammonite  Zone 


Stage 


Dogger  (sandstones  and  ironstones) 

- unconformity 


Cement  Shales 

5-5] 

Main  Alum  Shales 

22-0 

Hard  Shales 

4-5  J 

Bituminous  Shales 

23-5  1 

Jet  Rock  Shales 

8-5  ) 

Grey  Shales 

13-5} 

Hildoceras  bifrons 

Harpoceras  falciferum 
Dactylioceras  tenuicostatum 


Bajocian 


Toarcian 


The  succession  consists  of  shales,  with  minor  argillaceous  limestones  and  rare  seams  of  siderite  mudstone. 
Nodules  and  concretions  are  extremely  abundant;  some  are  scattered  at  random  through  the  shales,  while  others 
occur  in  constant  bands  which  serve  as  useful  marker  horizons  (see  Howarth  1962).  Concretions  in  these  marker 
horizons  often  have  highly  distinctive  shapes,  and  have  been  named  accordingly— ‘cannon  balls’,  ‘cheese 
doggers’,  ‘curling  stones’,  ‘pseudovertebrae’,  and  so  on.  Unfortunately  the  nodule  described  by  Seeley  cannot  be 
referred  with  certainty  to  any  of  these  marker  horizons.  Seeley’s  specimen  has  a matrix  of  soft,  grey,  flaky,  and 
non-bituminous  shale,  and  is  unlikely,  for  that  reason,  to  have  been  obtained  from  the  zone  of  Harpoceras 
falciferum— where,  the  shales  are  usually  brown  in  colour  and  often  have  a characteristic  ‘oily’  smell.  Nor  is  it 
very  likely  that  Seeley’s  specimen  came  from  the  Hard  Shales  or  the  Main  Alum  Shales:  in  many  places  these 
shales  weather  to  a brown  or  reddish  colour,  and  they  sometimes  produce  efflorescent  alum.  The  specimen 
probably  originated  from  the  Grey  Shales  or  the  Cement  Shales. 

DESCRIPTION 

The  supposed  plesiosaur  embryos  are  rounded  masses  of  grey-brown  mudstone  protruding  from  a core  of  flaky 
grey  shale  (text-fig.  1).  The  shale  is  very  soft,  and  contains  tiny  blebs  and  veins  of  white  calcite,  together  with 
occasional  flecks  of  black  plant  material.  The  bodies  of  the  embryos  are  slightly  harder  than  the  shale  matrix, 
extremely  fine-grained,  and  rich  in  finely  divided  pyrite.  Seeley  originally  believed  the  embryos  to  be  phosphatic 
(1888),  but  later  determined  that  they  were  in  fact  pyritic  (1896).  The  embryos,  and  some  areas  of  intervening 
shale,  have  a greasy  lustre  which  has  probably  been  enhanced,  if  not  produced,  by  repeated  handling  of  the 
specimen. 

Seeley  identified  and  numbered  four  principal  embryos,  with  ‘indications  of  three  or  four  others’  (1896,  p.  20). 
His  numbering  is  still  visible  on  the  specimen,  in  faded  red  ink,  and  will  be  followed  here  (see  text-fig.  2).  For  the 
sake  of  brevity  I will  also  adhere  to  Seeley’s  descriptive  terminology  (e.g.  the  ‘head’  of  embryo  1,  the  ‘limbs’  of 
embryo  3,  and  so  on);  and  to  assist  my  discussion  the  specimen  will  be  oriented  as  a spheroid  with  the  ‘head’  of 
embryo  1 directed  to  the  ‘North  Pole’. 

Seeley’s  description  (1896)  is  both  detailed  and  accurate  and  need  not  be  repeated  here.  However,  the  specimen 
has  certain  features  that  were  not  mentioned  by  Seeley.  First,  it  is  badly  damaged  in  the  region  of  the  ‘North 
Pole’,  where  it  seems  that  several  large  flakes  have  been  removed  by  hammer  blows  (text-figs.  1a,  2a).  The 
fragment  detached  to  provide  thin  sections  was  clearly  the  last  piece  to  be  removed,  for  it  can  still  be  fitted  on  to 
the  main  nodule,  where  it  forms  part  of  an  older  fracture-scar.  In  some  places  the  specimen  appears  to  have  been 
subjected  to  rather  rough  mechanical  preparation;  there  are,  for  example,  very  deep  needle-marks  along  the  right 
side  of  the  ‘neck’  in  embryo  1 . 

The  embryos  are  clustered  in  a partly  overlapping  arrangement;  the  ‘neck’  of  embryo  3,  for  example,  is  largely 
concealed  by  the  ‘body’  of  embryo  1 . Elsewhere  one  embryo  may  be  joined  to  another  without  the  slightest  trace 
of  a dividing  line  (see  text-fig.  1b).  The  arrangement  of  the  supposed  embryos  is  not  entirely  random:  they  extend 
in  a radiating  pattern,  along  meridians,  when  viewed  from  the  ‘South  Pole’  (text-fig.  2b). 

Seeley  could  find  no  definite  trace  of  organic  structure  in  the  specimen,  and  his  identification  of  ‘embryonic 
plesiosaurs’  rested  almost  entirely  on  the  criterion  of  shape.  He  considered  that  embryo  1 showed  ‘the  head, 
neck,  body,  tail  and  limbs  of  such  a shape  as  a plesiosaur  would  show’,  and  pointed  out  that  the  other  embryos 
were  basically  similar  in  form;  differences  in  shape  between  one  embryo  and  another  were  taken  to  reflect 
‘various  stages  of  development’  (Seeley  1896,  p.  23).  Nevertheless,  Seeley  did  attempt  to  identify  anatomical 


THULBORN:  ‘PLESIOSAUR  EMBRYOS’ 


353 


features  such  as  bones  and  teeth  within  the  embryos — though  he  did  so  with  extreme  circumspection.  These 
possible  organic  structures  merit  close  inspection. 

Placenta.  A ‘smooth  film  of  hard  clay’  over  parts  of  the  specimen  was  considered  by  Seeley  to  have  ‘much  the 
aspect  of  a defining  membrane  of  a placenta-like  character’  (1896,  p.  20).  This  apparently  refers  to  the  greasy 
lustre  on  the  embryos  and  on  some  intervening  areas  of  shale.  The  lustre  is  almost  certainly  artificial,  for  it  is 
easily  reproduced  by  gently  rubbing  a finger-tip  over  a fresh  patch  of  the  shale. 

Median  dorsal  ridge.  ‘Each  of  the  four  principal  specimens  is  characterized  by  a median  longitudinal  ridge  or 
blunt  angle  which  extends  down  . . . what  I regard  as  the  dorsal  surface;  corresponding  in  position  to  the  neural 
spines  of  the  vertebrae’  (Seeley  1896,  p.  21).  The  ridge  is  visible  in  unweathered  and  undamaged  portions  of  the 
embryos,  though  it  is  sometimes  very  faint.  Nowhere  is  there  any  trace  of  organic  structure  in,  or  beneath,  the 
ridge;  and  there  is  no  evidence  to  confirm  that  the  ridge  is  ‘dorsal’  in  position. 

Sclerotic  ring.  On  the  right  side  of  the  ‘head’  in  embryo  1 is  a shallow  oval  pit  which  was  tentatively  identified  by 
Seeley  as  the  orbit.  This  pit  originally  contained  ‘a  scale  with  a radiated  structure,  which  had  the  appearance  of 
being  a sclerotic  circle  of  bones  about  the  eye’,  but  this  was  subsequently  lost  from  the  specimen  (Seeley  1896, 
pp.  22, 24).  The  supposed  orbit  is  now  floored  by  a paper-thin  layer  of  finely  crystalline  pyrite,  which  is  overlain 
near  the  anterior  margin  by  a patch  of  shale  containing  flecks  of  white  calcite.  The  floor  of  the  orbit  appears  to  be 
featureless — though  Seeley  maintained  that  there  was  possibly  ‘some  evidence  of  a radiating  structure’  (1896, 
p.  25).  The  ‘radiated  scale’  described  by  Seeley  is  unlikely  to  have  been  a sclerotic  ring,  because  it  was  originally 
applied  to  the  concave  floor  of  the  ‘orbit’:  sclerotic  plates  are  usually  embedded  in  the  lateral  half  of  the  eyeball, 
and  therefore  tend  to  be  arched  outwards  from  the  orbit  (Walls  1942).  If  the  ‘radiated  scale’  were  a series  of 
sclerotic  plates  it  would  be  necessary  to  suppose  that  the  entire  eyeball  had  been  squashed  flat  in  the  embryo;  and  if 
the  eyeball  had  been  squashed  or  ruptured  it  would  be  reasonable  to  expect  evidence  of  similar  distortion  in  other 
regions  of  the  soft-bodied  embryos.  In  any  case,  it  seems  improbable  that  delicate  and  superficial  ossifications 
would  have  been  preserved,  while  deeper  and  presumably  more  robust  bones  were  obliterated.  I suspect  that  the 
‘radiated  scale’  was  no  more  than  a thin  vein  of  calcite,  resembling  similar  veins  in  other  parts  of  the  nodule. 

Scapula.  The  ‘left  forelimb’  of  embryo  2 is  truncated  by  a fracture,  behind  which  ‘the  dorsal  aspect  of  the  limb 
appears  to  include  a surface  bone  in  the  position  of  the  ascending  process  of  the  plesiosaurian  scapula’  (Seeley 
1896,  p.  23).  I can  find  no  trace  of  bone  in  this  area;  there  is  a slight  surface  irregularity  postero-dorsal  to  the 
‘limb’,  but  this  is  not  particularly  reminiscent  of  the  outline  of  a scapula.  Moreover,  the  ascending  ramus  of  the 
plesiosaurian  scapula  is  situated  in  front  of  the  glenoid  cavity,  and  not  behind  it. 

Muscle  segments.  According  to  Seeley,  ‘there  appear  to  be  some  faint  indications  of  transverse  segmentation  like 
that  of  muscles,  in  the  region  of  the  neck  in  the  specimen  No.  1 , and  in  the  dorsal  region  in  specimen  No.  2’  ( 1 896, 
p.  23).  There  seems  to  be  no  trace  whatsoever  of  segmental  structure  in  embryo  2.  Much  of  the  ‘neck’  of  embryo  1 
has  an  irregular,  flaky,  and  pitted  surface.  In  its  posterior  third  the  right  side  of  the  neck  carries  a series  of  deep 
needle-marks  which  extend  into  the  adjoining  shale;  it  is  improbable  that  these  would  be  taken  as  evidence  of 
segmentation,  even  at  a casual  glance.  Above  this  line  of  needle-marks  is  a smooth  area  which  proves,  on  close 
inspection  ( x 50),  to  carry  an  ornament  of  extremely  fine,  straight,  and  parallel  scratches.  These  microscopic 
scratches  were  most  probably  produced  with  a mild  abrasive:  they  are  so  regular  in  depth,  spacing,  and  orienta- 
tion (antero-dorsal  to  postero-ventral)  that  they  cannot  have  been  produced  individually,  and  they  are  so 
unvaryingly  straight  and  parallel  that  they  do  not  seem  to  be  a natural  feature.  Over  all,  there  is  no  clear  evidence 
of  segmentation  in  any  of  the  embryos. 

Mandible  and  teeth.  ‘It  is  possible  . . . that  the  lower  jaw  may  be  indicated  in  the  film  of  clay,  which  is  imperfectly 
preserved  beneath  the  head  [of  embryo  1],  and  that  some  small  badly  preserved  white  spots  arranged  in  linear 
succession  are  indications  of  teeth’  (Seeley  1896,  p.  24).  The  ‘head’  of  embryo  1 meets  the  shale  matrix  without 
any  indication  of  a separate  mandible.  Alongside  the  right  surface  of  the  ‘head’  is  a string  of  tiny  white  spots, 
each  about  0T5  mm  in  diameter;  these  are  blebs  of  calcite  similar  to  those  scattered  throughout  the  specimen, 
and  their  roughly  linear  arrangement  seems  to  be  fortuitous. 

Skull  roofing  bones,  external  nostrils,  and  parietal  foramen.  All  these  features  were  tentatively  identified  by  Seeley 
on  the  ‘head’  of  embryo  1 . The  surface  of  the  ‘head’  is  marked  with  a random  assortment  of  pits,  furrows,  and 
scratches,  all  of  which  seem  to  be  preservational  and  weathering  effects  comparable  to  those  in  other  parts  of  the 
embryos.  None  of  the  grooves  or  scratches  can  justifiably  be  regarded  as  a mid-line  suture  between  skull  roofing 
bones,  and  none  of  the  pits  can  be  matched  up  into  a pair  that  could  convincingly  represent  the  external  nostrils. 


354 


PALAEONTOLOGY,  VOLUME  25 


text-fig.  1.  Four  views  of  nodule  from  the  Upper  Liassic  of  Whitby,  Yorkshire,  showing  supposed  plesiosaur 
embryos.  BM(NH)  R.3585;  scale  indicates  2 cm. 


Other  anatomical  features  of  the  embryos  (heads,  necks,  bodies,  tails,  and  limbs)  were  identified  by  Seeley 
purely  on  the  basis  of  their  shape  and  their  position.  In  summary,  there  seems  to  be  no  evidence  of  organic 
structure  in  any  of  the  supposed  embryos. 

DISCUSSION 

There  are  several  reasons  why  the  specimen  is  unlikely  to  be  a cluster  of  plesiosaur  embryos.  First,  it 
lacks  any  definite  organic  structure.  Some  of  the  anatomical  features  identified  by  Seeley  (1896)  are 
no  more  than  surface  irregularities,  scratches,  pits,  and  similar  weathering  effects  (e.g.  ‘scapula’, 
‘nostrils’,  and  ‘skull  bones’);  other  such  features  are  patches  of  calcite  (e.g.  ‘teeth’)  or  artifacts  (e.g. 
‘placenta’).  Second,  the  supposed  embryos  show  differences  in  shape  and  proportions.  Seeley 
accounted  for  this  variation  by  suggesting  that  the  embryos  were  in  different  stages  of  development, 
but  this  is  not  a particularly  convincing  explanation.  Where  embryos,  neonates,  or  juveniles  of 
reptiles  are  preserved  as  natural  groups  of  fossils  the  siblings  are  invariably  in  the  same  state  of 
development— as,  for  example,  in  ichthyosaurs  (Hauff  1921;  Hoffmann  1958;  Urlichs,  Wild,  and 
Ziegler  1979),  in  nothosaurs  (Halstead  1969),  and  in  ornithopod  dinosaurs  (Horner  and  Makela 
1979).  Next,  and  most  important,  it  is  highly  improbable  that  soft-bodied  embryos  could  have  been 
preserved  ‘in  the  round’  in  a shale  matrix.  Sediment  compaction  usually  ensures  that  soft  organisms 
are  preserved  in  a partly  or  completely  flattened  state,  yet  the  supposed  plesiosaur  embryos  form  a 


THULBORN:  PLESIOSAUR  EMBRYOS’ 


355 


text-fig.  2.  Diagrammatic  interpretation  of  text-fig.  1.  Shale  matrix  unshaded,  and  supposed  plesiosaur 
embryos  (here  regarded  as  sediment-filled  crustacean  burrows)  indicated  by  solid  shading.  Fracture  surfaces 
indicated  by  oblique  shading.  Supposed  embryos  (=  burrows)  are  numbered  following  Seeley  (1896),  and  each 
has  its  ‘dorsal  ridge’  (=  groove  in  burrow  floor)  shown  schematically.  Abbreviations:  c— cut  surface;  f—  area 
from  which  a fragment  was  detached  to  provide  thin  sections  for  Seeley  (1896);  h ‘head’  of  supposed  embryo 
(=part  of  shaft  leading  to  burrow  chamber);  m— needle-marks  (shown  schematically)  along  the  side  of 
supposed  embryo  1 (=  burrow  1);  p— spots  indicating  location  of  possible  faecal  pellets;  t— ‘tail’  of  supposed 
embryo  (=  burrow  termination).  Scale  indicates  2 cm. 

spheroidal  mass  which  is  not  obviously  flattened  or  distorted.  If  this  specimen  were  a group  of 
embryos  it  would  be  necessary  to  suppose  that  it  had  been  enclosed  in  a syngenetic  or  diagenetic 
concretion  (following  terminology  of  Pantin  1958),  or  that  the  embryos  had  been  mineralized  before 
compaction  of  the  surrounding  sediment.  The  specimen  was  clearly  not  enclosed  in  a concretion,  for 
the  supposed  embryos  are  separated  and  overlain  by  soft  shale;  and  it  seems  unlikely  that  embryos 
could  have  been  mineralized  in  utero  or  in  an  unconsolidated  shale  matrix.  In  rare  circumstances 
human  embryos  are  known  to  undergo  spontaneous  abortion  and  calcification— to  become  ‘stone 
babies’  or  ‘lithopedia’  (Halstead  and  Middleton  1972)— but  there  are  no  reports  of  comparable 
abnormalities  among  fossil  vertebrates.  It  seems  a very  remote  possibility  that  the  supposed 
plesiosaur  embryos  could  be  reptilian  equivalents  of ‘lithopedia’. 

What,  then,  is  the  most  probable  identity  of  the  specimen?  It  is  quite  unlike  a concretion,  or  even  an 
aggregate  of  small  concretions.  Some  parts  of  the  supposed  embryos  bear  a resemblance  to 
coprolites,  but  the  specimen  as  a whole  seems  too  complicated  in  structure  to  be  a mass  of  coprolites 


356 


PALAEONTOLOGY,  VOLUME  25 


or  of  intestinal  fillings.  Each  of  the  supposed  embryos  is  a multi-lobed  object  (with  several  lobes 
representing  the  ‘tail’  and  the  ‘limbs’),  whereas  individual  coprolites  are  normally  sausage-shaped, 
spiral,  fusiform,  or  pellet-like  objects  which  are  not  developed  into  lobes  (see  Amstutz  1958,  and 
many  references  therein).  Nor  does  the  specimen  resemble  any  known  endocranial  cast  of  a verte- 
brate (Edinger  1929;  Jerison  1973). 

From  a review  of  literature  dealing  with  trace  fossils  and  problematica  I strongly  suspect  that  the 
supposed  plesiosaur  embryos  are  actually  the  sediment-filled  burrows  of  a crustacean— most 
probably  a shrimp  resembling  those  of  the  decapod  superfamily  Thalassinoidea.  The  thalassinoid 
shrimp  Callianassa  excavates  distinctive  burrow  systems  in  modern  marine  sediments  (see 
Braithwaite  and  Talbot  1972),  and  comparable  burrows  have  been  reported  as  far  back  as  the 
Triassic  (Fiege  1944;  Ireland,  Pollard,  Steel,  and  Thompson  1977)  and  even  the  Upper  Carboniferous 
(Warme  and  Olson  1971;  Chamberlain  and  Clark  1973).  Fossil  thalassinoid  burrows  are  often 
referred  to  the  ichnogenus  Thalassinoides  (reviewed  by  Bromley  and  Frey  1974);  they  vary 
considerably  in  size  and  architecture,  according  to  the  specific  identity  of  the  burrower  (e.g.  see 
Braithwaite  and  Talbot  1972)  and  to  the  nature  of  the  sediments  (e.g.  see  Bromley  1967). 
Nevertheless,  it  is  possible  to  give  a brief  generalized  description  (based  on  Weimer  and  Hoyt  1964; 
Glaessner  1969;  and  sources  mentioned  above).  A thalassinoid  burrow  system  comprises  one  or  more 
chambers  (or  tunnels)  opening  to  the  sea  floor  by  a steeply  inclined  shaft  which  is  wide  enough  to 
permit  free  passage  of  the  burrower.  A second  shaft,  which  is  often  narrower  and  distinctly  tapered, 
extends  to  the  surface  from  another  part  of  the  chamber  system.  The  shrimp  occupies  one  of  the 
chambers  and  creates  a current  by  movements  of  its  swimmerets;  water  enters  the  wider  shaft  and  is 
expelled  through  the  narrower  shaft,  often  serving  to  flush  out  faecal  pellets.  The  chamber  occupied 
by  the  shrimp  is  big  enough  for  the  animal  to  turn  round  completely  in  somersault  fashion.  The 
various  chambers  are  connected  by  constricted  passages,  and  the  shrimp  may  excavate  several  short 
galleries  from  the  sides  of  a chamber.  In  some  cases  the  chambers  occur  at  random,  while  in  others 
they  may  form  a spiral  or  radial  pattern.  In  soft  sediments  the  shafts  may  be  deep,  and  there  may  be 
many  chambers  excavated  at  several  horizons;  in  more  resistant  sediments  the  burrow  system  may 
comprise  little  more  than  ramifying  tunnels  confined  to  a preferred  horizon.  When  such  burrow 
systems  are  encountered  as  fossils  they  are  commonly  filled  with  minerals  or  sediments  which  differ 
from  the  surrounding  rock. 

The  supposed  plesiosaur  embryos  are  readily  interpreted  as  thalassinoid  burrow  fillings  (see  text- 
fig.  2).  Their  ovoid  ‘bodies’  seem  to  represent  a series  of  chambers  arranged  in  a roughly  radial 
pattern,  while  their  ‘necks’  may  be  identified  as  steeply  inclined  shafts  connecting  chambers  at 
different  levels  or  leading  towards  the  former  sea  floor.  Some  of  the  ‘necks’  have  a distinctly  tapered 
tip  and  may  be  regarded  as  excurrent  shafts.  There  is  no  trace  of  a broader  incurrent  shaft,  but  this 
would  most  likely  have  been  situated  in  the  region  of  the  badly  damaged  ‘North  Pole’.  The  ‘tails’  are 
probably  burrow-ends,  which  are  ‘usually  somewhat  conical,  coming  to  a blunt  point’  in  thalassinoid 
burrows  from  the  English  Chalk  (Bromley  1967,  p.  162).  The  ‘limbs’  of  the  supposed  embryos  may  be 
identified  as  short  galleries  developed  from  the  walls  of  the  main  chambers. 

Nearly  all  major  features  of  the  supposed  embryos  can  be  matched  in  thalassinoid  burrows  of 
one  sort  or  another.  And  several  other  facts  would  seem  to  confirm  that  Seeley’s  specimen  is  a 
thalassinoid  burrow  system.  The  shale  matrix  of  the  specimen  contains  a few  tiny  spheroidal  objects 
(about  0-5  mm  in  diameter)  which  may  possibly  be  faecal  pellets  flushed  out  from  the  interior  of  the 
burrow;  several  of  these  occur  in  the  shale  around  embryo  4.  Next,  the  invertebrate  fauna  of  the 
Whitby  Lias  includes  a variety  of  small  decapod  crustaceans,  and  some  of  these  ( Glyphea  spp.  and 
Eryma  sp.)  are  similar  in  size  and  configuration  to  modern  burrowing  shrimps.  Blake  listed  seven 
species  of  crustaceans  from  the  Whitby  Lias,  and  mentioned  that  ‘numerous  crustacean  claws  occur 
in  many  zones’  (1876,  p.  429).  Glyphea  and  Eryma  are  not  close  relatives  of  the  living  thalassinoid 
shrimps,  but  ‘it  must  be  remembered  that  similar  burrows  indicate  possibly  similar  body  shape  and 
behaviour,  not  systematic  identity  of  the  burrower’  (Glaessner  1969,  p.  430).  Bromley  and  Frey 
(1974)  listed  many  crustaceans  know  to  construct  burrows  of  Thalassinoides  type,  and  among  these 
were  Jurassic  shrimps  of  the  superfamily  Glypheoidea  (Sellwood  1971;  Bromley  and  Asgaard  1972). 


THULBORN:  PLESIOSAUR  EMBRYOS’ 


357 


It  is  possible,  then,  that  one  or  more  of  the  crustaceans  reported  from  the  Whitby  Lias  (most  probably 
Glyphea  spp.)  may  have  constructed  thalassinoid  burrows.  Remains  of  crustaceans  are  rarely  found 
inside  thalassinoid  burrows,  probably  because  the  animals  left  their  burrows  to  moult  (see  Glaessner 
1969,  p.  434).  Finally,  burrows  are  abundant  at  some  horizons  in  the  Whitby  Lias.  The  distribution 
of  these  burrows  was  studied  by  Morris  (1979),  who  did  not  specify  the  occurrence  of  Thalassinoides 
but  mentioned  that  the  trace  fossil  assemblages  were  dominated  by  Chondrites.  In  other  sedimentary 
settings  Chondrites  may  be  associated  with  another  trace  fossil,  Gyrolithes,  which  is  sometimes  linked 
with  Thalassinoides  to  form  a single  burrow  system  (Bromley  and  Frey  1974).  This  very  indirect 
evidence  may  hint  at  the  existence  of  thalassinoid  burrows  in  the  Upper  Lias  at  Whitby;  such  an 
occurrence  would  not  be  surprising,  since  Thalassinoides  is  widespread  in  the  British  Lower  Lias 
(Sellwood  1970;  Sellwood,  Durkin,  and  Kennedy  1970).  It  is  worth  noting  that  Morris  found 
Chondrites  restricted  to  a ‘normal’  shale  facies — ‘a  homogeneous  bioturbated  sediment  often 
containing  sideritic  nodules  or  horizons’  (1979,  p.  117).  Morris  studied  only  the  lower  part  of  the 
Toarcian  succession  at  Whitby  (as  high  as  the  Hard  Shales),  so  that  his  ‘normal’  shales  are  in  fact 
represented  by  the  Grey  Shales.  It  has  already  been  deduced,  on  other  grounds,  that  the  supposed 
plesiosaur  embryos  came  from  the  Grey  Shales  or  from  the  Cement  Shales  (which  were  not 
considered  by  Morris).  In  other  words,  it  seems  probable  that  Seeley’s  specimen  came  from  a 
‘normal’  shale  horizon  in  which  burrows  are  abundant. 

In  summary,  the  nodule  interpreted  by  Seeley  (1896)  as  a group  of  plesiosaur  embryos  bears  a 
strong  resemblance  to  a system  of  thalassinoid  burrow  fillings.  Crustaceans  known  from  the  Whitby 
Lias  may  well  have  included  forms  capable  of  excavating  such  burrows;  and  burrows  (though  not 
specifically  identified  as  thalassinoid)  are  abundant  in  the  ‘normal’  shale  facies  from  which  the  nodule 
is  likely  to  have  come.  The  specimen  has  only  one  structural  feature  that  I have  been  unable  to  match 
in  any  other  thalassinoid  burrow:  this  is  the  distinct  ‘dorsal  ridge’  noted  by  Seeley.  If  the  specimen  is 
interpreted  as  a burrow  system  this  ‘dorsal  ridge’  will  actually  represent  a longitudinal  groove  in  the 
floor  of  each  burrow  chamber.  It  is  reasonable  to  suppose  that  this  groove  was  produced  by  the 
appendages  or  the  telson  of  the  crustacean  inhabitant. 

Thalassinoid  burrow  fillings  have  commonly  been  mistaken  for  quite  different  objects;  Bromley 
reported  (1967,  p.  158)  that  thalassinoid  burrows  of  Upper  Cretaceous  age  have  been  interpreted  as 
benthonic  algae,  plant  roots,  sponges,  solution  channels,  concretions,  and  trace  fossils  of  unidentified 
marine  organisms.  Seeley’s  misinterpretation  is  all  the  more  understandable  when  one  considers  that 
his  specimen  had  been  ‘improved’  by  artificial  means.  Parts  of  the  supposed  embryos  seem  to  have 
been  smoothed  off  and  polished  with  a mild  abrasive,  leaving  a microscopic  pattern  of  unnaturally 
straight  and  parallel  scratches.  And  in  places  it  appears  that  the  main  burrow  chambers  (or  embryo 
‘bodies’)  have  been  trimmed  to  a more  suggestive  shape  by  removal  of  some  side-galleries  (which 
would  otherwise  have  remained  as  rather  puzzling  supernumerary  ‘limbs’).  Such  trimming  is 
apparent  along  the  right  side  of  embryo  1 , where  there  is  a series  of  deep  needle-marks.  These  marks 
extend  from  the  fossil  into  the  shale  matrix,  whereas  needle-marks  produced  in  normal  preparation 
would  extend  in  the  opposite  direction.  Either  this  work  with  a needle  was  intended  to  remove 
portions  of  the  fossil,  or  it  was  an  unbelievably  clumsy  attempt  at  preparing  the  specimen.  Next,  the 
suitably  constricted  appearance  of  the  ‘neck’  in  embryo  2 (text-fig.  2b)  is  due  partly  to  the  fact  that  it 
has  been  cut  down  with  a knife-blade  or  similar  instrument.  The  cut  surface  is  easily  visible;  it  is  flat 
and  sharp-edged,  in  contrast  to  the  irregular  and  broadly  rounded  surfaces  elsewhere.  Finally  there  is 
evidence  of  a much  coarser,  but  none  the  less  effective,  ‘improvement’:  several  hammer  blows  in  the 
region  of  the  ‘North  Pole’  removed  any  trace  that  might  have  existed  of  an  incurrent  shaft  leading  to 
the  burrow  system.  It  is  not  altogether  surprising  to  discover  these  alterations  to  Seeley’s  specimen, 
for  many  such  ‘improved’  or  ‘repaired’  fossils  are  known  to  have  come  from  the  dealers  at  Whitby 
(see,  for  example,  Blake  1876,  p.  428;  Dance  1976,  p.  103). 

Acknowledgements.  I thank  Dr.  Angela  Milner  and  Dr.  Alan  Charig  for  providing  help  and  facilities  at  the 
British  Museum  (Natural  History).  Text-fig.  1 was  prepared  by  Mandy  Cilento  (Zoology  Department, 
University  of  Queensland),  and  Susan  Turner  (Queensland  Museum)  helped  me  in  searching  the  literature. 


358 


PALAEONTOLOGY,  VOLUME  25 


REFERENCES 

abel,  o.  1935.  Vorzeitliche  Lebensspuren,  Gustav  Fischer,  Jena.  644  pp. 

amstutz,  G.  c.  1958.  Coprolites:  a review  of  the  literature  and  a study  of  specimens  from  southern  Washington. 
J.  sediment,  petrol.  28,  498-508. 

blake,  j.  f.  1876.  Palaeontology.  In  tate,  r.  and  blake,  j.  f.  The  Yorkshire  Lias.  Van  Voorst,  London,  pp.  243- 
475. 

braithwaite,  c.  j.  r.  and  talbot,  m.  r.  1972.  Crustacean  burrows  in  the  Seychelles,  Indian  Ocean.  Palaeogeogr., 
Palaeoclimatol.,  Palaeoecol.  11,  265-285. 

bromley,  r.  G.  1967.  Some  observations  on  burrows  of  thalassinidean  Crustacea  in  chalk  hardgrounds.  Q.  Jl 
geol.  Soc.  Lond.  123,  157-182. 

— and  asgaard,  u.  1972.  Notes  on  Greenland  trace  fossils.  II.  The  burrows  and  microcoprolites  of  Glyphea 
rosenkrantzi,  a Lower  Jurassic  palinuran  crustacean  from  Jameson  Land,  East  Greenland.  Rept  geol.  Surv. 
Greenland , 49,  15-21. 

— and  frey,  r.  w.  1974.  Redescription  of  the  trace  fossil  Gyrolithes  and  taxonomic  evaluation  of 
Thalassinoides,  Ophiomorpha  and  Spongeliomorpha.  Bull.  geol.  Soc.  Denmark,  23,  311-335. 

chamberlain,  c.  k.  and  clark,  d.  l.  1973.  Trace  fossils  and  conodonts  as  evidence  for  deep-water  deposits  in  the 
Oquirrh  Basin  of  central  Utah.  J.  Paleont.  47,  663-682. 
dance,  p.  1976.  Animal  fakes  and  frauds.  Sampson  Low,  Maidenhead.  128  pp. 
edinger,  t.  1929.  Die  fossilen  Gehirne.  Ergebn.  d.  Anat.  28,  1-249. 

fiege,  k.  1944.  Lebensspuren  aus  dem  Muschelkalk  Nordwestdeutschlands.  N.  Jb.  Miner.  Geol.  Palaont.  Abh. 
B88,  401-426. 

GLAESSNER,  M.  F.  1969.  Decapoda.  In  moore,  r.  c.  (ed.),  Treatise  on  Invertebrate  Paleontology,  Part  R, 
Arthropoda  ( Crustacea  exclusive  of  Ostracoda,  Myriapoda,  Hexapoda ).  Geol.  Soc.  Amer.  and  Kansas  Univ., 
pp.  400-533. 

hallam,  a.  1962.  A band  of  extraordinary  calcareous  concretions  in  the  Upper  Lias  of  Yorkshire,  England. 
J.  sediment,  petrol.  32,  840-847. 

halstead,  L.  b.  1969.  The  pattern  of  vertebrate  evolution.  Oliver  and  Boyd,  Edinburgh.  209  pp. 

— and  Middleton,  j.  1972.  Bare  bones,  an  exploration  in  art  and  science.  Oliver  and  Boyd,  Edinburgh.  1 19  pp. 
hauff,  b.  1921.  Untersuchung  der  Fossilfundstatten  von  Holzmaden  im  Posidonienschiefer  des  oberen  Lias 

Wiirttemburgs.  Palaeontographica,  64,  1 -42. 

hemingway,  j.  e.  1974.  Jurassic.  In  rayner,  d.  h.  and  Hemingway,  j.  e.,  The  geology  and  mineral  resources  of 
Yorkshire.  Yorks,  geol.  Soc.,  Leeds,  pp.  161-223. 

hoffmann,  j.  1958.  Einbettung  und  Zerfall  der  Ichthyosaurier  im  Lias  von  Holzmaden.  Meyniana,  6,  10-55. 
horner,  j.  r.  and  makela,  r.  1979.  Nest  of  juveniles  provides  evidence  of  family  structure  among  dinosaurs. 
Nature , Lond.  282,  296-298. 

howarth,  m.  k.  1962.  The  Jet  Rock  Series  and  Alum  Shale  Series  of  the  Yorkshire  Coast.  Proc.  Yorks,  geol.  Soc. 
33,381-422. 

1973.  The  stratigraphy  and  ammonite  fauna  of  the  Upper  Liassic  Grey  Shales  of  the  Yorkshire  coast.  Bull. 

Brit.  Mus.  ( nat . Hist.),  Geol.  24,  238-277. 

Ireland,  r.  j.,  pollard,  j.  e.,  steel,  R.  J.  and  Thompson,  d.  b.  1977.  Intertidal  sediments  and  trace  fossils  from  the 
Waterstones  (Scythian -Anisian?)  at  Daresbury,  Cheshire.  Proc.  Yorks,  geol.  Soc.  41,  399-436. 
jerison,  h.  j.  1973.  Evolution  of  the  brain  and  intelligence.  Academic  Press,  New  York  and  London. 
morris,  K.  a.  1979.  A classification  of  Jurassic  marine  shale  sequences:  an  example  from  the  Toarcian  (Lower 
Jurassic)  of  Great  Britain.  Palaeogeogr.,  Palaeoclimatol.,  Palaeoecol.  26,  117-126. 
pantin,  H.  m.  1958.  Rate  of  formation  of  a diagenetic  calcareous  concretion.  J.  sediment,  petrol.  28,  366-371 . 
robinson,  J.  a.  1975.  The  locomotion  of  plesiosaurs.  N.  Jb.  Geol.  Palaont.  Abh.  149,  286-332. 
saint-seine,  p.  de.  1955.  Sauropterygia.  In  piveteau,  j.  (ed.),  Traite  de  Paleontologie,  V ( Amphibiens , Reptiles, 
Oiseaux).  Masson,  Paris,  pp.  420-458. 

seeley,  h.  g.  1888.  On  the  mode  of  development  of  the  young  in  Plesiosaurus.  Rept.  Brit.  Assn  Adv.  Sci., 
Manchester  1887,  Section  C,  697-698.  [Abstract;  a second,  edited,  version  is  in  Geol.  Mag.  (3),  4,  562- 
563.] 

— 1 896.  On  a pyritious  concretion  from  the  Lias  of  Whitby,  which  appears  to  show  the  external  form  of  the 
body  of  embryos  of  a species  of  Plesiosaurus.  Ann.  Rept  Yorks,  philos.  Soc.  1895,  20-29. 

sellwood,  b.  w.  1970.  The  relation  of  trace  fossils  to  small  scale  sedimentary  cycles  in  the  British  Lias.  In  crimes, 
t.  p.  and  harper,  j.  c.  (eds.).  Trace  Fossils  (Geol.  Soc.  London,  Special  Issues,  No.  3).  Seel  House  Press, 
Liverpool,  pp.  489-504. 


THULBORN:  PLESIOSAUR  EMBRYOS’ 


359 


sellwood,  B.  w.  1971.  A Thalassinoides  burrow  containing  the  crustacean  Glyphea  udressieri  (Meyer)  from  the 
Bathonian  of  Oxfordshire.  Palaeontology , 14,  589-591. 

— durkin,  m.  K.  and  Kennedy,  w.  j.  1970.  Field  meeting  on  the  Jurassic  and  Cretaceous  rocks  of  Wessex: 
report  by  the  directors.  Proc.  geol.  Assn .,  81,  715-732. 

urlichs,  m.,  wild,  r.  and  ziegler,  b.  1979.  Fossilien  aus  Holzmaden.  Stuttgarter  Beit,  zur  Naturk.  C 11,  1-34. 
walls,  G.  l.  1942.  The  vertebrate  eye  and  its  adaptive  radiation.  Bull.  Cranbrook  Inst.  Sci.  19,  1-785. 
warme,  J.  E.  and  olson,  r.  w.  1 97 1 . Lake  Brown  wood  spillway.  In  perkins,  b.  (ed.),  Trace  fossils:  a field  guide. 

Louisiana  State  Univ.  School  Geosci.,  misc.  Publns  No.  71.  pp.  27-43. 
weimer,  r.  j.  and  hoyt,  j.  h.  1964.  Burrows  of  Callianassa  major  Say,  geologic  indicators  of  littoral  and  shallow 
neritic  environments.  J.  Paleont.  38,  761-767. 

williston,  s.  w.  1914.  Water  reptiles  of  the  past  and  present.  Chicago  Univ.  Press,  Chicago.  251  pp. 
woodward,  A.  s.  1898.  Outlines  of  vertebrate  palaeontology  for  students  of  zoology.  Cambridge  Univ.  Press, 
Cambridge.  470  pp. 


RICHARD  A.  THULBORN 


Original  typescript  received  7 January  1981 
Revised  typescript  received  14  April  1981 


Department  of  Zoology 
University  of  Queensland 
St.  Lucia 
Queensland  4067 
Australia 


LIMPET  GRAZING  ON  CRETACEOUS 
ALGAL-BORED  AMMONITES 

by  ETIE  BEN  AKPAN,  GEORGE  E.  FARROW,  and  NOEL  MORRIS 


Abstract.  Phosphatized  internal  moulds  of  Anahoplites  sp.  and  Euhoplites  sp.  reveal  sets  of  six-toothed  radula 
marks  closely  comparable  to  those  scratched  on  to  Recent  molluscs  by  Acmaea  ( Tectura ) virginea  in  water 
depths  of  up  to  25  m on  the  muddy  inshore  shelf  of  western  Scotland.  Both  the  Recent  and  Cretaceous  acmaeid 
marks  truncate  algal  borings.  Depth  of  the  Gault  sea  during  the  Spathi  subzone  (M.  Albian)  was  probably 
between  8 and  30  m,  i.e.  within  the  euphotic  zone.  Subsequent  to  lithification  and  phosphatization  the  moulds 
were  bored  by  phoronids  and  scraped  by  regular  echinoids. 

Farrow  and  Clokie  (1979)  have  recently  drawn  attention  to  the  close  association  between  the 
Recent  limpet  Acmaea  ( Tectura ) virginea  (Muller  1776)  and  the  shell-boring  alga  ‘conchocelis’  on 
which  it  feeds.  ‘Conchocelis’  is  the  resting  phase  of  red  seaweeds  of  the  family  Bangiaceae:  use  of  the 
term  is  as  in  Farrow  and  Clokie  ( 1 979).  In  this  paper  we  describe  a similar  association  from  the  Lower 
Cretaceous  which  gives  an  indication  of  the  depth  of  deposition  of  part  of  the  Gault  Clay  and 
additionally  demonstrates  the  persistence  of  acamaeid  radula  design.  The  fossil  material  consists  of 
both  shells  and  ammonite  steinkerns  bearing  grazing  traces:  nearly  all  come  from  the  Spathi  Subzone 
of  the  Middle  Albian,  only  one  specimen  from  Dunton  Green  being  of  uncertain  horizon.  The 
grazing  traces  were  collected  from  Copt  Point,  Folkestone,  and  are  found  on  ten  out  of  fifty-eight 
phosphatic  internal  moulds  of  hoplitid  ammonites. 

We  outline  the  evidence  for  our  interpretation  of  these  traces,  and  document  their  occurrence 
alongside  the  only  known  species  of  Acmaeidae  from  the  British  Gault. 

CHARACTERISTICS  OF  THE  ASSOCIATED  FAUNA 

The  form  of  the  grazing  traces  from  the  Gault  with  their  six  parallel  incised  grooves  preserved  as 
negatives  has  led  us  to  conclude  that  they  probably  belonged  to  an  acmaeid.  Patelliform  shells  are 
uncommon  in  the  north-west  European  Albian,  and  particularly  so  in  the  Gault  facies.  There  is  one 
exception,  however,  Patella  tenuistriata  Michelin,  1838.  This  species  is  recorded  from  Folkestone  by 
Price  (1879)  from  his  beds  2 II,  2 V,  2 VI,  2 VII,  and  7 VIII,  that  is,  scattered  through  the  Middle 
Albian.  We  have  examined  nine  specimens  from  the  Gault  at  Folkestone  with  no  further  locality 
information.  Of  these,  two  are  preserved  on  ammonite  or  nautiloid  shells.  Their  protoconchs  appear 
to  have  been  lost  during  life  and  no  epifauna  or  epiphyte  is  apparent.  Further  specimens  of  particular 
interest  are  two  specimens  from  the  Hampden  Park  borehole,  Sussex,  from  the  Spathi  Subzone,  the 
same  subzone  as  our  grazing  trails.  Another  specimen,  from  an  uncertain  horizon  of  the  Gault  at 
Dunton  Green,  Kent  (IGS  Z5013),  has  a well-preserved  muscle  scar  which  confirms  their 
identification  as  Acmaeidae.  The  only  other  limpet-like  forms  from  the  British  Gault  are  attributed  to 
the  Calyptraeacea.  Since  these  are  suspension  feeders,  they  are  unlikely  to  have  produced  the  incised 
traces  we  describe. 

The  clays  of  the  Spathi  Subzone  occur  widely  across  south-east  England  and  have  an  extremely 
restricted  benthonic  fauna  preserved.  We  are  indebted  to  Adrian  Morter  of  the  I.G.S.  for  information 
concerning  the  faunas  of  this  horizon,  put  together  from  information  from  several  boreholes  in  East 
Anglia.  Morter’s  Bed  3 contains  common  Birostrina  concentrica  and  'O  steed  cf.  incurva,  and  less 
commonly  the  brachiopod  Moutonithyris,  the  gastropods  Anticonulus  and  Rissoina,  the  bivalves 
Entolium,  Inoceramus,  Pycnodonte,  Ludbrookia , Neithea,  Pectinucula , and  the  echinoid  Hemiaster. 


(Palaeontology,  Vol.  25,  Part  2,  pp.  361-367.] 


362 


PALAEONTOLOGY,  VOLUME  25 


Morter’s  Bed  2,  a grey  clay  bioturbated  with  Chondrites  and  other  traces,  contains  the  bivalves 
Birostrina,  Inoceramus,  Mesosacella,  Pectinucula,  Pseudolimea,  Pycnodonte , Rastellum,  the 
brachiopods  Kingena , Moutonithyris,  and  Tamarella,  and  the  solitary  coral  Cyclocyathus.  These 
faunas  are  typical  of  the  Spathi  Subzone  benthos.  One  or  two  additions  are  known  from  Folkestone, 
including  Oistotrigonia  fittoni  (Deshayes),  which  seems  to  be  restricted  to  the  clay  facies. 

We  suggest  that  the  fauna  was  limited  by  the  fine-grained  sediments.  These  were,  however,  not 
always  oxygen-deficient,  as  nuculoid  infaunal  deposit  feeders  form  a significant  part  of  the  fauna. 
Shells  of  nektonic  cephalopods  are  present  in  sufficient  numbers  to  have  provided  a substrate  for 
epifaunal  grazers  and  this  would  have  been  the  case  for  the  acmaeids.  The  trochaceans  and 
rissoaceans  may  also  have  browsed  on  shell  fragments,  or  storm-derived  algae  or  sponges. 

We  feel  that  the  coincidence  of  the  occurrence  of  Acmaea-Mke  grazing  traces  and  the  records  of  the 
only  known  species  of  Acmaeidae  from  the  British  Middle  Albian  is  sufficient  to  justify  our 
interpretation  as  one  being  the  product  of  the  other. 

SYSTEMATIC  DESCRIPTION 
Acmaea  tenuistriata  (Michelin,  1838) 

Material.  Shells:  nine  specimens  from  the  Gault  of  Folkestone  (M.-U.  Albian)  BM(NH)  GG.  19971  -19978;  two 
specimens  from  the  Spathi  Subzone,  M.  Albian,  Hampden  Park  borehole,  Sussex,  IGS,  BDL  6727  and  6731; 
one  specimen  from  an  unspecified  nodule  bed  in  the  Gault,  Dunton  Green,  Kent,  IGS  Z5013. 

Description.  Broad,  oval,  bilaterally  symmetrical,  limpet-like  shells  with  the  apex  approximately  half-way 
between  the  mid-point  and  the  anterior  margin.  The  height  is  approximately  one-third  of  the  length.  The  largest 
specimen  is  about  15  mm  in  length.  There  is  a radiating  ornament  of  fine  ribs  with  wide  inter-spaces.  No 
protoconchs  are  preserved.  The  shell  structure  is  partly  preserved  in  two  specimens,  BM(NH)  GG.  19977-19978 
(text-fig.  1 d),  and  consists  of  concentrically  arranged  cross  lamellae  which  appear  like  vertical  prisms  in  radial 
section.  The  shell  is  preserved  entirely  in  a pinkish  form,  typical  of  Gault  shells  originally  formed  entirely  of 
aragonite.  The  body  and  foot  attachment  scar  is  well  preserved  in  IGS  Z5013  and  is  illustrated  in  text-fig.  la-c.  It 
consists  of  a broad  horseshoe-shaped  area  of  scar  joined  at  its  anterior  by  a thinner  line. 


text-fig.  1 . Acmaea  tenuistriata  (Michelin).  a-c  steinkern  with  muscle  attachment  scar  inked  in: 
IGS  Z5013,  unspecified  nodule  bed,  Gault,  Dunton  Green,  x4.  ( a ) view  of  left  side,  ( b ) dorsal, 
(c)  view  of  anterior.  ( d ) specimen  with  part  of  the  shell  preserved:  BM(NH)  GG.  19977,  Middle- 
Upper  Albian,  Copt  Point,  Folkestone,  x 3,  dorsal  view. 


AKDAN  ET  AL.:  LIMPET  GRAZING 


363 


Discussion.  The  shape  of  this  shell  and  the  form  of  its  muscle  scar  are  characteristic  of  the  Family 
Acmaeidae.  The  general  shape  of  this  species,  with  its  apex  placed  anterior  to  the  mid-point,  and  its 
fine  radial  striae,  are  reminiscent  of  the  subgenus  Acmaea  ( Tectura ) Gray,  type  species  A.  virginea 
(Muller),  quoted  as  Patella  parva  da  Costa  (jun.  sub.  syn.)  rather  than  Acmaea  s.s.,  type  species 
A.  mitra  (von  Eschscholtz),  a Californian  species  with  a taller  shell  and  central  apex.  However, 
without  a thorough  revision  of  the  fossil  members  of  the  family  we  feel  unable  to  place  this  species 
in  a particular  subgenus. 

Grazing  traces 

Seven  specimens  are  in  the  BM(NH);  four  collected  by  EBA  and  GEF,  GG  19979-19982,  and  three 
collected  by  H.  Lister,  GG  19983-19985.  The  grazing  traces  are  preserved  as  exquisitely  detailed 
protruberances  on  the  surfaces  of  the  phosphatic  moulds  (PI.  37,  figs.  2-4).  Individual  markings,  up 
to  200  ju.m  wide  by  400  jum  long  when  fully  developed,  consist  of  six  relatively  broad  ridges  separated 
by  narrow  furrows.  Overlapping  of  adjacent  marks,  however,  rarely  permits  the  full  complement  of 
ridges  to  be  appreciated,  and  sets  of  sub-parallel  ridges  are  the  norm,  their  orientation  being  more 
consistent  on  some  specimens  (PI.  37,  fig.  4)  than  on  others  (PI.  37,  fig.  3).  Successive  truncation  of  sets 
of  ridges  suggests  at  least  three  phases  of  trace  activity,  the  lowest  being  poorly  preserved  and  having 
a smoothed  appearance. 

Borings 

Three  types  seem  to  be  present.  The  first  consists  of  a widespread  fine  granulation  on  the  surface  of 
moulds  where  radula  traces  occur,  in  the  form  of  small  upstanding  mounds  of  circular  cross-section 
(PI.  37,  fig.  4).  These  represent  the  infilling  of  borings  into  the  original  ammonite  shell.  Having  an 
approximate  diameter  of  6 fxm  they  are  probably  algal  in  origin. 

The  second  and  third  types  represent  unfilled  borings  into  the  lithified  phosphatic  moulds.  The 
former  are  oval  in  cross-section,  500  jum  in  diameter,  and  penetrate  deeply  into  the  moulds:  the  latter 
are  tunnels  very  close  to  the  surface  of  the  moulds  which  clearly  truncate  grazing  traces  in  places 
(PI.  37,  fig.  3).  The  tunnels  measure  about  100  in  depth  and  between  180  and  300  /urn  in  maximum 
diameter,  and  are  visible  for  a few  millimetres  before  disappearing,  only  to  reappear.  These  are 
probably  the  work  of  phoronids  (Bromley  1970). 


COMPARISON  WITH  RECENT  TRACES 

The  Gault  traces  are  similar  to  Recent  A.  ( Tectura ) virginea  radula  marks  incised  into  mollusc  shells 
(PI.  37,  fig.  1)  but  are  preserved  in  a negative  form.  The  comparison  extends  from  over-all  similarity  of 
shape  to  details  of  six  broadly  incised  elements,  demonstrating  a closer  affinity  in  radula  design 
between  the  presumed  Gault  acmaeid  and  the  Recent  A.  (77)  virginea  than  between  living  A.  virginea 
and  A.  tessulata  (Muller  1776),  whose  radula  is  of  four-pronged  construction. 

However,  the  Gault  grazing  traces  are  wider  and  less  consistent  in  orientation  than  are  those  of  the 
Recent  A.  ( T .)  virginea.  This  size  difference  is  a reflection  of  the  sizes  of  the  causative  limpets.  For 
example,  the  longest  Gault  acmaeid  measures  15  mm  whereas  the  Recent  British  A.  ( T .)  virginea  is 
scarcely  more  than  7 mm  long.  Variation  in  the  size  of  Recent  Acmaea  dictates  the  width  of  the 
grazing  traces  produced,  ‘scoops’  ranging  from  80  to  100  /urn  wide. 


DEPTH  OF  DEPOSITION  OF  THE  GAULT  CLAY 

The  similarity  between  the  fossil  and  Recent  grazing  traces  is  the  more  remarkable  on  account  of  the 
further  association  between  the  grazing  marks  and  the  fine  granulation  seen  on  the  Gault 
specimens — the  negative  of  the  algal  borings  seen  pitting  the  Recent  examples.  It  is  this  association 
that  sheds  light  on  the  depositional  environment  of  the  Gault  during  the  Spathi  Subzone,  and  in  order 
to  assess  the  probable  depth  of  water  it  is  necessary  to  consider  something  of  the  ecology  of  living 
Acmaea. 


364 


PALAEONTOLOGY,  VOLUME  25 


The  genus  Acmaea  occurs  widely  in  intertidal  and  shallow  water-shelf  habitats  and  there  is  much 
evidence  for  an  algal  grazing  habit.  Craig  (1968)  has  described  A.  pelta  living  on  micro-  and  macro- 
algae. Black  (1976)  described  A.  incessa  living  exclusively  on  the  kelp  Egregia,  and  many  intertidal 
species  graze  upon  epilithic  algae  (Kozloff  1973;  Carefoot  1976).  Fretter  and  Graham  (1976) 
recorded  A.  virginea  on  algal-coated  shell  debris  off  the  British  coast.  While  these  observations 
suggest  that  the  depth  was  within  the  photic  zone,  it  is  possible  to  make  a more  refined  estimate  by 
looking  in  more  detail  at  a single  area. 

Living  Acmaea  virginea  on  the  Scottish  continental  shelf 

Farrow  and  Clokie  (1979,  Table  1)  working  in  the  Firth  of  Clyde,  found  A.  virginea  down  to  only 
1 8 m,  coincident  with  the  local  limit  of  the  euphotic  zone  as  measured  biologically  by  the  last  record  of 
‘conchocelis’  (Clokie,  Boney,  and  Farrow  1979).  This  correlation  is  well  established  regionally, 
Acmaea  extending  into  deeper  water  where  substrates  are  coarser,  environmental  energy  higher,  and 
water  clearer,  as  off  the  Orkney  Islands  where  the  euphotic  limit  is  38  m (Farrow,  Allen,  Akpan,  and 
Brown  1 98 1 ).  A muddier,  Firth-of-Clyde  or  Firth-of-Lorne-type  environment  is  more  appropriate  as 
a Gault  analogy,  although  the  present  latitude  of  56°  N is  further  north  than  that  of  Folkestone  in  the 
Albian,  which  was  more  like  33°  N (Smith,  Briden,  and  Drewry  1973,  fig.  7,  map  2). 

Evidence  on  A.  virginea  distribution  in  one  area  of  the  Firth  of  Clyde  is  summarized  on  text-fig.  2.  It 
is  clear  that  the  limpet  extends  down  to  within  a few  metres  of  the  euphotic  limit.  Since,  however, 
Acmaea  grazing  is  intensive  only  where  algal  infestation  is  heavy  we  may  conclude  that  deposition  of 
the  Gault  Clay  during  this  part  of  the  Spathi  Subzone  was  well  within  the  euphotic  zone.  Depth  was 
unlikely  to  have  been  greater  than  30  m and  could  even  have  been  as  shallow  as  8 m. 

TAPHONOMY 

The  evidence  is  clear  that  these  Gault  ammonites  were  infested  with  shell-boring  algae  to  such  an 
extent  that  the  full  thickness  of  shell  must  have  been  riddled  with  them,  for  the  limpet  grazing  marks 
were  made  on  the  inside  of  the  body  chambers:  any  on  the  outside  would  have  had  negligible 
fossilization  potential.  It  remains  to  be  established,  however,  whether  the  algae  infested  dead  shells 
drifting  near  the  ocean  surface,  maybe  for  several  years  (House  1973,  p.  309;  though  cf.  Reyment 
1958),  or  whether  they  bored  into  the  shells  when  the  latter  were  on  the  sea  bed.  Since  the 
environmental  interpretation  for  this  part  of  the  Gault  hinges  on  this  point,  further  consideration  is 
necessary.  There  is  some  supporting  circumstantial  evidence. 

First,  there  is  the  occurrence  of  probable  regular  echinoid  scratches  both  on  the  phosphate  moulds 
and  on  associated  shell  debris:  this  scratching  is  commonest  today  within  the  euphotic  zone. 
However,  the  occurrence  of  these  traces  on  the  ammonite  moulds  post-dates  not  only  the  grazing 
marks  but  also  lithification  of  the  moulds,  and  could  be  ruled  misleading.  Second,  it  is  to  be  doubted 
whether  the  ‘conchocelis’  phase  of  members  of  the  family  Bangiaceae  (coastal  red  algae)  would  be 
sufficiently  dispersed  to  infest  a multiplicity  of  widely  scattered  floating  cephalopods.  It  seems  easier 


EXPLANATION  OF  PLATE  37 

Fig.  1.  Grazing  traces  of  Recent  Acmaea  ( Tectura ) virginea  on  algal  bored  Dosinia  shell:  9 m:  Isle  of  Bute,  Firth 
of  Clyde,  Scotland.  For  locality  see  text-fig.  2. 

Fig.  2.  Grazing  traces  of  Lower  Cretaceous  acmaeid  on  surface  of  hoplitid  ammonite  skeinkern:  note  the  grazing 
front  (cf.  fig.  1):  BM(NH)  GG.19983:  Spathi  Subzone,  Middle  Albian,  Copt  Point,  Folkestone. 

Fig.  3.  Phoronid  boring  cutting  through  acmaeid  grazing  traces  on  surface  of  hoplitid  ammonite  steinkern: 
BM(NH)  GG.  19980:  Spathi  Subzone,  Middle  Albian,  Copt  Point,  Folkestone. 

Fig.  4.  Grazing  traces  of  Lower  Cretaceous  acmaeid  on  surface  of  hoplitid  ammonite  steinkern:  note  their 
relatively  consistent  orientation  (cf.  fig.  3):  BM(NH)  GG.  19979:  Spathi  Subzone,  Middle  Albian,  Copt  Point, 
Folkestone. 


PLATE  37 


AKPAN  et  al..  Limpet  grazing 


366 


PALAEONTOLOGY,  VOLUME  25 


• with  Acmaea  GRAZING  MARKS 

text-fig.  2.  Depth  of  occurrence  of  Recent  shells  showing  algal 
borings  and  Acmaea  grazing  marks:  Firth  of  Clyde,  Scotland. 
Note  that  the  limit  of  grazing  is  slightly  shallower  than  the  limit 
of  algal  boring  by  ‘conchocelis’  (taken  as  a biological  indication 
of  the  limits  of  the  euphotic  zone). 


to  envisage  punctured  shells  being  concentrated  on  the  sea  bed  by  winnowing  and  wave  action,  thus 
providing  a typical  habitat  for  the  acmaeids,  which  are  unknown  pseudo-planktonically. 

The  following  taphonomic  sequence  is  therefore  suggested: 

1 . Damaged  ammonite  shells  sink  on  to  muddy  sea  floor  and  are  colonized  by  boring  algae  on  their  inner  and 
outer  surfaces  (damage  is  supported  by  incomplete  preservation:  Reyment  1958;  Seilacher  1971). 

2.  Acmaeid  limpets  colonize  the  dead  shells,  feeding  on  the  endolithic  algae  in  ‘sweeping’  fashion  where 
infestation  is  high,  but  with  more  clearly  defined  ‘scoops’  where  infestation  is  lower. 

3.  Draught  filling  of  very  fine  mud  replicates  the  body  chamber  before  the  limpet  grazing  has  reached  such  an 
advanced  stage  that  the  weakened  shell  falls  apart:  many  ammonites  may  have  been  grazed  to  destruction  in  the 
absence  of  an  intervening  smothering  of  mud.  (Attempts  artificially  to  replicate  modem  grazing  traces  with 
Epotex  resin  have  been  markedly  less  successful  than  Nature’s  amazingly  high-fidelity  moulds  in  the  Gault!) 

4.  Burial,  lithification,  and  phosphatization  (with  dissolution  of  aragonite?). 

5.  Re-exhumation  by  winnowing  of  mud:  concentration  of  nodules  without  transportation. 

6.  Boring  by  polychaetes?  and  phoronids:  biting  by  regular  echinoids. 

CONCLUSIONS:  BORING  ALGAE  AND  MOLLUSCAN  GRAZING  TRACES 
IN  THE  FOSSIL  RECORD 

The  long  geological  record  of  both  the  boring  algae  (e.g.  Palaeoconchocelis  starmachii  from  the 
Silurian  of  Poland,  Campbell  1980)  and  the  patellaceans  (Trias-Rec.)  suggests  that  earlier  radula 
marks  may  soon  come  to  light.  Boekschoten  (1967)  and  Voigt  (1977)  have  recorded  a range  of 


AKPAN  ET  AL.:  LIMPET  GRAZING 


367 


molluscan  traces  as  far  back  as  the  upper  Jurassic,  with  chiton  marks  more  frequently  encountered 
than  those  of  limpets.  Taylor  (1981,  pi.  2/3)  has  figured  what  are  clearly  acmaeid  radula  marks  on  an 
Isognomon  from  the  Portlandian,  kerberus  Zone.  Again  they  seem  to  be  associated  with  an  algal- 
bored  shell. 

The  implications  of  the  kind  of  approach  reported  in  this  paper  are  twofold.  First,  it  shows  that 
when  reasonably  stringent  sampling  programmes  are  put  into  effect  on  present-day  continental  shelves 
it  is  possible  to  refine  the  determination  of  palaeoecological  parameters,  in  this  instance  water  depth. 
As  this  parameter  is  a difficult  one  to  assess  in  clay  sequences,  this  is  a step  forward,  and  the  result  for 
the  Spathi  Subzone  of  the  Gault  of  8-30  m may  come  as  something  of  a surprise.  The  second 
implication  lies  in  the  evidence  which  is  afforded  for  a remarkable  lack  of  change  in  radula  design  or 
pattern  of  feeding  in  the  acmaeids  since  the  Lower  Cretaceous. 


black,  r.  1976.  The  effect  of  grazing  by  limpet  Acmaea  insessa  on  the  kelp,  Egregia  laevigata  in  the  intertidal 
zone.  Ecology,  57,  265-277. 

boekschoten,  G.  s.  1967.  Palaeoecology  of  some  Mollusca  from  the  Tielrode  sands  (Pliocene,  Belgium). 
Palaeogeogr.,  Palaeoclimat.,  Palaeoecol.,  3,  311-362. 

bromley,  R.  G.  1970.  Borings  as  trace  fossils  and  Entobia  cretacea  Portlock,  as  an  example.  In  T.  p.  crimes  and 
j.  c.  harper  (eds.)  Trace  fossils.  Geol.  J.  Spec.  Issue,  3,  49-90. 

Campbell,  s.  e.  1980.  Palaeoconchocelis  starmachii,  a carbonate  boring  micro-fossil  from  the  Upper  Silurian  of 
Poland  (435  million  years  old):  implications  for  the  evolution  of  Bangiaceae  (Rhodophyta).  Phycologia,  19, 
25-36. 

carefoot,  t.  1976.  Pacific  seashores.  J.  J.  Douglas.  Vancouver. 

clokie,  J.,  boney,  a.  d.  and  farrow,  G.  e.  1979.  Use  of  Conchocelis  as  an  indicator  organism:  data  from  Firth  of 
Clyde  and  N.W.  Shelf.  Br.  Phycol.  J.  14,  120-121. 

craig,  p.  a.  1968.  The  activity  pattern  and  food  habits  of  the  limpet  Acmaea  pelta.  Veliger,  11  ( Suppl .),  13-19. 
farrow,  G.  e.  and  clokie,  j.  1979.  Molluscan  grazing  of  sub-littoral  algal-bored  shells  and  the  production  of 
carbonate  mud  in  the  Firth  of  Clyde,  Scotland.  Trans.  R.  Soc.  Edinb.  70,  139-148. 

— allen,  n.  h.,  akpan,  e.  b.  and  brown,  b.  j.  1981.  Bioclastic  carbonate  sedimentation  and  biofacies  on  a 
high-latitude,  tide-dominated  shelf:  NE  Orkney  Islands,  Scotland.  [In  prep.] 
fretter,  v.  and  graham,  a.  1976.  The  Prosobranch  molluscs  of  Britain  and  Denmark.  Pleurotomariacea, 
Fissurellacea  and  Patellacea.  J.  Mollusc.  Stud.,  Suppl.  1,  36  pp. 
house,  M.  r.  1973.  An  analysis  of  Devonian  goniatite  distribution.  Spec.  Pap.  Palaeont.  12,  305-317. 
kozloff,  e.  m.  1973.  Seashore  life  of  Puget  Sound,  the  Strait  of  Georgia,  and  the  San  Juan  Archipelago.  Seattle  and 
London:  Univ.  Washington  Press.  282  pp.  +28  pis. 

michelin,  h.  1838.  Sur  une  argile  dependant  du  Gault  observee  au  Gaty,  Communede  Gerodot,  Departement  de 
L’Aube.  Mem.  Soc.  Geol.,  Paris,  3,  97-103. 
price,  f.  G.  h.  1879.  The  Gault,  viii  + 81  pp.  London. 

reyment,  r.  a.  1958.  Some  factors  in  the  distribution  of  fossil  cephalopods.  Stockh.  Contr.  Geol.  1,  98-185. 
seilacher,  a.  1971.  Preservational  history  of  ceratite  shells.  Palaeontology,  14,  16-21. 
smith,  a.  g.,  briden,  j.  c.  and  drewry,  G.  e.  1973.  Phanerozoic  world  maps.  Spec.  Pap.  Palaeont.  12,  1-42. 
taylor,  p.  d.  1981.  Bryozoa  of  British  Portland  beds  (Upper  Jurassic).  Palaeontology,  24, 863-875,  pis.  121-122. 
voigt,  e.  1977.  On  grazing  traces  produced  by  the  radula  of  fossil  and  Recent  gastropods  and  chitons.  In  t.  p. 
crimes  and  j.  c.  harper  (eds.)  Trace  fossils  2,  Geol.  J.  Spec.  Issue,  9,  335-346. 


REFERENCES 


ETIE  B.  AKPAN 
GEORGE  E.  FARROW 


Department  of  Geology 
University  of  Glasgow 
Glasgow  G12  8QQ,  Scotland 


Typescript  received  9 October  1980 
Revised  typescript  received  10  February  1981 


NOEL  MORRIS 

Department  of  Palaeontology 
British  Museum  (Natural  History) 
Cromwell  Road 
London  SW7  5BD,  England 


' 


A REVIEW  OF  RECENT  AND  QUATERNARY 
ORGANIC-WALLED  DINOFLAGELLATE  CYSTS 
OF  THE  GENUS  PROTO PERIDINIUM 

by  REX  HARLAND 


Abstract.  A review  is  given  of  Recent  and  Quaternary  organic-walled  dinoflagellate  cysts  belonging  to  the 
genus  Protoperidinium  Bergh  emend.  Balech  1974.  There  is  a similarity  between  the  taxonomies  based  upon 
thecal  and  cyst  morphologies,  the  major  difference  being  one  of  hierarchy.  A scheme  is  presented  which  attempts 
to  amalgamate  the  two  taxonomic  systems,  and  which  shows  the  range  of  cyst  morphology  attributable  to  a 
single  living  genus,  the  general  non-conservative  nature  of  certain  cyst  morphologies,  and  the  use  of  common 
cyst  forms  amongst  different  dinoflagellates.  Two  new  names,  Fuscusasphaeridium  and  Asymmetropedinium,  are 
introduced,  and  Quinquecuspis  Harland  is  emended  and  diagnosed  in  Latin. 

Recently  a number  of  important  reviews  have  appeared  describing  the  taxonomy  of  peridiniacean 
dinoflagellates  and  their  cysts.  These  include  Lentin  and  Williams  (1975)  and  Stover  and  Evitt  (1978), 
both  of  which  were  mainly  concerned  with  dinoflagellate  cysts,  their  morphology,  and  particularly 
their  archeopyle  shape  and  structure.  Unfortunately  little  or  no  attention  was  paid  to  the  taxonomy 
of  extant  dinoflagellates  and  their  cysts,  or  to  the  recently  fossilized  cysts  found  in  Quaternary 
sediments.  Similarly,  in  the  field  of  phycology  a major  revision  of  a part  of  the  genus  Peridinium 
Ehrenberg  1832  was  published  by  Balech  (1974),  but  with  little  reference  to  dinoflagellate  cysts.  This 
paper  attempts  to  look  at  the  classification  of  living  peridiniacean  dinoflagellates,  particularly  those 
associated  with  the  genus  Peridinium,  in  relation  to  Recent  and  Quaternary  dinoflagellate  cysts. 

TAXONOMY  OF  THE  GENUS  PERIDINIUM 

The  genus  Peridinium  was  erected  in  1 832  by  Ehrenberg  with  Peridinium  cinctum  (O.  F.  Muller)  as  the 
type  species.  In  the  early  part  of  the  twentieth  century,  with  the  advent  of  major  oceanographical 
expeditions  and  the  increasing  number  of  species  attributed  to  the  genus  (now  well  over  200)  a 
number  of  schemes  to  subdivide  the  genus  were  published.  Jorgensen  (1913)  comprehensively 
reclassified  the  genus,  and  used  the  number  of  plates  that  border  the  first  apical  as  the  principal 
character  in  the  formation  of  subgenera,  and  the  pattern  of  the  dorsal  plates  on  the  epitheca,  in 
particular  the  relationship  with  the  precingular  plate  series,  for  the  division  of  the  subgenera  into 
sections. 

In  the  first  instance  the  genus  was  divided  into  two  subgenera:  Orthoperidinium  [T  contacts  four  plates  1",  2',  4', 
and  7”]  and  Metaperidinium  [1 ' contacts  five  plates  2',  1 ",  2",  4',  and  7"].  The  third  condition  Paraperidinium  [T 
contacts  six  plates  2',  1 ",  2",  4',  6",  and  7"]  was  not  given  subgeneric  status  because  of  the  considerable  variation 
in  the  length  of  the  sutures  between  the  plates.  The  subgenus  Orthoperidinium  was  then  divided  into  the 
following  three  sections: 

Tabulata  2a  plate  contacts  3"  and  4"  or  4"  and  5"  plates 

Conica  2a  plate  contacts  3",  4",  and  5" 

Oceanica  2a  plate  contacts  4"  only, 

Similarly  the  subgenus  Metaperidinium-. 

Pyriformia  2a  plate  contacts  3"  and  4"  or  4"  and  5" 

Paraperidinium  see  discussion  above 

Humilia  2a  plate  contacts  4"  only;  with  solid  antapical  horns 

Divergens  as  above  with  hollow  antapical  horns. 


| Palaeontology,  Vol.  25,  Part  2,  1982,  pp.  369-397,  pis.  38-42.| 


370 


PALAEONTOLOGY,  VOLUME  25 


This  system  was  revised  by  Paulsen  ( 1 93 1 ) to  take  into  account  a greater  number  of  characters,  and  he 
used  the  terms  ‘ortho’-contacts  1",  2',  3',  4',  and  7",  ‘meta’ -contacts  l",  + 2”,  2',  3',  4',  + 6",  and  7", 
and  ‘para’-contacts  1 ",  2",  2',  3',  4',  6",  and  7"  to  characterize  the  first  apical  plate  (see  text-fig.  1);  and 
‘quadra’-contacts  4"  only,  ‘hexa’-contacts  3",  4",  and  5"  and  ‘penta’-contacts  3"  and  4"  to 
characterize  the  dorsal  epithecal  configuration,  especially  the  contacts  at  the  posterior  border  (see 
text-fig.  2).  Paulsen  included  in  his  treatment  of  Peridinium,  forms  with  two  intercalary  plates  as 


text-fig.  1 . Species  of  Protoperidinium  illustrating  the  three  different  types  of 
first  apical  plates  (after  Graham  1942).  The  first  apical  plate  is  stippled.  The 
ortho  species  also  shows  growth  bands. 


text-fig.  2.  Examples  of  the  three  different  styles  of  dorsal  epithecal  tabulation  (after  Graham 
1942).  The  second  dorsal  intercalary  plate  is  hachured. 

opposed  to  the  normal  three  referring  them  to  the  subgenus  Archaeperidinium  ( = genus 
Archaeoperidinium  of  Jorgensen  (1913)).  These  taxonomic  schemes  all  rely  upon  the  ‘stability’  of 
plate  patterns  in  the  region  of  the  first  apical  and  dorsal  epithecal  plates  and  in  general  on  the 
possession  of  three  dorsal  intercalary  plates.  Graham  (1942)  noted,  however,  that  the  number  of 
plates  in  the  cingulum  could  be  important  in  the  taxonomy  of  Peridinium  species  and  that  their  study 
had  been  neglected  despite  the  fact  that  the  dissection  of  the  plates  was  not  difficult. 

Balech  (1974),  in  his  major  revision  of  a part  of  the  genus  Peridinium , used  the  nature  of  the 
cingular  plates  to  differentiate  between  freshwater  Peridinium  sensu  stricto  with  five  cingular  plates, 
and  marine  Peridinium  species  with  four  cingular  plates.  The  generic  name  Protoperidinium  Bergh 
was  used  to  accommodate  these  marine  species.  Protoperidinium  was  then  subdivided  into  three 
subgenera  based  upon  the  number  of  precingular  and  intercalary  plates.  Further  subdivisions  were 
recognized  using  the  criteria  first  employed  by  Jorgensen  (1913)  and  Paulsen  (1931). 

Although  the  basis  of  this  taxonomic  subdivision  is  the  number,  shape,  and  size  of  the  plates 
together  with  the  tabulation  pattern,  I have  never  seen  a reasoned  account  explaining  why  the  number 
of  cingular  plates  and  the  number  of  intercalary  plates  should  be  given  hierarchical  priority  over  the 
configuration  of  the  first  apical  plate,  or  why  the  configuration  of  the  first  apical  plate  should  be  given 


HARLAND:  DINOFLAGELLATE  CYSTS 


37: 


priority  over  the  mutual  relationship  of  the  second  intercalary  plate  and  its  neighbours.  Indeed, 
evidence  from  the  cyst  morphology  suggests  that  the  shape  and  relative  position  of  the  second 
intercalary  plate  is  possibly  more  significant,  because  it  is  through  this  site  that  excystment  occurs. 
Recent  work  by  Gocht  and  Netzel  (1974,  1976)  on  the  overlap  system  in  peridiniacean  and 
gonyaulacacean  dinoflagellates  (Durr  and  Netzel  1974)  has  suggested  that  this  second  intercalary  site 
is  the  keystone  for  both  overlap  or  imbrication  (Dorhofer  and  Davies  1980),  and  in  archeopyle 
formation,  and  that  it  must  be  genetically  determined.  This  is  in  contrast  to  the  first  apical  plate  which 
does  not  appear  to  play  such  a major  role  in  thecal  or  cyst  function. 

It  is,  however,  Balech’s  (1974)  scheme  that  is  used  herein  as  a basis  for  discussing  the  contribution 
of  cyst  morphology.  We  are  obliged  to  use  thecal  morphology  as  the  starting-point  since  the 
information  is  potentially  complete  and  only  a small  proportion  of  the  dinoflagellates  produce 
fossilizable  cysts  as  a part  of  their  life-cycle  (Dale  1976).  Cysts  of  the  genus  Protoperidinium  are 
known  to  have  a varied  morphology  ranging  from  simple  spheres  to  quite  complex  cysts  with 
processes  and  horns.  Often  the  common  unifying  feature  is  the  style  of  archeopyle  formation,  which 
always  involves  the  use  of  an  intercalary  paraplate  with  or  without  its  adjacent  paraplates.  Only  in 
rare  cases  is  it  possible  to  distinguish  a clear  paratabulation.  In  this  study  emphasis  is  necessarily 
placed  upon  gross  morphology  and  archeopyle  formation  in  investigating  the  relationships  of  the 
various  Protoperidinium  cysts  to  their  respective  thecal  stages. 

SYSTEMATIC  DESCRIPTIONS 

In  each  of  the  subsequent  discussions  of  genera,  subgenera,  sections,  and  species,  the  relevant  cyst 
morphologies  will  be  particularly  noted,  with  further  comment  reserved  for  later  sections.  Cysts 
known  to  the  author  are  described  in  some  detail,  otherwise  the  reader  is  referred  to  the  best  published 
description.  The  accompanying  thecal  tabulation  diagrams  have  been  standardized,  do  not 
necessarily  correspond  to  details  of  the  particular  taxa  in  nature,  and  should  therefore  be  treated  as 
diagrammatic. 

All  the  cysts  illustrated  here  are  registered  in  the  MPK  series  and  are  housed  in  the  Palynological 
Collections  of  the  Institute  of  Geological  Sciences  (IGS),  Leeds. 

The  taxonomic  changes  that  have  resulted  from  this  review,  including  the  erection  of  two  new  taxa 
and  the  change  in  status  of  several  others,  are  all  handled  within  the  Appendix.  The  taxonomic  system 
used  herein  is,  however,  summarized  in  Table  1. 

Division  pyrrhophyta  Pascher  1914 
Class  DINOPHYCEAE  Fritsch  1929 
Order  peridiniales  Haeckel  1 894 
Family  peridiniaceae  Ehrenberg  1832 
Genus  protoperidinium  Bergh  emend  Balech  1974 

Type  species.  Protoperidinium  pellucidum  Bergh  1881;  S.D.  by  Loeblich,  Jr.  and  Loeblich,  III,  1966. 

Remarks.  This  genus  accommodates,  for  the  most  part,  the  marine  species  formerly  belonging  to  the 
genus  Peridinium  Ehrenberg.  They  have  four  cingular  plates  in  contrast  to  the  five  in  Peridinium  sensu 
stricto.  An  exception  is  P.  faeroense  Paulsen  which  has  five  cingular  plates  (Dale  1978).  The  type 
species  possesses  a para  first  apical  plate  arrangement  and  a hexa  dorsal  epithecal  configuration  (text- 
fig.  3).  Cysts  capable  of  fossilization  have  not  been  recorded  for  the  type  species. 

Subgenus  Minusculum  (Lebour)  Balech  1974 
Type  species.  Protoperidinium  ( Minusculum ) bipes  (Paulsen)  Balech  1974;  O.D. 

Remarks.  This  subgenus  is  characterized  by  six  precingular  plates  and  three  intercalary  plates.  The 
unique  ‘boomerang’  shape  of  plate  6"  (Balech  1974)  is  also  a significant  feature  (text-fig.  4).  No 
fossilizable  cysts  have  been  observed  from  species  attributable  to  this  subgenus. 


table  1.  Taxonomy  of  Protoperidinium  species  that  produce  fossilizable  cysts 


GENUS 

SUBGENUS 

SECTION 

SPECIES 

Protoperidinium 

Minus  culum 

A rchaeperidinium 

A rchaeperidinium  stat.  nov. 

minuium 

Stelladinium  stat.  nov. 

compressum 

Fuscusasphaeridium  nomen  nov. 

ave liana 

denticulatum 

excentricum 

Protoperidinium 

Para 

Quadra 

Penta 

Hexa 

Protoperidinium  stat.  nov. 

latissimum 

Meta 

Quadra 

Penta 

Hexa 

- 

- 

Ortho 

Quadra 

Votadinium  stat.  nov. 

oblongum 

Penta 

Asymmetropedinium  nomen  nov. 

punctulatum 

Hexa 

Brigantedinium  stat.  nov. 

conicoides 

Selenopemphix  stat.  nov. 

nudum 

subinerme 

Quinquecuspis  stat.  nov. 

leonis 

Trinovantedinium  stat.  nov. 

pentagonum 

table  1.  Taxonomy  of  Protoperidinium  species  that  produce  fossilizable  cysts 


HARLAND:  DINOFLAGELLATE  CYSTS 


373 


text-fig.  3 {left).  Epithecal  tabulation  of  Protoperidinium 
{ Protoperidinium ) pellucidum  Bergh  ex  Loeblich,  Jr.  and 
Loeblich,  III  (after  Lebour  1925). 
text-fig.  4 {right).  Epithecal  tabulation  of  Protoperidinium 
{Minusculum)  bipes  (Paulsen)  Balech  (after  Balech  1974). 


Subgenus  Archaeperidinium  (Jorgensen)  Balech  1974 
Type  species.  Protoperidinium  {Archaeperidinium)  minutum  (Kofoid)  Loeblich,  III,  1969;  O.D. 

Remarks.  Archaeperidinium  has  seven  precingular  plates,  an  ortho  first  apical  and  two  intercalary 
plates,  and  contains  at  least  five  species  that  are  known  to  produce  fossilizable  cysts.  These  are 
described  and  discussed  herein. 

Protoperidinium  {Archaeperidinium)  minutum  (Kofoid)  Loeblich,  III  1969 
Text-fig.  5 

Remarks.  The  thecal  plate  configuration  of  this  the  type  species  shows  the  distinct  ortho  first  apical 
and  two  symmetrically  placed  intercalary  plates.  The  cysts  produced  by  this  species  are  described  in 
Wall  and  Dale  (1968)  and  illustrated  on  their  pi.  4,  fig.  7 but  they  have  not  been  seen  by  the  present 
author.  The  archeopyle,  however,  is  reported  to  be  intercalary  (?  la  or  2a)  but  is  rarely  seen  as  an 
aperture;  Wall  and  Dale’s  (1968)  illustration  appears  to  show  an  attached  operculum.  The  cyst  is 
otherwise  characterized  by  short,  hollow  processes  with  a flat-topped  distal  extremity  that  bears 
spinules  around  the  circumference.  The  cysts  are  known  from  Recent  sediments.  Incubation 
experiments  (Wall  and  Dale  1968)  have  clearly  established  the  link  between  this  theca  and  cyst. 

Protoperidinium  {Archaeperidinium)  avel/ana  (Meunier)  Balech  1974 
Text-fig.  6,  7b;  Plate  38,  figs.  4-9 


text-fig.  5 {left).  Epithecal  tabulation  of  Protoperidinium 
{Archaeperidinium)  minutum  (Kofoid)  Loeblich,  III  (after 
Wall  and  Dale  1968). 

text-fig.  6 {right).  Epithecal  tabulation  of  Proto- 
peridinium {Archaeperidinium)  avellana  (Meunier)  Balech 
(after  Wall  and  Dale  1968). 


374 


PALAEONTOLOGY,  VOLUME  25 


Remarks.  The  thecal  tabulation  illustrated  shows  the  salient  features  but  it  is  interesting  to  note  that 
there  is  some  asymmetry  in  the  position  of  the  intercalary  plates  and  especially  plate  4".  The  cysts  of 
this  species  have  been  described  and  illustrated  by  Wall  and  Dale  (1968,  pi.  4,  fig.  2)  and  consist  of 
spherical,  smooth-walled  brown  bodies  with  a single  dorsal  intercalary  archeopyle.  The  shape  of  the 
operculum  is  seen  in  Wall  and  Dale  (1968),  Reid  (1977),  Harland  (1977)  and  in  the  present 
illustrations.  The  degree  of  reflection  of  the  detailed  thecal  plate  shape  is  often  quite  remarkable. 

There  can  be  some  confusion  between  these  cysts  and  those  of  P.  (A.)  denticulatum  (Gran  and 
Braarud)  Balech  1974  although  the  archeopyles  are  distinct  (text-fig.  7).  In  palynological 
preparations  many  such  cysts  are  often  so  crumpled  that  a proper  analysis  of  the  archeopyle  shape  is 
not  always  possible.  Also,  the  archeopyle  shape  can  be  somewhat  modified  and  need  not  be  a perfect 
reflection  of  the  thecal  plate  counterpart.  Incubation  experiments  (Wall  and  Dale  1968)  have 
established  the  link  between  this  theca  and  this  cyst. 


text-fig.  7.  Apical  view  of  the  cysts  A — Protoperidinium 
( Archaeperidinium ) denticulatum  (Gran  and  Braarud) 
Balech  (after  Wall  and  Dale  1968),  and  B— P.  (A.) 
avellana  (Meunier)  Balech  showing  the  distinct  archeo- 
pyle shapes,  both  symmetrical. 

The  cyst  was  originally  described  as  Chytroeisphaeridia  cariacoensis  Wall  1965,  but  may  be  placed 
in  Reid’s  (1977)  genus  Brigantedinium  recently  validated  by  Harland  and  Reid  (1980)  in  Harland, 
Reid,  Dobell,  and  Norris  (1980).  However,  the  type  species  of  Brigantedinium , B.  simplex , is  the  cyst 
of  P.  ( Protoperidinium ) conicoides  (Paulsen)  Balech,  a member  of  a separate  subgenus.  The  placement 
of  P.  (A.)  avellana  in  Brigantedinium  may,  therefore,  be  inappropriate  and  perhaps  a separate 
designation  needs  consideration  (see  later  discussion). 

Cyst  description.  Simple  spheroidal  brown  cyst  made  up  of  autophragm  or  with  two  very  closely  adpressed  wall 
layers.  Paratabulation  not  present  although  the  archeopyle  shape  does  hint  at  the  configuration  of  the  2a,  4',  3', 
and  la  contacts  (text- fig.  7b).  Cyst  wall  generally  smooth  but  may  appear  somewhat  shagreenate;  the  possible 
indications  of  two  indentations  on  the  ? sulcal  area  may  be  a reflection  of  flagellar  pores.  Archeopyle  intercalary 
formed  by  the  loss  of  paraplate  2a  which  is  transversely  elongate  and  symmetrical  (see  text-fig.  7b).  Operculum 
free.  Cyst  diameters  fall  generally  within  the  range  50-0-55-0  ^m. 


EXPLANATION  OF  PLATE  38 

All  figures  are  illustrated  at  a magnification  of  x 500  and  were  photographed  in  plain  transmitted  light  unless 

otherwise  indicated. 

Figs.  1-3.  Protoperidinium  ( Protoperidinium  sect.  Brigantedinium)  conicoides  (Paulsen)  Balech,  Nomarski 
interference  contrast.  1,  dorsal  view  of  MPK  1232  showing  simple  morphology  and  standard  hexa 
archeopyle.  2,  ventral  view  of  MPK  1232  with  the  two  flagellar  indentations.  3,  dorsal  view  with  the 
archeopyle  and  operculum,  specimen  MPK  1243. 

Figs.  4-9.  Protoperidinium  ( Archaeperidinium  sect.  Fuscusasphaeridium)  avellana  (Meunier)  Balech,  figs.  4 and  5 
by  Nomarski  interference  contrast.  4,  apical  view  showing  symmetrical  archeopyle  with  operculum  in  place, 
specimenMPK  1236.  5,  ditto,  specimen  MPK  1238.  6,  oblique  view  with  operculum  coming  free,  specimen 

MPK  1235.  Figs.  7-9,  various  levels  of  focus  showing  operculum  free  and  more  crumpled  nature  of  the  cyst, 
specimen  MPK  2768. 

Figs.  10-12.  ?Indet.  Protoperidinium  cyst,  various  levels  of  focus  to  show  the  general  morphology  and  the  zigzag 
archeopyle  split,  specimen  MPK  2769. 


PLATE  38 


HARLAND,  Protoperidinium 


376 


PALAEONTOLOGY,  VOLUME  25 


Protoperidinium  ( Archaeperidinium ) compressum  (Abe)  Balech  1974 
Text-figs.  8,  9;  Plate  39,  fig.  12 

Remarks.  The  thecal  tabulation  shows  the  features  of  this  subgenus,  but  in  particular  the  asymmetry 
in  the  disposition  of  the  intercalary  and  4"  plates  is  worth  noting.  Text-fig.  8 shows  the  theca  as 
expanded  whereas  in  life  the  species  is  compressed  dorso-ventrally,  hence  the  name.  The  cyst  of  this 
species  is  particularly  striking  in  its  stellate  morphology,  and  has  been  described  by  Wall  and  Dale 
(1968),  Bradford  (1975),  Reid  (1977),  and  Harland  (1977).  The  archeopyle  is  intercalary  and  formed 
by  the  displacement  of  two  paraplates  which  remain  attached  laterally  (text-fig.  9).  They  are 


text-fig.  8 (left).  Epithecal  tabulation  of  Protoperidinium 
(Archaeperidinium)  compressum  (Abe)  Balech  (after  Wall 
and  Dale  1968). 

text-fig.  9 (right).  Archeopyle  formation  in  Protoperidinium 
(Archaeperidinium)  compressum  (Abe)  Balech,  A — Speci- 
men MPK  1256  (see  PI.  39,  fig.  12),  B— Idealized  scheme. 

symmetrically  placed  and  fold  back  to  open  a large  archeopyle.  It  is  thought  that  both  la  and  2a 
paraplates  are  involved.  They  do  not  appear  to  reflect  the  asymmetry  of  the  theca.  These  cysts  are 
unique  and  to  date  at  least  two  species  (herein  PI.  39,  fig.  12,  and  in  Wall  and  Dale  1968,  pi.  2,  figs. 
15-17)  are  known,  only  one  of  which  can  be  definitely  assigned  to  P.  (A.)  compressum.  Incubation 
experiments  (Wall  and  Dale  1968)  have  established  the  link  between  this  theca  and  the  illustrated  cyst 
described  below. 

Cyst  description.  Stellate,  peridinioid  cysts,  compressed  dorso-ventrally  made  up  of  autophragm.  Wall  smooth, 
and  epitract  smaller  than  hypotract.  Cyst  carries  one  apical,  two  antapical,  and  two  lateral  horns  that  are  long, 
solid,  and  acicular.  Thickening  of  the  horn  bases  is  sometimes  apparent.  Paratabulation  not  present.  Archeopyle 
intercalary  formed  by  the  opening  of  two  intercalary  paraplates  that  remain  attached  laterally.  Cysts  range  in 
length,  excluding  the  processes,  from  26-0-37-0  ^m. 

Protoperidinium  (Archaeperidinium)  denticulatum  (Gran  and  Braarud)  Balech  1974 
Text-figs.  7a,  10 

Remarks.  The  tabulation  of  this  species  shows  a remarkable  similarity  to  that  of  P.  (A.)  avellana  j 
especially  with  respect  to  the  slight  asymmetry  of  the  two  intercalary  plates;  best  seen  in  relation  to  the  ] 

position  of  plate  4".  It  is,  therefore,  perhaps  not  surprising  that  the  cyst  illustrated  and  described  by 
Wall  and  Dale  (1968,  pi.  3 fig.  30)  is  very  similar  to  that  produced  by  P.  (A.)  avellana  (see  Wall  and 
Dale  1968,  p.  277)  although  the  identification  is  admittedly  not  positive.  Archeopyle  formation  is  by  | 
loss  of  a single  transversely  elongate  intercalary  paraplate  whose  configuration  is  symmetrical,  but 
unlike  that  of  P.  (A.)  avellana  (text-fig.  7).  In  palynological  preparations  confusion  between  the  two 
forms  is  common.  The  form  illustrated  here  has  a provisional  assignment  to  this  species.  Incubation  ] 
experiments  by  Wall  and  Dale  (1968)  failed  to  positively  identify  the  theca  therefore  a clearly  ' 

established  link  between  this  theca  and  cyst  is  not  possible. 


HARLAND:  DINOFLAGELL ATE  CYSTS 


377 


Protoperidinium  ( Archaeperidinium ) excentricum  (Paulsen)  Balech  1974 
Text-fig.  1 1 

Remarks.  The  thecal  tabulation  of  this  species  is  very  similar  to  both  those  of  P.  (^4.)  avellana  and  P. 
(A.)  denticulatum.  The  cysts  are  brown  oval  bodies  that  Wall  and  Dale  (1968,  p.  278)  describe  as 
being  flattened  in  a polar  direction.  The  archeopyle  has  not  been  described.  Cysts  attributable  to  this 
species  have  not  been  seen  by  the  present  author.  Although  cysts  enclosed  by  thecae  were  recorded  by 
Wall  and  Dale  (1968)  it  is  not  clear  if  any  thecae  were  obtained  by  incubation  and  positively  identified. 


text-fig.  10  {left).  Epithecal  tabulation  of  Proto- 
peridinium ( Archaeperidinium ) denticulatum  (Gran 
and  Braarud)  Balech  (after  Wall  and  Dale  1968). 
text-fig.  1 1 (right).  Epithecal  tabulation  of  Proto- 
peridinium (Archaeperidinium)  excentricum  (Paulsen) 

Balech  (after  Lebour  1925). 

Discussion  of  the  subgenus  Archaeperidinium 

Dinoflagellate  cysts  of  the  subgenus  Archaeperidinium  fall  into  three  distinct  categories.  The  first  is 
exemplified  by  the  type  species,  a cyst  with  short  processes  and  a single  intercalary  archeopyle;  the 
second  a stellate  cyst  with  a two  paraplate  intercalary  archeopyle;  and  finally  the  third  with  brown 
spherical/spheroidal  cysts  with  single  intercalary  transversely  elongate  archeopyle.  These  differences 
in  gross  cyst  morphology  underline  differences  in  the  tabulation  pattern;  the  type  species  is 
symmetrical;  the  second  is  very  markedly  asymmetrical  and  divided  laterally  into  a clearly  marked 
dorsal  and  ventral  epitheca,  and  finally  those  that  are  only  slightly  asymmetrical  especially  in  regard  to 
the  position  of  plate  4".  I believe  this  is  a sufficiently  natural  genotypic  division  to  warrant  the 
application  of  particular  names.  The  names  chosen  are  derived  from  both  biological  and 
palaeontological  literature  and  include  Stelladinium  Bradford  1975,  a name  coined  to  apply  to  the 
unique  stellate  morphology  of  P.  compressum  cysts.  Herein  I shall  treat  them  as  separate  sections  such 
that: 

1.  Protoperidinium  (Archaeperidinium  sect.  Archaeperidinium  stat.  nov.)  includes  minutum 

2.  Protoperidinium  (Archaeperidinium  sect.  Stelladinium  stat.  nov.)  includes  compressum 

3.  Protoperidinium  (Archaeperidinium  sect.  Fuscusasphaeridium  nomen  nov.)  includes  avellana, 
denticulatum , and  excentricum 

The  only  new  name,  Fuscusasphaeridium,  is  formally  erected  in  the  appendix  together  with  the  change 
of  status  of  the  other  two  taxa.  This  approach  hopefully  embodies  the  general  ideas  expressed  by  Wall 
and  Dale  (1968),  Dale  (1976  and  1978),  and  Reid  and  Harland  (1977).  An  amalgamation  of  systems 
should  be  possible  since  we  are  dealing  with  common  organisms,  albeit  at  different  stages  of  their  life- 
cycle. 

Some  difficulties  are  still  apparent,  for  instance  P.  (P.)  punctulatum  (Paulsen)  also  has  brown 
spherical  cysts  with  transversely  elongate  archeopyles  but  with  a very  different  thecal  tabulation. 
Possible  differences  in  cyst  archeopyle  morphology,  i.e.  ? attached  and  an  asymmetrical  opercula 
(see  later  discussion)  may  serve  to  differentiate  them  from  cysts  of  P.  punctulatum. 


378 


PALAEONTOLOGY,  VOLUME  25 


Subgenus  Protoperidinium  (Bergh)  Balech  1974 

Type  species.  Protoperidinium  ( Protoperidinium ) pellucidum  Bergh  1881,  S.  D.  by  Loeblich,  Jr.  and  Loeblich,  III, 
1966. 

Remarks.  The  thecal  tabulation  of  P.  (P.)  pellucidum  has  already  been  discussed,  see  text-fig.  3.  Balech 
(1974)  does,  however,  subdivide  the  subgenus  into  a number  of  units  based  upon  the  pattern  of  the 
first  apical  plate  and  its  adjacent  plates,  and  then  again  on  the  dorsal  intercalary  configuration.  Since 
the  type  species  has  a para  first  apical  plate  that  group  will  be  discussed  first. 

A.  Para-species 

i.  Quadra.  No  fossilizable  cysts  attributable  to  dinoflagellates  in  this  group  are  known. 

ii.  Penta.  No  fossilizable  cysts  attributable  to  dinoflagellates  in  this  group  are  known. 

iii.  Hexa.  Only  one  species  within  this  unit  is  known  to  produce  fossilizable  cysts  and  that  is 
described  below.  The  type  species,  P.  ( P .)  pellucidum,  also  belongs  in  this  group. 

Protoperidinium  { Protoperidinium ) latissimum  (Kofoid)  Balech  1974 

Remarks.  The  thecal  tabulation  of  this  species  clearly  demonstrates  the  para  first  apical  plate  and  the 
hexa  dorsal  intercalary  arrangement  and  the  symmetry  of  the  pattern.  The  fossilizable  cyst  for  this 
species  has  been  described  by  Wall  and  Dale  (1968)  as  large,  pentagonal,  and  dorso-ventrally 
compressed.  A paracingulum  is  represented  by  broad  weakly  excavated  lateral  lobes.  The  epitract  is 
triangular  in  outline  and  large,  while  the  hypotract  is  much  smaller  and  has  two  small  antapical  horns. 
The  archeopyle  is  intercalary,  has  a hexa  shape  but  is  asymmetrical  {see  Wall  and  Dale  1968,  pi.  2,  fig. 
7),  is  large,  taking  up  most  of  the  dorsal  surface  of  the  cyst.  The  operculum  is  free.  Incubation 
experiments  (Wall  and  Dale  1968)  have  firmly  linked  this  theca  with  this  cyst. 

The  cyst  genus  Leipokatium  Bradford  1975  was  erected  to  accommodate  cysts  very  similar  to  that 
described  above.  Since,  however,  the  type  species  for  the  subgenus  also  is  assigned  to  this  group  it 
might  be  prudent  to  suppress  Leipokatium  in  favour  of  Protoperidinium  {Protoperidinium  sect. 


EXPLANATION  OF  PLATE  39 

All  figures  are  illustrated  at  a magnification  of  x 500  and  were  photographed  in  plain  transmitted  light  unless 

otherwise  indicated. 

Figs.  1-3.  Protoperidinium  {Protoperidinium  sect.  Selenopemphix)  conicum  (Gran)  Balech.  1,  apical  view  to 
show  the  reniform  ambitus  and  offset  standard  hexa  archeopyle,  specimen  MPK  2770.  2,  ditto,  specimen 

MPK  2771.  3,  ditto,  specimen  MPK  2772. 

Figs.  4,  5.  Indet.  ?Protoperidinium  cyst,  Xandarodinium  xanthum  Reid.  4,  ?apical  view  showing  over-all 
morphology,  Nomarski  interference  contrast,  specimen  MPK  1261.  5,  ditto,  specimen  MPK  2773. 

Fig.  6.  Protoperidinium  {Protoperidinium  sect.  Selenopemphix)  subinerme  (Paulsen)  Loeblich,  III,  apical  view 
showing  broad  paracingular  zone,  lack  of  ornament,  and  offset  archeopyle  with  operculum,  specimen  MPK 
1634. 

Figs.  7-11.  Protoperidinium  {Protoperidinium  sect.  Trinovantedinium ) pentagonum  (Gran)  Balech.  7,  dorsal 
view  showing  the  broad  hexa  archeopyle,  hyaline  nature  of  the  cyst,  and  the  paracingulum.  8,  ventral  view 
with  parasulcus  and  planar  paracingulum.  9,  optical  section  illustrating  the  nature  of  the  apical  boss, 
specimen  MPK  1240.  10,  dorsal  view  with  broad  hexa  operculum  and  nature  of  the  parasutural  and 

intratabular  spines.  1 1,  ventral  view  and  deeply  indented  parasulcus,  specimen  MPK  2774,  all  figures  with 
Nomarski  interference  contrast. 

Fig.  12.  Protoperidinium  {Archaeperidinium  sect.  Stelladinium)  compressum  (Abe)  Balech  1974,  dorsal  view 
showing  stellate  morphology  and  the  21  archeopyle  with  the  two  opercular  paraplates  attached  laterally, 
Nomarski  interference  contrast,  specimen  MPK  1256. 


PLATE  39 


HARLAND,  Protoperidinium 


380 


PALAEONTOLOGY,  VOLUME  25 


text-fig.  12  {left).  Epithecal  tabulation  of  Proto- 
peridinium  {Protoperidinium)  latissimum  (Kofoid)  Balech 
(after  Wall  and  Dale  1968). 

text-fig.  13  {right).  Epithecal  tabulation  of  Proto- 
peridinium {Protoperidinium)  claudicans  (Paulsen)  Balech 
(after  Wall  and  Dale  1968). 

Protoperidinium  stat.  nov.).  However,  no  fossilizable  cysts  are  known  from  the  type  species  and 
therefore  the  thecal  and  cyst  morphologies  cannot  be  compared.  In  fact  certain  thecal  tabulation 
details  are  different  between  P.  pellucidum  and  P.  latissimum,  but  it  is  difficult  to  know  if  these  are 
taxonomically  significant  at  this  level  (compare  text-figs.  3 and  12). 

B.  Meta-species 

Fossilizable  cysts  have  not  been  described  from  Modern  species  attributable  to  the  quadra,  penta,  or 
hexa  subdivisions  of  this  unit.  However,  it  is  of  interest  that  a new  species  of  Phthanoperidinium 
described  by  Edwards  and  Bebout  (1981)  has  a meta  configuration. 

C.  Ortho-species 

This  group  contains  the  most  numerous  fossilizable  Protoperidinium  cysts  and  like  those  above  can  be 
subdivided  as  follows: 

i.  Quadra.  This  includes  the  species  P.  {P.)  claudicans  (Paulsen)  Balech  1974  and  P.  {P.) 
oblongum  (Aurivillius)  Balech  1974.  Balech  (1974)  considers  the  latter  to  be  a part  of  P.  {P.) 
oceanicum  Vanhoffen.  Both  species  produce  fossilizable  cysts. 

Protoperidinium  {Protoperidinium)  claudicans  (Paulsen)  Balech  1974 
Text-fig.  13 

Remarks.  The  roughly  symmetrical  thecal  tabulation  clearly  shows  the  ortho  first  apical  plate  and  a 
penta  dorsal  configuration.  This  perhaps  indicates  that  the  dorsal  configuration  is  subject  to  some 
phenotypic  variation  as  this  species  is  usually  placed  in  the  quadra  group  (Balech  1974)  although  it  is 
admitted  that  some  possess  the  penta  pattern.  The  cyst  of  this  species  has  been  described  by  Wall  and 
Dale  (1968)  and  Reid  (1977).  In  dorsal  view  it  is  a chordate  cyst  that  bears  numerous  short-pointed 
spines.  Paratabulation  not  present  but  the  parasulcus  is  deep  and  separates  the  two  broad 
asymmetrical  antapical  lobes.  Paracingulum  not  observed.  Archeopyle  formation  is  reported  to  be  by 
loss  of  the  2a  intercalary  paraplate  which  is  subapical  in  position  and  has  a tendency  to  truncate  the 
apex.  Its  shape  tends  to  be  pentagonal  and  if  indeed  only  the  one  paraplate  is  involved  then  it  is  surely 
an  enlarged  archeopyle.  Incubation  experiments  have  confidently  established  the  theca/cyst 
relationship  in  this  species  (Wall  and  Dale  1968).  These  cysts  have  in  the  past  been  included  in  Reid’s 
(1977)  genus  Votadinium. 

Protoperidinium  {Protoperidinium)  oblongum  (Aurivillius)  Balech  1974 
Text-fig.  14;  Plate  40,  figs.  10-12 

Remarks.  The  thecal  tabulation  as  illustrated  clearly  demonstrates  the  ortho  first  apical  plate  and 
quadra  2a  plate  and  the  symmetrical  nature  of  the  over-all  pattern.  The  cyst  of  this  species  has  been 


HARLAND:  DINOFLAGELL ATE  CYSTS 


38: 


described  by  Wall  and  Dale  (1968),  Reid  (1977),  and  Harland  (1977),  with  the  last  two  authors  refer- 
ring it  to  Reid’s  (1977)  genus  Votadinium.  It  may  be  that  this  species  also  exhibits  variation  in  the 
dorsal  thecal  configuration,  as  in  P.  claudicans,  since  there  appears  to  be  some  variation  in  cyst 
morphology  (see  Wall  and  Dale  1 968,  pi.  1 , figs.  23-29).  Incubation  experiments  have  established  the 
theca/cyst  relationship  within  the  species,  but  some  experiments  have  shown  that  relationships  exist 
with  the  varieties  latidorsale  Dangeard,  inaequale  Dangeard,  and  symmetricum  Dangeard  and  this 
may  in  part  explain  some,  if  not  all,  of  the  cyst  morphological  variations. 


text-fig.  14.  Epithecal 
tabulation  of  Protoperi- 
dinium  ( Protoperidinium ) 
oblongum  (Aurivillius) 

Balech  (after  Wall  and 
Dale  1968). 

Cyst  description.  The  cysts  are  chordate  and  smooth,  formed  of  autophragm.  They  possess  a clearly  defined 
deeply  indented  parasulcus  which  effectively  divides  the  two  rounded  antapical  lobes.  Faint  traces  of  probable 
reflected  flagellar  pores  may  sometimes  be  seen  in  the  sulcus.  Paracingulum  not  seen.  The  archeopyle  is  broad 
and  quite  large  and  appears  to  reach  the  apex  of  the  cyst  on  the  dorsal  surface,  giving  the  cyst  a truncated 
appearance  similar  to  P.  claudicans.  Whether  this  archeopyle  is  an  enlarged  intercalary  involving  the  loss  of 
paraplate  2a  or  whether  it  includes  paraplates  la,  2a,  3a,  and  3'  is  not  known.  Cyst  diameters  range  from  54-0  to 
75-0  /xm. 

Discussion  of  ortho  quadra  Protoperidinium  species 

The  two  species  of  this  subdivision  of  the  subgenus  are  clearly  related  in  terms  of  both  their  thecal 
tabulation  and  cyst  morphology.  The  cysts  have  in  common  their  over-all  chordate  morphology  and 
archeopyle  style  despite  the  uncertainty  as  to  which  paraplates  may  be  involved.  The  use  of  a 
particular  name  to  identify  the  condition  is,  therefore,  clearly  necessary.  In  this  instance  Votadinium  is 
available  for  use  as  a section  of  the  subgenus;  hence  Protoperidinium  ( Protoperidinium  sect. 
Votadinium  stat.  nov.). 

The  variation  in  the  dorsal  tabulation  of  P.  claudicans  and  the  variation  in  the  cysts  of  P.  oblongum 
may  point  to  a deficiency  in  the  taxonomy  of  these  species,  with  perhaps  a lack  of  sufficient  evidence 
on  the  varieties  that  exist  and  their  respective  cysts.  Further  incubation  experiments  are  essential  to 
more  closely  define  the  apparent  variation. 

ii.  Penta.  Only  one  species  Protoperidinium  ( Protoperidinium ) punctulatum  (Paulsen)  Balech 
1974  produces  fossilizable  cysts  and  is  included  here. 

Protoperidinium  ( Protoperidinium ) punctulatum  (Paulsen)  Balech  1974 
Text-figs.  15-17;  Plate  42,  figs.  3-6 

Remarks.  This  dinoflagellate,  although  usually  penta,  can  have  a hexa  dorsal  configuration  as  seen  in 
the  illustrations,  and  can  appear  slightly  asymmetrical.  The  cyst  has  been  described  by  Wall  and  Dale 
(1968)  and  is  spherical,  brown,  with  a large  intercalary  archeopyle  formed  by  the  loss  of  paraplate  2a. 
Operculum  may  be  attached  (Wall  and  Dale,  pi.  2,  fig.  27).  The  aperture  is  transversely  elongate  but  is 
asymmetrical  and  this  may  be  a significant  difference  compared  with  the  cysts  of  P.  (A.)  avellana  and 
P.  ( A .)  denticulatum  (compare  text-figs.  7 and  17).  Confusion  between  all  these  cysts  can  occur  if  well- 
preserved,  three-dimensional  orientated  specimens  are  not  available.  However,  the  nature  of  the 
archeopyle  and  thecal  tabulation  does  serve  to  distinguish  this  dinoflagellate  from  other  species.  Wall 
and  Dale  (1968)  have  established  by  incubation  the  link  between  this  theca  and  cyst. 


382 


PALAEONTOLOGY,  VOLUME  25 


text-fig.  15  (left).  Epithecal  tabulation  of  a hexa  Protoperidinium  (Protoperidinium) 
punctulatum  (Paulsen)  Balech  (after  Wall  and  Dale  1968). 


text-fig.  16  (centre).  Epithecal  tabulation  of  a penta  Protoperidinium  (Protoperidinium) 
punctulatum  (Paulsen)  Balech  (after  Lebour  1925). 
text-fig.  17  (right).  Apical  view  of  the  cyst  of  Protoperidinium  (Protoperidinium)  punctulatum 
(Paulsen)  Balech  showing  the  laterally  elongate,  asymmetrical  archeopyle. 


A name  is  required  to  accommodate  these  dinoflagellates  and  to  differentiate  them  from  section 
Fuscusasphaeridium  which  is  erected  herein  to  hold  the  two  Archaeperidinium  species  discussed 
previously.  The  name  proposed  is  Asymmetropedinium  nom.  nov.  (see  appendix),  hence 
Protoperidinium  (Protoperidinium  sect.  Asymmetropedinium)  punctulatum  for  this  species. 

Cyst  description.  Spheroidal  brown  cyst  made  up  of  autophragm  or  two  closely  adpressed  wall  layers.  Surface 
smooth,  shagreenate,  or  slightly  granulate.  Paratabulation  only  revealed  by  archeopyle  formation.  Archeopyle 
intercalary  formed  by  the  loss  of  paraplate  2a.  Operculum  large,  transversely  elongate  but  asymmetrical.  Cyst 
diameter  ranges  from  40  0 to  60  fim. 

iii.  Hexa.  This  group  contains  the  following  species  that  produce  fossilizable  cysts. 

Protoperidinium  (Protoperidinium)  conicoides  (Paulsen)  Balech  1974 
Text-fig.  18;  Plate  38,  figs.  1-3 

Remarks.  The  epithecal  tabulation  of  this  species  illustrates  the  nature  of  the  ortho  first  apical  plate 
and  the  hexa  dorsal  intercalary  arrangement  together  with  the  symmetry.  The  cyst  of  this  species  has 


EXPLANATION  OF  PLATE  40 

All  figures  are  illustrated  at  a magnification  of  x 500  and  were  photographed  in  plain  transmitted  light  unless 

otherwise  indicated. 

Figs.  1-8.  Protoperidinium  cyst,  Lejeunia  paratenella  Benedek.  1,  low-focus  dorsal  view  with  attenuated  hexa 
archeopyle.  2,  high-focus  dorsal  view  showing  the  paracingulum  and  nature  of  aciculate  antapical 
horns.  3,  ventral  view  with  slightly  displaced  paracingulum,  specimen  MPK  2775.  4,  dorsal  view  with 
archeopyle  and  denticulate  cingular  parasutures.  5,  ditto  in  phase  contrast,  specimen  MPK  2776.  6,  dorsal 

view  showing  attenuated  hexa  archeopyle  with  operculum  in  place,  Nomarski  interference  contrast.  7, 
ventral  view,  Nomarski  interference  contrast,  specimen  MPK  1247.  8,  dorsal  view  with  attenuated  hexa 

archeopyle,  Nomarski  interference  contrast,  specimen  MPK  1245. 

Fig.  9.  Protoperidinium  (Protoperidinium  sect.  Quinquecuspis)  leonis  (Pavillard)  Balech,  dorsal  view  with 
standard  hexa  archeopyle  exhibiting  slight  apical  tongue.  Difference  in  archeopyle  style  between  this  and 
previous  species  is  marked,  specimen  MPK  2777. 

Figs.  10-12.  Protoperidinium  (Protoperidinium  sect.  Votadinium)  oblongum  (Aurivillius)  Balech.  10,  ventral 
view  showing  overall  chordate  morphology.  11,  dorsal  view  with  archeopyle  truncating  apex.  12,  ditto, 
Nomarski  interference  contrast,  specimen  MPK  2778. 


PLATE  40 


HARLAND,  Protoperidinium 


384 


PALAEONTOLOGY,  VOLUME  25 


been  described  by  Wall  and  Dale  (1968),  Reid  (1977),  and  Harland  (1977),  and  was  originally 
placed  in  the  cyst  genus  Chytroeisphaeridia  as  C.  simplicia  by  Wall  (1965).  More  recently  it  has  been 
chosen  as  the  type  species  for  the  cyst  genus  Brigantedinium  Reid  which  has  been  validated  by 
Harland  and  Reid  (1980)  in  Harland  et  al.  (1980).  The  cyst  morphology  is,  like  many  other 
Protoperidinium  cysts,  a simple  brown  ball.  This  cyst  has  not  been  successfully  incubated  but  the 
link  between  theca  and  cyst  is  almost  certain  (Wall  and  Dale  1968). 


text-fig.  18.  (left)  Epithecal  tabulation  of  Proto- 
peridinium (Protoperidinium)  conicoides  (Paulsen) 
Balech  (after  Wall  and  Dale  1968). 
text-fig.  19  (right).  Epithecal  tabulation  of  Proto- 
peridinium (Protoperidinium)  conicum  (Gran)  Balech 
(after  Wall  and  Dale  1968). 


Cyst  description.  A spherical  brown  cyst  made  up  of  autophragm  which  may  be  smooth,  shagreenate,  or  loosely 
reticulate.  May  exhibit  a slightly  indented  paracingular  and  parasulcal  area,  the  latter  possessing  indications  of 
the  flagellar  pores.  Archeopyle  formed  by  loss  of  a single  intercalary  paraplate  (2a)  of  standard  hexa  shape  which 
may  show  interesting  detail  around  the  margin  (PI.  38,  fig.  1).  This  detail,  especially,  toward  the  apex,  perhaps 
suggests  the  loss  of  more  than  just  a single  paraplate.  Operculum  free.  Cyst  diameters  range  from  30-0  to  500  jum. 

This  cyst  can  be  distinguished  from  other  similar  cysts  by  the  nature  of  its  archeopyle  (see  earlier  sections)  but 
it  is  often  difficult  to  recognize  if  the  cysts  are  crushed. 

Protoperidinium  (Protoperidinium)  conicum  (Gran)  Balech  1974 
Text-fig.  19;  Plate  39,  figs.  1-3;  Plate  42,  figs.  1,10 

Remarks.  The  thecal  epitabulation  shows  the  main  features  of  this  group  and  is  distinctly 
symmetrical.  An  interesting  feature  is,  however,  the  almost  rectangular  shape  of  plates  2'  and  4'.  The 
cysts  of  this  species  have  been  described  by  Wall  and  Dale  (1968),  Bradford  (1975),  Reid  (1977),  and 
Harland  (1977).  The  cyst  genus  Multispinula  Bradford  has  been  applied  to  them  although  this  is  a 
junior  synonym  of  Selenopemphix  Benedek  which  was  originally  erected  for  cysts  showing  polar 
compression  and  an  offset  archeopyle  (Stover  and  Evitt  1978).  Bujak  (1980)  formally  emended 
Selenopemphix  to  draw  attention  to  this  offset  archeopyle  and  also  to  include  spinose  forms.  Wall  and 
Dale  (1968)  established  the  theca/cyst  link  by  incubation  studies. 

Cyst  description.  Ovoidal  to  reniform  cysts  probably  made  up  of  two  wall  layers,  the  outer  making  up  the  solid 
aciculate  processes.  Cyst  usually  has  a polar  compression  due  to  its  general  cyst  morphology  of  low  epicystal  and 
hypocystal  cones  (see  Wall  and  Dale  1968,  pi.  2,  figs.  4,  5).  This  compression  may  also  have  the  effect  of  slightly 
rotating  the  epicyst  to  accentuate  the  offset  archeopyle.  PI.  39,  figs.  1 -3  demonstrates  the  variable  amount  of 
offset  of  the  archeopyle.  The  sulcal  area  is  exhibited  as  an  indentation  in  the  ventral  side  of  the  ambitus  and  the 
paracingulum  is  displayed  as  a circumferential  band.  Processes  are  generally  most  common  at  the  apex  and 
circumference  although  they  also  occur  all  over  the  cyst.  Archeopyle  intercalary  formed  by  the  loss  of  paraplate 
2a  which  exhibits  a standard  hexa  shape  and  is  asymmetrically  placed.  Operculum  free.  The  offset  archeopyle 
may  indicate  the  utilization  of  an  additional  paraplate  to  the  2a  in  archeopyle  formation.  Cyst  diameter  40  0 to 
60  0 ^m. 


HARLAND:  DINOFLAGELL ATE  CYSTS 


385 


There  is  some  confusion  between  these  cysts  and  those  of  P.  { P .)  nudum  that  have  a similar  morphology  but  are 
smaller.  Bradford  (1975)  and  Reid  (1977)  argued  that  a continuous  size  gradation  existed  between  the  two  such 
that  only  the  one  cyst  species  can  be  recognized  in  the  absence  of  knowledge  of  the  thecal  morphology. 

Protoperidinium  ( Protoperidinium ) leonis  (Pavillard)  Balech  1974 
Text-fig.  20;  Plate  41,  figs.  1-14;  Plate  42,  figs.  7,  9. 

Remarks.  The  epithecal  tabulation  of  this  species  indicates  a possible  assignment  to  the  penta  group 
and  not  to  the  hexa  as  Balech  (1974)  suggests,  but  Lebour  (1925)  has  figured  a specimen  from 
Plymouth  Sound  with  a clear  hexa  dorsal  arrangement.  Again  some  slight  variation  in  the  dorsal 
tabulation  pattern  is  apparent.  A symmetrical  arrangement  of  plates  on  the  epitheca  is  normal.  The 
cysts  of  this  species  have  been  described  by  Evitt  and  Davidson  (1964),  Wall  and  Dale  (1968), 
Bradford  (1977),  Reid  (1977),  and  Harland  (1977)  using  such  generic  taxa  as  Lejeunia  Gerlach, 
Trinovantedinium  Reid  and  Quinquecuspis  Harland  (see  Harland  1977  for  some  synonymies).  Wall 
and  Dale  (1968)  established  the  theca/cyst  link  by  incubation  but  had  difficulty  in  distinguishing 
P.  leonis  from  such  species  as  P.  marielebourae  (Paulsen)  and  P.  obtusum  (Karsten).  The  range  of  cyst 
morphologies  place  within  P.  (P.)  leonis  reinforces  the  need  for  further  study  as  it  is  perfectly  possible 
that  there  are  a number  of  separate  species  involved,  including  those  represented  in  PI.  40,  fig. 
9— species  1,  PI.  41,  figs.  1,  2,  7,  8— species  2,  PI.  41,  figs.  3-6— species  3,  PI.  41,  figs.  9,  10— species  4 
and  PI.  41,  figs.  11-14 — species  5. 

All  these  forms  can  be  attributed  to  the  genus  Lejeunia  Gerlach  emend  Stover  and  Evitt  1978  now 
Lejeunecysta  Artzner  and  Dorhofer  1978  but  I believe  there  is  a justification  in  keeping  them  within  a 
separate  taxon,  Quinquecuspis.  This  is  based  upon  their  brown  colour,  thick  wall,  deeply  indented 
parasulcus,  often  discontinuous  paracingulum,  and  the  standard  hexa  archeopyle,  and  an  archeopyle 
index  of  c.  0-35-0-45.  By  contrast,  Lejeunia  cysts  are  often  paler  in  colour,  thin  walled,  have  a planar 
continuous  paracingulum  with  an  attenuated  hexa  archeopyle,  and  an  archeopyle  index  of  c.  0-2-0-3. 

Cyst  description.  Peridinioid  acavate  brown  cysts  made  up  of  autophragm  which  is  often  thickened  at  the  apex 
and  antapex.  The  cyst  surface  is  generally  smooth  but  may  be  somewhat  shagreenate.  Epitract  may  be  conical  or 
have  shoulders,  whilst  the  hypotract  carries  two  asymmetrical  horns.  Paratabulation  may  be  represented  by 
distinctly  indented  paracingulum  and  sulcus,  the  former  delimited  by  low  discontinuous  or  continuous  ridges. 
Archeopyle  intercalary  by  loss  of  paraplate  2a.  Archeopyle  shape  can  vary  from  standard  hexa  (PI.  41 , figs.  3, 14) 
to  standard  hexa  with  an  apical  tongue  (PI.  41,  fig.  1).  The  antapical  margin  of  the  archeopyle  is  often  at,  or  very 
close  to,  the  paracingulum.  Operculum  free.  Cyst  diameter  varies  from  60  0 to  75  0 (im. 

Protoperidinium  ( Protoperidinium ) nudum  (Meunier)  Balech  1974 
Text-fig.  21 


text-fig.  20  {left).  Eptithecal  tabulation  of  Proto- 
peridinium {Protoperidinium)  leonis  (Pavillard)  Balech 
(after  Wall  and  Dale  1968). 

text-fig.  21  {right).  Epithecal  tabulation  of  Proto- 
peridinium ( Protoperidinium ) nudum  (Meunier)  Balech 
(after  Wall  and  Dale  1968). 


386 


PALAEONTOLOGY,  VOLUME  25 


Remarks.  The  epithecal  tabulation  pattern  shows  all  the  usual  features  of  this  group;  the  ortho  first 
apical  and  the  hexa  dorsal  intercalary  arrangement.  The  cyst  has  been  described  by  Wall  and  Dale 
( 1 968)  as  being  similar  to  those  of  P.  ( P .)  conicum  but  smaller  (31.0  /x m,  excluding  spines  as  compared 
to  50.0  /xm  for  P.  (P.)  conicum ),  and  possessing  relatively  longer  spines  with  conical  bases.  The 
archeopyle  is  formed  by  the  displacement  of  the  dorsal  intercalary  paraplate  (2a).  The  attached 
operculum  carries  two  spines.  The  confusion  in  distinguishing  this  cyst  from  those  of  conicum  in 
palynological  preparations  has  already  been  noted.  Wall  and  Dale  (1968)  incubated  two  thecae  from 
cysts  but  some  doubt  in  identification  of  the  thecae  is  apparent  from  their  Inudum  designation. 

Protoperidinium  ( Protoperidinium ) pentagonum  (Gran)  Balech  1974 
Text-fig.  22;  Plate  39,  figs.  7-11;  Plate  42,  fig.  8 

Remarks.  This  species  fits  well  into  this  group  by  virtue  of  its  epithecal  tabulation  whereas  its  cyst, 
described  by  Wall  and  Dale  (1968),  Bradford  (1977),  Reid  (1977),  and  Harland  (1977),  is  somewhat 
different  from  those  described  above.  It  has  in  fact  been  described  under  the  names  Lejeunia 
applanata  by  Bradford  (1977)  and  Trinovantedinium  capitatum  by  Reid  (1977)  and  Harland  (1977). 

This  cyst  is  quite  distinct  from  others  described  within  this  group  by  virtue  of  its  hyaline  wall, 
possession  of  short  sutural  and  intratabular  spines,  and  a broad  hexa  archeopyle.  The  cyst  of  P.  (P.) 
pentagonum  was  chosen  by  Reid  (1977)  as  the  type  for  his  genus  Trinovantedinium.  This  genus  was 
used  to  accommodate  other  cysts,  especially  those  having  thick  brown  walls,  a peridinioid  outline, 
and  standard  hexa  archeopyle.  I prefer  to  restrict  the  genus  Trinovantedinium  to  include  the  cyst  of 
P.  ( P .)  pentagonum  and  any  similar  forms  (see  Wall  and  Dale  1968,  pi.  2,  figs.  9-10  and  11-12  and 
Bradford  1977),  but  not  to  include  brown  cysts  with  standard  hexa  archeopyles.  The  cyst 
Sumatradinium  hispidum  (Drugg)  Lentin  and  Williams  could  be  related.  Wall  and  Dale  (1968) 
incubated  a cyst  to  establish  the  cyst/theca  relationship  for  P.  (P.)  pentagonum.  However,  other 
similar  cysts  produced  thecae  with  minor  differences  such  that  a number  of  separate  species  or 
varieties  may  be  involved. 

Cyst  description.  Pentagonal  peridinioid  cyst  made  up  of  autophragm,  epicyst  with  apical  boss,  and  sometimes 
shoulders.  Hypocyst  has  two  slightly  asymmetrical  antapical  horns  or  bulges.  Paratabulation  not  easily 
recognized  except  for  the  planar  non-indented  paracingulum  marked  by  sutural  line  of  processes,  the  deeply 
indented  parasulcus,  and  the  pattern  of  intratabular  processes  defining  paraplate  areas.  Processes  are  short,  rigid 
with  hollow  bulbous  bases,  fine,  aciculate  with  bifid,  capitate,  or  infundibular  solid  tips.  Processes  are  both 
sutural  and  intratabular  but  their  dispositions  are  not  sufficiently  clear  to  reveal  the  paratabulation.  Archeopyle 
intercalary  formed  by  the  loss  of  2a  paraplate,  operculum  free,  archeopyle  is  a broad  hexa  style  (PI.  39,  fig.  10). 
Cyst  maximum  length  ranges  from  65-0  to  70-0  /im. 


EXPLANATION  OF  PLATE  41 

All  figures  are  illustrated  at  a magnification  of  x 500  and  were  photographed  in  plain  transmitted  light  unless 
otherwise  indicated. 

Figs.  1-14.  Protoperidinium  ( Protoperidinium  sect.  Quinquecuspis)  leonis  (Pavillard)  Balech.  1 , dorsal  view  with 
archeopyle  and  operculum  showing  a clear  apical  tongue.  2,  ventral  view  with  deeply  indented  parasulcus 
and  somewhat  discontinuous  paracingulum,  specimen  MPK  1230.  3,  ventral  view  with  standard  hexa 

archeopyle,  apical  margin  by  transparency.  4,  dorsal  view,  specimen  MPK  2779.  5,  ventral  view  with 

deeply  indented  parasulcus  and  discontinuous  paracingulum.  6,  dorsal  view,  specimen  MPK  2780.  7, 

dorsal  view  showing  posterior  archeopyle  margin  almost  affecting  the  paracingulum.  8,  ventral  view, 
specimen  MPK  278 1 . 9,  ventral  view.  1 0,  dorsal  view,  specimen  MPK  2782.  1 1 , ventral  view  with  deep 

parasulcus  and  discontinuous  paracingulum.  12,  dorsal  view,  specimen  MPK  2783.  13,  ventral  view  with 

two  flagellar  parapores.  14,  dorsal  view  showing  standard  hexa  archeopyle  with  no  apical  tongue,  specimen 
MPK  2784. 


PLATE  41 


2 


3 


4 


10 


14 


HARLAND,  Protoperidinium 


388 


PALAEONTOLOGY,  VOLUME  25 


Protoperidinium  ( Protoperidinium ) subinerme  (Paulsen)  Loeblich,  III,  1969 
Text-fig.  23;  Plate  39,  fig.  6 

Remarks.  The  thecal  tabulation  reveals  the  ortho  first  apical  plate  and  the  dorsal  epithecal  style.  The 
cyst  of  this  species  has  been  described  by  Wall  and  Dale  (1968)  as  being  characterized  by  a relatively 
wide  and  deeply  indented  cingulum.  The  cyst  illustrated  here  (PI.  39,  fig.  6)  is  believed  to  be 
attributable  to  P.  (P.)  subinerme.  Wall  and  Dale  (1968)  established  the  theca/cyst  relationship  of  this 
species  although  they  recognized  that  the  resultant  theca  from  incubation  was  more  elongate  in  the 
polar  direction  than  other  thecae  of  this  species. 

Cyst  description.  Ovoidal  to  reniform  cysts,  usually  with  a polar  compression  due  to  the  low  cone  morphology  of 
the  epi-  and  hypocyst.  Marked  paracingular  zone  characterized  by  a broad  circumferential  band.  Cyst  smooth, 
light  brown  in  colour,  and  possessing  an  offset  standard  hexa  intercalary  archeopyle  formed  by  the  displacement 
of  ?2a  paraplate.  Operculum  may  be  attached.  Cyst  diameter  ranges  from  45  0 to  60-0  ^m. 

Discussion  of  the  subgenus  Protoperidinium 

Within  the  para  species  of  this  subgenus  the  cyst  morphology  is  of  a peridinioid  form  with  a reduced 
hypocyst  and  an  intercalary  standard  hexa  archeopyle.  Unfortunately  only  one  cyst  type  is  known 


text-fig.  22  (left).  Epithecal  tabulation  of  Proto- 
peridinium ( Protoperidinium ) pentagonum  (Gran)  Balech 
(after  Wall  and  Dale  1968). 

text-fig.  23  (right).  Epithecal  tabulation  of  Proto- 
peridinium (Protoperidinium)  subinerme  (Paulsen) 
Loeblich,  III  (after  Wall  and  Dale  1968). 


EXPLANATION  OF  PLATE  42 

All  figures,  except  the  stereoscan  photomicrographs,  are  illustrated  at  a magnification  of  x 500  and  were  photo- 
graphed in  Nomarski  interference  contrast. 

Fig.  1.  Protoperidinium  (Protoperidinium  sect.  Selenopemphix)  conicum  (Gran)  Balech,  specimen  MPK  2949.  A 
small  specimen  at  the  lower  end  of  the  size  range  that  could  be  confused  with  P.  (P.  sect.  Selenopemphix)  nudum 
(Meunier)  Balech. 

Fig.  2.  Indet.  Protoperidinium  cyst,  specimen  MPK  2950,  with  a well  developed  zigzag  split  archeopyle. 

Figs.  3-6.  Protoperidinium  (Protoperidinium  sect.  Asymmetropedinium)  punctulatum  (Paulsen)  Balech,  specimen 
MPK  2951,  2952  and  2953)  respectively.  Specimens  showing  the  over-all  cyst  morphology  and  nature  of 
operculum,  particularly  its  asymmetrical  morphology. 

Fig.  7.  Protoperidinium  (Protoperidinium  sect.  Quinquecuspis)  leonis  (Pavillard)  Balech,  specimen  MPK  2954 
showing  cyst  morphology  and  archeopyle,  together  with  a continuous  paracingulum,  x e.  1000. 

Fig.  8.  Protoperidinium  (Protoperidinium  sect.  Trinovantedinium)  pentagonum  (Gran)  Balech,  specimen  MPK 
2956,  showing  cyst  morphology  and  the  broad  hexa  archeopyle,  x c.  1000. 

Fig.  9.  Indet.  Protoperidinium  cyst,  specimen  MPK  2957.  Could  be  part  of  the  cyst  variation  as  currently 
attributed  to  P.  (P.  sect.  Quinquecuspis)  leonis  (Pavillard)  Balech,  x c.  1000. 

Fig.  10.  Protoperidinium  (Protoperidinium  sect.  Selenopemphix)  conicum  (Gran)  Balech,  specimen  MPK  2958, 
showing  over-all  cyst  morphology,  x c.  1000. 


PLATE  42 


HARLAND,  Protoperidinium 


390 


PALAEONTOLOGY,  VOLUME  25 


from  this  group  at  the  present  time.  The  designation  Protoperidinium  ( Protoperidinium  sect. 
Protoperidinium ) may  be  used  for  thecae  with  a para  first  apical  and  hexa  dorsal  intercalary 
arrangement  and  for  peridinioid  cysts  with  reduced  hypocysts.  However,  it  must  be  admitted  that  the 
type  species  does  not  appear  to  have  fossilizable  cysts. 

In  the  ortho  species  we  have  most  of  the  living  cyst  species  including,  in  the  quadra  subsection,  the 
chordate  cysts,  with  or  without  spines  and  having  an  archeopyle  that  truncates  the  apex,  i.e.  the 
species  P.  (P.)  claudicans  and  P.  (P.)  oblongum  respectively.  The  designation  Protoperidinium 
{Protoperidinium  sect.  Votadinium)  can  be  used  for  these  ortho,  quadra  species. 

Amongst  the  ortho  penta  species  only  the  cyst  of  P.  (P.)  punctulatum  is  known,  and  like  so  many 
Protoperidinium  cysts  it  is  a brown  sphere.  It  is,  however,  characterized  by  its  archeopyle  which  is 
transversely  elongate  and  asymmetrical.  The  name  Asymmetropedinium  is  proposed  at  section  level  to 
accommodate  this  species. 

In  the  ortho  hexa  species  a fourfold  division  based  upon  cyst  morphology  is  possible.  The  thecal 
tabulation  does  not  in  itself  support  this  division  but  perhaps  other  factors  need  to  be  considered.  The 
first  division  is  represented  by  P.  (P.)  conicoides  which  has  a spherical  brown  cyst  with  a standard 
hexa  archeopyle;  the  second  by  P.  (P.)  conicum,  P.  (P.)  nudum,  and  P.  (P.)  subinerme  that  have  cysts 
commonly  showing  polar  compression  because  of  the  low  conate  shape  of  the  epi-  and  hypocyst,  and 
an  offset  standard  hexa  archeopyle.  A third  type  represented  by  P.  (P.)  leonis  contains  brown 
peridinioid  cysts  with  standard  hexa  archeopyles  and  finally  the  fourth  is  characterized  by  P.  (P.) 
pentagonum  with  a hyaline  peridinioid  cyst  with  a broad  hexa  archeopyle.  The  names  Brigantedinium 
Reid,  Selenopemphix  Benedek,  Quinquecuspis  Harland,  and  Trinovantedinium  Reid  respectively  are 
used  here  to  designate  these  dinoflagellates  at  section  level  and  are  names  that  were  first  established  in 
the  palaeontological  literature. 


Unattributed  ? Protoperidinium  cysts 

Remarks.  There  are  many  ? Protoperidinium  cysts  described  in  the  literature  that  have  been  observed 
in  palynological  assemblages  of  Recent  sediments,  but  whose  attribution  to  the  thecate  stage  is 
unknown.  A number  of  such  forms  are  illustrated  and  commented  upon  here.  The  first  (PI.  38,  figs. 
10-12,  and  PI.  42,  fig.  2)  is  in  essence  a brown  spherical  cyst  that  opens  by  means  of  a zigzag  split. 
Reid  (1972)  described  such  cysts  from  intertidal  sediments  around  the  British  coast.  It  is  interesting  to 
note  that  the  brown  spherical  morphology  is  again  apparent,  but  that  the  archeopyle  morphology  is 
sufficiently  different  to  differentiate  these  cysts  from  the  other  forms  described  earlier.  It  is,  however, 
quite  possible  that  these  cysts  are  from  glenodiniacean  dinoflagellates  such  as  Diplopsalis  (see  Wall 
and  Dale  1986,  pi.  4,  fig.  20). 

A second  type,  that  has  been  described  under  the  name  of  Xandarodinium  xanthum  Reid  1977,  is 
illustrated  on  PI.  39,  figs.  4-5.  This  cyst  has  a unique  morphology  with  hollow  processes  carrying 
distal  furcate  and  bifid  tips.  Reid  (1977)  describes  these  cysts  as  having  a possible  single  paraplate 
intercalary  archeopyle,  but  an  archeopyle  has  not  been  seen  by  me.  The  morphology  in  regard  to  the 
ovoidal/reniform  ambitus  suggests  a polar  compression  and  hints  at  an  attribution  to  sect. 
Selenopemphix,  but  confirmation  and  details  of  archeopyle  formation  must  await  further  study. 
Artemisiocysta  cladodichotoma  Benedek  may  be  related  to  the  cyst  form  described  here.  This  cyst 
might  be  a representative  of  a gymnodinialean  dinoflagellate  (see  Wall  and  Dale  1968,  pi.  4,  fig.  29). 

The  final  type  (PI.  40,  figs.  1-8)  has  been  referred  to  as  Lejeunia  paratenella  Benedek  1972  by 
Harland  (1977)  and  Trinovantedinium  olivum  Reid  1977  by  Reid  (1977).  It  is  characterized  by  its 
peridinioid  outline,  pale  brown  colour,  thickened  apex  and  antapex,  the  latter  often  developed  into 
acicular  horns,  planar  paracingulum  delimited  by  denticulate  sutures,  and  an  attenuated  hexa 
archeopyle  (Lentin  and  Williams  1975).  This  cyst  differs,  therefore,  from  those  referred  to  as 
Quinquecuspis,  and  I prefer  to  see  it  accommodated  in  Lejeunecysta  Artzner  and  Dorhofer  1978. 
Lentin  and  Williams  (1975)  have  attributed  a standard  hexa  archeopyle  to  Lejeunecysta  (as  Lejeunia), 
but  the  nature  of  the  archeopyle  in  the  type  species  has  not  been  clearly  demonstrated.  The  specimen 
figured  by  Benedek  (1972)  does,  however,  appear  to  have  a standard  hexa  archeopyle  which  may 


HARLAND:  DINOFLAGELLATE  CYSTS 


391 


indicate  that  Lejeunecysta  is  a senior  of  Quinquecuspis  and  that  a new  name  needs  to  be  erected  for 
these  cysts.  Until  the  thecate  form  is  identified  further  speculation  is  unwarranted. 

GENERAL  DISCUSSION 

A taxonomic  system  capable  of  amalgamating  both  cyst  and  thecal  data  must  be  one  of  the  aims  of 
dinoflagellate  research  and  would  inevitably  lead  to  a rationalization  of  much  of  our  fossil  data.  The 
exercise,  outlined  above,  tests  that  possibility  and  shows  this,  or  a similar  system,  is  practical.  It  is 
clear  that  both  the  major  and  minor  divisions  of  the  genus  Protoperidinium  Bergh  are  largely  upheld 
by  differing  cyst  morphologies  (text-fig.  24).  The  genus  is  basically  characterized  by  cysts  of  spherical 
to  peridinioid  shape,  possessing  an  intercalary  archeopyle.  The  subgenus  Archaeperidinium  is 
characterized  by  such  unique  cysts  as  those  of  P.  (A.)  minutum,  P.  (A.)  compressum,  and  simple  brown 
spherical  cysts  distinguished  by  broad  hexa,  transversely  elongate,  symmetrical  archeopyles.  Cyst 
morphology  can  also  be  used  to  further  subdivide  the  subgenus  at  section  level.  These  subdivisions 
need,  however,  further  testing  in  relation  to  total  dinoflagellate  morphology. 

Similarly  the  subdivisions  of  the  subgenus  Protoperidinium  based  upon  thecal  morphology  appear 
to  be  largely  substantiated  by  differences  in  cyst  construction.  The  para  species  may  have  cysts 
with  reduced  hypocysts  and  no  paracingulum  (but  the  evidence  is  based  upon  one  species  only). 
The  different  divisions  of  the  ortho  forms  have  the  following  morphologies;  the  quadra  cysts  are 
■ chordate  in  shape  with  an  archeopyle  that  is  basically  intercalary  but  truncates  the  apex  and  may 
involve  additional  paraplates;  the  penta  cysts  are  brown  and  spherical,  with  a broad  hexa  archeo- 
pyle that  is  asymmetrical;  and  the  hexa  cysts  are  both  spherical  and  peridinioid,  with  standard  or 
broad  hexa  archeopyles.  In  this  last  case  the  cyst  morphologies  may  serve  to  further  subdivide  the 
group. 

Dale  (1978)  has  pointed  out  that  dinoflagellate  cysts  are  often  non-conservative,  with  large 
differences  in  cyst  morphology  being  reflected  by  small  ‘minor’  differences  in  thecae.  This  is  supported 
by  the  present  study;  indeed  in  the  ortho  hexa  cysts  large  differences  in  cyst  morphology  are  seemingly 
not  reflected  in  thecal  morphology  at  all.  In  my  view,  major  and  minor  differences  in  morphology  are 
relative  terms.  What  should  be  remembered  is  that  the  genetic  information  is  common  to  both  stages 
of  the  life  cycle,  and  that  all  morphological  differences  must  be  critically  evaluated.  Dinoflagellate 
taxonomy  should  be  based  upon  the  holomorph  where  possible  or  on  as  much  information  as  is 
available  such  that  a common  taxonomy  can  evolve.  In  this  way  it  is  necessary  to  evaluate  both  the 
cyst-based  taxonomy,  often  ‘overclassified’  and  the  thecate-based  taxonomy,  often  ‘underclassified’ 
(Dale  1978),  in  order  to  arrive  at  the  best  amalgamation.  Whilst  agreeing  with  Dale  (1978)  that  new 
cyst-based  nomenclature  should  not  be  created  to  artificially  maintain  the  cyst-based  taxonomy, 
much  useful  information  can  be  derived  from  good  taxonomic  work  based  upon  cysts,  thecae,  or  on 
both. 

From  a review  of  this  nature  the  major  difference  between  the  thecal-based  taxonomy  and  the  cyst- 
based  system  is  one  of  hierarchy  (see  text-fig.  24).  The  palynologist  would  regard  his  cyst  taxa  as 
separate  genera  whereas  the  phycologist  might  be  prepared  to  accept  them  at  section  level.  Once  this 
hierarchical  difference  is  recognized,  the  conflict  between  the  two  systems  largely  vanishes  and  both 
theca  and  cyst  data  can  point  to  areas  where  further  study  is  needed  to  help  resolve  or  enlarge  upon 
taxonomic  decisions.  In  addition,  it  should  assist  the  palynologist  to  view  his  fossil  taxa  in  a better 
perspective.  It  is  clear,  however,  that  the  species  is  the  common  denominator,  is  constant,  and  should 
be  the  basis  of  all  our  mutual  research. 

This  study  has  also  highlighted  the  ‘variation’  in  thecal  and  cyst  morphologies,  especially  in  the 
ortho  quadra  species,  which  appear  to  have  both  quadra  and  penta  dorsal  configurations  as  in  P.  (P.) 
claudicans,  the  ortho  hexa  species  such  as  P.  (P.)  leonis  with  both  penta  and  hexa  configurations,  and 
the  ortho  penta  species  such  as  P.  (P.)  punctulatum  that  have  both  penta  and  hexa  dorsal  epithecal 
patterns.  In  the  ortho  quadra  group  the  data  also  suggest  variation  in  cyst  morphologies,  from 
chordate  cysts  with  archeopyles  truncating  the  apex  to  cysts  with  a peridinioid  shape  and  standard 
hexa  archeopyles.  Recent  studies  have  tended  to  show  that  the  dorsal  epithecal  and  epicystal  pattern 


392 


PALAEONTOLOGY,  VOLUME  25 


Genus  PROTOPERIDINIUM 


text-fig.  24.  Recent  and  Quaternary  Protoperidinium  cyst  types  and  their  respective  positions  in  the  taxonomic 

subdivision  of  the  genus. 


is  of  considerable  importance  in  dinoflagellate  organization  as  both  the  keystone  plate  and  overlap 
system,  together  with  archeopyle  position,  all  relate  to  that  pattern,  and  suggest  a common  genetic 
control. 

Some  of  the  ‘variation’  discussed  may  in  fact  be  more  artificial  than  real,  because  much  of  it  is 
inextricably  linked  to  taxonomic  difficulties.  The  examples  of  P.  ( P .)  oblongum  and  P.  (P.)  leonis  are 
cases  in  point.  Much  of  these  difficulties  will  need  detailed  research  on  both  thecal  and  cyst 
morphologies,  together  with  population  studies,  before  any  satisfactory  conclusions  can  be  reached. 

There  is  a range  of  cyst  types  associated  with  a single  living  genus,  i.e.  from  simple  brown  spheres  to 
peridinoid  and  quite  complex  stellate  morphologies.  The  cyst  types  are,  however,  dominated  by  two 
cyst  morphologies — the  simple  brown  spheres  and  the  peridinoid  cysts,  the  intercalary  archeopyle  is  a 
common  factor  throughout,  although  its  shape,  style,  and  involvement  of  adjacent  paraplates  is 
different  in  the  separate  groups.  The  occurrence  of  the  simple  brown  spherical  cyst  morphology  in  the 
subgenera  Archaeperidinium  and  Protoperidinium  (and  in  both  ortho  penta  and  ortho  hexa  groups) 
may  be  cited  as  a good  example  of  the  development  of  a common  cyst  form  by  different  dinoflagellates 
(homoeomorphy).  Indeed  this  morphology  is  also  seen  in  cysts  of  unknown  affinity  (see  PI.  38,  figs. 
10-12)  and  in  the  glenodiniacean  dinoflagellates  (Reid  1977).  There  is  difficulty  in  recognizing  species 
among  these  brown  spheres  because  of  a usual  lack  of  well-preserved  and  well-orientated  specimens. 
In  addition,  brown  peridinioid  cysts  also  occur  in  more  than  one  dinoflagellate  group. 


HARLAND:  DINOFLAGELLATE  CYSTS 


393 


It  is  interesting  to  observe  the  lack  of  correspondence  between  some  dorsal  epithecal  tabulations 
and  the  partial  paratabulations  revealed  by  archeopyle  formation.  In  particular,  the  stellate  cyst  of  P. 
(A.)  compressum  appears  to  have  a symmetrically  placed  type  21  archeopyle  but  not  a symmetrical 
dorsal  epithecal  tabulation  pattern.  Similarly  P.  (P.)  conicum,  P.  (P.)  nudum , and  P.  (P.)  subinerme 
have  symmetrically  placed  2a  thecal  plates,  but  offset  type  I archeopyles.  Although  the  over-all 
genetic  control  is  apparent  there  does  not  seem  to  be  a strict  template  mechanism.  Whether  or  not 
additional  paraplates  are  involved  in  these  cases  in  archeopyle  formation,  there  is  an  obvious  need  for 
caution.  The  mechanism  and  control  of  archeopyle  formation  is  a fascinating  question  and  is  clearly 
more  complex  than  at  first  appears. 

Unfortunately  there  are  still  a number  of  cysts,  described  in  the  literature,  of  possible 
Protoperidinium  affinity,  whose  ‘parental’  thecae  are  not  known.  One  of  these  is  the  cyst  referred  to  as 
Lejeunia  paratenella  Benedek  by  Harland  (1977).  It  would  be  of  special  interest  to  know  the  thecal 
affinity  of  L.  paratenella  in  order  to  compare  it  with  other  Protoperidinium  species  known  to  produce 
Quinquecuspis  type  cysts.  This  should  serve  to  help  in  understanding  the  nature  of  Lejeunecystaj 
Quinquecuspis  species  in  the  fossil  record. 

Other  fossil  cysts  may  be  underrepresented  within  this  scheme  because  of  their  susceptibility  to 
oxidation  in  nature,  in  the  palynological  preparation  technique,  and  to  treatment  with  strong  acids 
(Dale  1976;  Reid  1972).  It  is  apparent  from  Lentin  and  Williams  (1975)  that  the  fossil  record  of  the 
genus  Protoperidinium  includes  only  those  cysts  attributable  to  the  subgenus  Protoperidinium, 
including  ortho  forms,  such  as  the  fossil  genera  Rhombodinium  Gocht  and  Wetzeliella  Eisenack, 
which  have  a geological  record  from  the  Palaeocene,  and  ortho  forms  such  as  Alterbia  Lentin  and 
Williams  and  Deflandrea  Eisenack,  with  records  from  at  least  the  Albian.  Other  peridiniacean  cysts 
with  more  complex  compound  archeopyles  are  not  represented  here.  The  cysts  portrayed  within  the 
sections  Selenopemphix  and  Quinquecuspis \Lejeunecysta  have  fossil  records  at  least  as  far  back  as  the 
Late  Eocene  and  Late  Cretaceous  respectively. 

CONCLUSIONS 

This  study  should  have  demonstrated  the  feasibility  of  combining  cyst  and  thecate  morphological 
data  into  a sensible  taxonomic  scheme  which  might  be  of  some  application  to  both  Quaternary  and 
some  Tertiary  peridiniacean  dinoflagellates.  It  is,  of  course,  not  complete,  because  much  more  thecal 
data  are  needed  before  a fully  comprehensive  taxonomy  can  be  attempted.  I hope  it  does  suggest  a 
possible  way  forward  and  one  that  could  be  equally  applied  to  gonyaulacacean,  glenodiniacean, 
ceratiacean,  or  pyrophacacaean  dinoflagellates  where  and  when  sufficient  evidence  is  available. 
Notable  points  of  interest  have  been  the  morphological  range  of  cysts  from  a single  living  genus,  and 
the  hierarchical  difference  between  the  sections  erected  herein  and  what  would  be  undoubtedly 
regarded  as  separate  genera  by  palynologists.  The  species  concept  does,  however,  appear  to  be  a 
mutually  acceptable  base. 

Acknowledgements.  The  author  would  like  to  acknowledge  the  various  discussions  with  Dr  J.  P.  Bujak, 
Geological  Survey  of  Canada,  Mr  B.  Dale,  University  of  Oslo,  Dr  P.  C.  Reid,  Institute  for  Marine 
Environment  Research,  Professor  W.  A.  S.  Sarjeant,  University  of  Saskatchewan,  and  Dr  L.  E.  Stover,  Exxon 
Production  Research,  on  the  nature  and  classification  of  Protoperidinium  cysts.  However,  the  views  expressed 
herein  are  solely  his  responsibility.  I thank  Barrie  Dale,  and  Drs  Evitt  and  Reid  for  their  critical  reading.  This 
paper  was  presented  at  the  V International  Palynological  Conference  in  Cambridge,  June  1980.  Thanks  are  also 
due  to  Mrs  Jane  Sharp  for  her  excellent  preparations  and  single  grain  mounting,  to  Mrs  Sue  Crook  for  her 
draughting  skills,  and  to  Mrs  Margaret  Metcalfe  for  her  typing.  Latin  translation  of  the  various  diagnoses  are 
by  Mr  John  A.  Rooney.  This  paper  is  published  with  permission  from  the  Director,  Institute  of  Geological 
Sciences  (N.E.R.C.). 

REFERENCES 

artzner,  d.  g.  and  dorhofer,  G.  1978.  Taxonomic  note:  Lejeunecysta  nom.  nov.  pro.  Lejeunia  Gerlach  1961 

emend.  Lentin  and  Williams  1976 — dinoflagellate  cyst  genus.  Can.  J.  Bot.  56,  1381-1382. 


394  PALAEONTOLOGY,  VOLUME  25 

balech,  E.  1974.  El  genero  ‘Protoperidinium’  Bergh,  1881  (‘Peridinium’  Ehrenberg,  1831,  partim.).  Rev.  Mus. 
Cienc.  natur.,  Hidrobiol.  4,  1-79. 

benedek,  p.  n.  1972.  Phytoplanktonten  aus  dem  Mittel-  und  Oberologozan  von  Tonisberg  (Niederrheingebeit). 
Palaeontographica,  Abt.  B,  137,  1-71. 

Bradford,  m.  r.  1975.  New  dinoflagellate  cyst  genera  from  the  recent  sediments  of  the  Persian  Gulf.  Can.  J.  Bot. 
53,  3064-3074. 

— 1977.  New  species  attributable  to  the  dinoflagellate  cyst  genus  Lejeunia  Gerlach,  1961  emend.  Lentin  and 
Williams  1975.  Grana,  16,  45-59. 

bujak,  j.  p.  1980.  Dinoflagellate  cysts  and  acritarchs  from  the  Eocene  Barton  Beds  of  southern  England.  In 
bujak,  j.  p.,  downie,  c.,  Eaton,  G.  l.  and  williams,  G.  l.  Dinoflagellate  cysts  and  acritarchs  from  the  Eocene  of 
southern  England.  Palaeontology , Spec.  Pap.,  24,  36-91. 
dale,  B.  1976.  Cyst  formation,  sedimentation,  and  preservation:  factors  affecting  dinoflagellate  assemblages  in 
Recent  sediments  from  Trondheimsfjord,  Norway.  Rev.  Palaeobot.  Palynol.  22,  39-60. 

— 1978.  Acritarchous  cysts  of  Peridinium  faeroense  Paulsen:  implications  for  dinoflagellate  systematics. 
Palynology,  2,  187-193. 

dorhofer,  g.  and  davies,  E.  h.  1980.  Evolution  of  archeopyle  and  tabulation  in  Rhaetogonyaulacinean 
dinoflagellate  cysts.  Life  Sciences  Miscell.  Publ.,  Royal  Ontario  Museum,  1-91. 
durr,  G.  and  netzel,  h.  1974.  The  fine  structure  of  the  cell  surface  in  Gonyaulax  polyedra  (Dinoflagellata).  Cell 
Tiss.  Res.  150,  21-41. 

edwards,  L.  E.  and  bebout,  j.  w.  1981.  Emendation  of  Phthanoperidinium  Drugg  and  Loeblich  1967,  and  a 
description  of  P.  brooksii  sp.  nov.  from  the  Eocene  of  the  mid-Atlantic  outer  continental  shelf.  Palynology, 
5,  29-41. 

enrenberg,  c.  G.  1832  [separate  1830].  Beitrage  zur  Kenntniss  der  Infusorien  und  ihrer  geographischen 
Verbrietung  besonders  in  Sibirien.  Abh.  preuss.  Akad.  Wiss.  1830,  1-88. 
evitt,  w.  r.  and  davidson,  s.  e.  1964.  Dinoflagellate  Studies  I.  Dinoflagellate  cysts  and  thecae.  Stanford  Univ. 
Pubis.,  Geol.  Sci.  10  (1),  1-12. 

gocht,  h.  and  netzel,  h.  1974.  Rasterelektronenmikroskipische  Untersuchungen  am  Panzer  von  Peridinium 
(Dinoflagellata).  Arch.  Protistenk.  116,  381-410. 

— 1976.  Reliefstrukturen  des  Kreide— Dinoflagellaten  Palaeoperidinium  pyrophorum  (Ehr.)  in  Ver- 
gleich  mit  Panzer-Merkmalen  rezenter  Peridinium- Arten.  Neues  Jb.  Geol.  Palaont.,  Abh.  153,  380- 
413. 

graham,  H.  w.  1942.  Studies  in  the  morphology,  taxonomy,  and  ecology  of  the  Peridiniales.  Carnegie  Institution 
of  Washington,  542,  1-129. 

harland,  r.  1977.  Recent  and  Late  Quaternary  (Flandrian  and  Devensian)  dinoflagellate  cysts  from  marine 
continental  shelf  sediments  around  the  British  Isles.  Palaeontographica,  Abt.  B,  164,  87-127. 

— reid,  p.  c.,  dobell,  p.  and  norris,  G.  1980.  Recent  and  sub-Recent  dinoflagellate  cysts  from  the  Beaufort 
Sea,  Canadian  Arctic.  Grana,  19,  211-225. 

jorgensen,  e.  (1912)  1913.  Bericht  fiber  die  von  der  schwedischen  Hydrographisch — Biologischen  Kommission 
in  den  schwedischen  Gewassern  in  den  Jahren  1909-1910  eingesammelten  Planktonproben.  Svenska 
hydrogr. — biol.  komm.  Skr.  4,  1-20. 

lebour,  m.  v.  1925.  The  dinoflagellates  of  northern  seas.  Mar.  Biol.  Assoc.  U.K.,  Plymouth,  1-250. 
lentin,  J.  k.  and  williams,  G.  l.  1975.  A monograph  of  fossil  peridinioid  dinoflagellate  cysts.  Bedford  Inst,  of 
Oceanogr.  Rept.  Bl-R-75-16,  1-237. 

paulsen,  o.  1931.  Etudes  sur  le  microplankton  de  la  Mer  d’Alboran.  Trab.  Inst,  esp  Oceanogr.  4,  5- 
108. 

reid,  p.  c.  1972.  The  distribution  of  dinoflagellate  cysts,  pollen  and  spores  in  Recent  marine  sediments  from  the 
coast  of  the  British  Isles.  Unpubl.  Ph.D.  thesis,  Univ.  of  Sheffield,  1-273. 

— 1977.  Peridiniacean  and  glenodiniacean  dinoflagellate  cysts  from  the  British  Isles.  Nova.  Hedwigia,  24, 
429-463. 

— and  harland,  R.  1977.  Studies  of  Quaternary  dinoflagellate  cysts  from  the  North  Atlantic.  Am.  Ass. 
stratigr.  Palynol.,  Contr.  Ser.  5A,  147-169. 

stafleu,  F.  A.  et  al.  1978.  International  Code  of  Botanical  Nomenclature.  Bohn,  Scheltema  & Holkema,  Utricht, 
1-457. 

stover,  l.  E.  and  evitt,  w.  r.  1978.  Analyses  of  pre-Pleistocene  organic-walled  dinoflagellates.  Stanford  Univ. 
Pubis.  Geol  Sci.  15,  1-298. 

wall,  D.  1965.  Modern  hystrichospheres  and  dinoflagellate  cysts  from  the  Woods  Hole  region.  Grana  palynol.  6, 
297-314. 


HARLAND:  DINOFLAGELLATE  CYSTS 


395 


wall  and  dale,  1968.  Modern  dinoflagellate  cysts  and  evolution  of  the  Peridiniales.  Micropaleontology,  14, 
265-304. 


Typescript  received  12  January  1981 
Revised  typescript  received  29  March  1981 


REX  HARLAND 


Institute  of  Geological  Sciences 
Ring  Road  Halton 
Leeds  LSI 5 8TQ 


APPENDIX 

The  appendix  is  used  here  to  indicate  new  names,  and  names  whose  status  has  been  altered.  The  order 
adopted  follows  that  presented  in  the  main  text. 

Subgenus  Archaeperidinium  (Jorgensen)  Balech  1974 
Section  Archaeperidinium  Jorgensen  stat.  nov. 

Type  species.  Protoperidinium  ( Archaeperidinium  sect.  Archaeperidinium)  minutum  (Kofoid)  Loeblich,  III,  1969 

Remarks.  This  name  has  been  used  at  generic  level  by  Jorgensen  ( 1 9 1 3),  at  sub-generic  level  by  Balech 
(1974),  and  here  is  used  at  section  level  as  one  of  the  sections  of  the  subgenus  Archaeperidinium.  It  is 
characterized  by  cysts  with  unique  process  structure  and  a symmetrical  epithecal  tabulation  pattern. 

Section  Stelladinium  Bradford  1975  ex  Harland  and  Reid  1980  stat.  nov. 

Type  species.  Stelladinium  reidii  Bradford  197 5 = Protoperidinium  ( Archaeperidinium  sect.  Stelladinium)  cf. 
compressum  (Abe)  Balech  1974. 

Remarks.  This  name,  originally  erected  as  a genus,  is  altered  in  status  to  become  a section  of  the 
subgenus  Archaeperidinium.  The  type  species  reidii  is  based  upon  a cyst  holotype  and  not  a thecate 
form  but  the  parent  must  be  a species  similar  to  P.  compressum.  It  is  characterized  by  stellate  cysts  and 
an  asymmetrical  thecal  tabulation. 

Section  Fuscusasphaeridium  nom.  nov. 

Derivation  of  name.  Latin:  fuscus,  brown,  sphaera,  ball. 

Diagnosis.  Spherical  to  spheroidal  brown  cyst  of  autophragm  with  a laterally  elongate,  symmetrical, 
archeopyle  formed  by  the  loss  of  a single  intercalary  paraplate  (2a).  Margin  of  archeopyle  often 
shows  its  configuration  with  paraplates  3',  4',  la,  and  6".  Operculum  free. 

Cista  autophragmatis  aurea  quae  forman  habet  sphericalem  vel  spheroidalem  et  latus  elongatum, 
symmetricalis  est.  Archaeopyla  facta  est  uno  paraplato  intercalari  amisso  (?  2a).  Margo  archaeopylae  saepe 
configurationem  suam  ostendit  cum  paraplatis  3',  4'  la  et  6".  Operculem  liberum  est. 

Type  species.  Protoperidinium  ( Archaeperidinium  sect.  Fuscusasphaeridium)  avellana  (Meunier)  Balech 
197 4 = Protoperidinium  avellana  Meunier  1919,  pp.  56,  pi.  18,  figs.  37-41.  Recent. 

Remarks.  The  new  section  erected  here  accommodates  those  dinoflagellates  possessing  brown 
spherical  cysts  with  symmetrical,  transversely  elongate  archeopyles  and  with  slightly  asymmetric 
dorsal  epithecal  tabulations. 

Subgenus  Protoperidinium  (Bergh)  Balech  1974 
Section  Protoperidinium  Bergh  1881  stat.  nov. 

Type  species.  Protoperidinium  ( Protoperidinium  sect.  Protoperidinium)  pellucidum  Bergh  1881. 


396 


PALAEONTOLOGY,  VOLUME  25 


Remarks.  Protoperidinium  is  used  here  at  section  level  as  a part  of  the  subgenus  Protoperidinium.  It 
may  be  characterized  by  cysts  with  inflated  epicysts  but  this  is  as  yet  unproven  since  the  type  species 
does  not  appear  to  have  fossilizable  cysts. 

Section  Votadinium  Reid  1977  stat.  nov. 

Type  species.  Votadinium  calvum  Reid  1977  = Protoperidinium  ( Protoperidinium  sect.  Votadinium ) oblongum 
(Aurivillius)  Balech  1974. 

Remarks.  This  name  is  given  the  new  status  of  section  within  the  subgenus  Protoperidinium,  and  is 
characterized  by  chordate  cysts  with  large  dorsal  intercalary  archeopyles  that  truncate  the  apex 
together  with  ortho  quadra  epithecal  tabulation  patterns. 

Section  Asymmetropedinium  nov. 

Derivation  of  name.  Greek:  asymmetro,  without  symmetry;  ope,  opening;  and  dinium  of  dinoflagellate  affinity, 
with  reference  to  the  asymmetrical  archeopyle. 

Diagnosis.  Spherical  to  spheroidal,  brown  cyst  of  autophragm  that  possesses  a laterally  elongate 
archeopyle  which  is  asymmetrical  about  the  dorso-ventral  plane,  formed  by  the  loss  of  a single 
intercalary  paraplate  (?2a).  Margin  of  archeopyle  may  show  its  configuration  with  adjacent  para- 
plates.  Operculum  ?attached. 

Cista  aurea  autophragmatis  quae  habet  et  formam  sphericalem  vel  spheroidalem  et  latus  elongatum 
asymmetricalis  est.  Archaeopyla  facta  est  uno  paraplato  amisso  (?2a).  Fieri  potest  ut  margo  archaeopylae 
ostendat  configurationem  cum  paraplatis  quae  proximae  sunt.  Operculum  ? adiunctum. 

Type  species.  Protoperidinium  ( Protoperidinium  sect.  Asymmetropedinium)  punctulatum  (Paulsen)  Balech 
\91A  = Peridinium  punctulatum  Paulsen  1907,  p.  19,  fig.  28.  Recent. 

Remarks.  Asymmetropedinium  is  used  here  for  ortho  penta  dinoflagellates  that  have  brown  spherical 
cysts  with  transversely  elongate  archeopyles  that  are  asymmetrical. 

Section  Brigantedinium  Reid  1977  ex  Harland  and  Reid  1980  stat.  nov. 

Type  species.  Brigantedinium  simplex  (Wall)  Reid  1977  = Protoperidinium  ( Protoperidinium  sect.  Brigantedinium ) 
conicoides  (Paulsen)  Balech  1974. 

Remarks.  This  section  is  characterized  by  an  ortho  hexa  epithecal  configuration  and  brown  spherical 
cysts  with  a standard  hexa  archeopyle.  Brigantedinium  was  originally  coined  at  generic  level  by  Reid 
(1977)  to  accommodate  all  brown  spherical  cysts. 

Section  Selenopemphix  Benedek  1972  stat.  nov. 

Type  species.  Selenopemphix  nephroides  Benedek  1972  = ? Protoperidinium  ( Protoperidinium  sect. 
Selenopemphix)  subinerme  (Paulsen)  Loeblich,  III,  1969. 

Remarks.  Selenopemphix,  originally  erected  as  a genus  by  Benedek  (1972),  is  here  altered  in  status  to  a 
section  of  the  subgenus  Protoperidinium  to  hold  ortho  hexa  dinoflagellates  whose  cysts  show  polar 
compression  and  offset  archeopyles.  The  type  species,  the  fossil  S.  nephroides  Benedek,  is  very  similar, 
if  not  identical  to  the  cyst  of  P.  ( P .)  subinerme  (Paulsen). 

Section  Quinquecuspis  Harland  ex  et  emend.  Harland  stat.  nov. 

Type  species.  Quinquecuspis  concretum  (Reid)  Harland  1977  =1  Protoperidinium  ( Protoperidinium  sect. 
Quinquecuspis)  leonis  (Pavillard)  Balech  1974. 

Emended  diagnosis.  Pentagonal,  peridinioid,  acavate,  brown  cyst,  made  up  of  autophragm  which 
thickens  towards  the  apex  and  antapex.  Epitract  conical  or  with  shoulders,  hypotract  with  well  or 


HARLAND:  DINOFL AGELL ATE  CYSTS 


397 


poorly  developed  asymmetrical  horns.  Paratabulation  absent,  paracingulum  conspicuous,  planar, 
and  delimited  by  low  continuous  or  broken  ridges.  Deeply  indented  parasulcus  sometimes  with 
flagellar  scars.  Archeopyle  intercalary,  standard  hexa,  formed  by  the  loss  of  paraplate  2a,  with  or 
without  an  apical  tongue.  Operculum  free. 

Cista  aurea  pentagonalis  peridinoidalis  et  acavata,  ex  autophragmate  facta  quod  densum  fit  versus  apicem  et 
antapicem.  Epitractum  habet  vel  formam  conicalem  vel  humeros,  hypotractum  habet  cornua  asymmetricalia 
quae  bene  vel  male  formata  sunt.  Paratabulatio  nulla  est  sed  paracingulum  et  videri  potest  et  planare  finitum  est 
rugis  quae  vel  imae  vel  fractae  sunt.  Parasulcus  alte  dentatum  cum  cicatricibus  flagellaribus.  Archaeopyla 
intercalaris,  hexa  normalis  quae  facta  est  amisso  paraplato  2a,  cum  vel  sine  lingua  in  apice.  Operculum  liberum. 

Remarks.  This  name  was  originally  erected  at  generic  level  (Harland  1977)  but  did  not  comply  with 
Article  36  of  the  International  Code  of  Botanical  Nomenclature  (Stafleu  et  al.  1978)  which  requires 
that  Recent  algal  taxa  be  furnished  with  a Latin  diagnosis.  The  type  species  had  a Latin  diagnosis  in 
its  original  publication  (Reid  1977)  and  was  valid  from  that  date.  As  a section  of  the  subgenus 
Protoperidinium  it  is  characterized  by  an  ortho  hexa  epithecal  tabulation  and  peridinioid  brown  cysts 
with  a standard  or  modified  standard  hexa  archeopyle. 

These  cysts  are  noted  in  particular  for  their  thick  brown  walls,  and  archeopyles  often  having  an 
apical  tongue  and  deeply  indented  sulcal  areas.  During  archeopyle  formation  a break  often  occurs 
down  across  the  paracingulum  (see  PI.  41,  figs.  4,  7)  from  the  posterior  archeopyle  margin. 

Section  Trinovantedinium  Reid  1977  stat.  nov. 

Type  species.  Trinovantedinium  capitatum  Reid  1977  = Protoperidinium  ( Protoperidinium  sect.  Trinovantedinium) 
pentagonum  (Gran)  Balech  1974. 

Remarks.  Trinovantedinium , a genus  erected  by  Reid  (1977),  is  herein  given  the  new  status  of  section 
within  the  subgenus  Protoperidinium.  It  accommodates  dinoflagellates  with  ortho  hexa  epithecal 
tabulations  and  hyaline  peridinioid  cysts  with  short  spines  and  a broad  hexa  archeopyle. 


A NEW  GENUS  OF  SHARK  FROM  THE 
MIDDLE  TRIASSIC  OF  MONTE  SAN  GIORGIO, 
SWITZERLAND 

by  O.  RIEPPEL 


Abstract.  Associated  material  from  the  Middle  Triassic  of  Monte  San  Giorgio,  Kt.  Tessin,  Switzerland, 
demonstrates  that  the  sharks  Nemacanthus  tuberculatus  and  Acrodus  bicarinatus  constitute  a single  taxon  that 
must  be  included  in  a new  genus,  Acronemus  tuberculatus.  Finspine  structure  of  Acronemus  is  distinctly 
different  from  that  of  Nemacanthus  monilifer,  but  does  correspond  to  the  general  ctenacanthiform  pattern. 
Other  features  of  A.  tuberculatus  such  as  tooth  structure  and  placoid  scales  have  previously  been  reported 
for  Triassic  hybodontiform  sharks  only.  A discussion  of  the  orders  Ctenacanthiformes  and  Hybodontiformes 
concludes  the  study. 

The  Grenzbitumenzone  of  Monte  San  Giorgio,  Kt.  Tessin,  Switzerland,  has  so  far  yielded  four 
hybodontiform  shark  genera,  viz.  Hybodus,  Acrodus , Aster  acanthus,  and  Palaeobates.  A fifth 
selachian  taxon,  the  most  frequently  found  one  in  the  deposits  at  Monte  San  Giorgio,  will  be 
described  in  the  present  contribution.  This  shark  is  important  as  it  highlights  some  problems  in 
the  distinction  of  the  Mesozoic  orders  Hybodontiformes  and  Ctenacanthiformes  as  defined  by 
Maisey  (1975). 

Kuhn  (1945)  first  described  the  genus  Acrodus  from  the  Middle  Triassic  of  Monte  San  Giorgio. 
Kuhn’s  (1945)  specimen  d was  not  fully  prepared  at  the  time.  The  radiograph  showed,  however, 
that  the  proportions  of  the  finspines  of  this  shark  were  very  different  from  those  of  Acrodus,  and 
that  the  palatoquadrate  was  cleaver-shaped  (Kuhn,  1945,  fig.  4),  a feature  which  is  otherwise  not 
known  in  the  genus  Acrodus.  Peyer  (1957)  mentioned  an  Acrodus  from  Monte  San  Giorgio  with  a 
tuberculate  finspine  ornamentation.  Acrodus  finspines  always  show  a costate  ornament. 

The  misidentifications  by  Kuhn  (1945)  and  Peyer  (1957)  are  based  on  the  fact  that  in  the  specimens 
mentioned  by  these  authors  typical  Acrodus  teeth  are  associated  with  stout  and  tuberculate  finspines. 
The  teeth  and  the  finspines  have  been  named  and  referred  to  different  genera  by  earlier  authors. 
Bellotti,  in  an  unpublished  manuscript  on  the  fossil  fishes  at  the  Museo  Civico  di  Milano  (1873), 
had  named  the  teeth  Acrodus  bicarinatus  and  the  spines  Nemacanthus  tuberculatus.  The  first  author 
to  use  these  names  in  a formal  publication  was  Bassani  (1886).  He  used  the  names  in  connection 
with  a valid  diagnosis,  referring  to  Bellotti’s  specimens  and  manuscript.  According  to  the  kind 
information  provided  by  Professor  G.  Pinna,  Museo  Civico  di  Storia  Naturale  Milano,  the  entire 
old  collection  on  which  Bellotti’s  research  was  based  has  been  destroyed  during  the  Second  World 
War.  No  type  material  has  been  preserved.  Likewise,  Bellotti’s  (1873)  manuscript  can  no  longer 
be  located.  The  only  material  still  available  is  unpublished  drawings  by  Bellotti  including  those  of 
N.  tuberculatus. 

The  associated  material  from  Monte  San  Giorgio  demonstrates  that  A.  bicarinatus  and  N. 
tuberculatus  are  a single  taxon.  Since  Bassani  (1886)  is  the  formal  author  of  both  names,  and  since 
in  his  publication  N.  tuberculatus  has  page  priority,  the  correct  species  name  for  this  shark  must 
be  tuberculatus.  The  description  of  the  Monte  San  Giorgio  material  will  make  it  clear  that  this 
shark  species  cannot  be  included  in  the  genus  Nemacanthus,  nor  in  the  genus  Acrodus.  A new  genus 
must  consequently  be  erected.  Since  no  original  type  material  is  preserved,  it  is  appropriate  to 
select  a neotype  for  that  species. 


IPalaeontology,  Vol.  25,  Part  2,  1982,  pp.  399-412,  pi.  43.| 


400 


PALAEONTOLOGY,  VOLUME  25 


SYSTEMATIC  PALAEONTOLOGY 

Class  CHONDRICHTHYES 
Subclass  ELASMOBRANCHII 
Order  ctenacanthiformes  incertae  familiae 
Genus  acronemus  n.gen. 

Type  and  only  known  species.  Acronemus  tuberculatus  (Bassani  1886). 

Revised  diagnosis.  Small  (30-35  cm  long)  ctenacanthiform  shark;  teeth  Acrodus-Wke,  with  a single 
and  blunt  main  cusp,  crown  ornamented  with  prominent  longitudinal  and  transverse  carinae,  root 
without  lingual  torus;  palatoquadrate  cleaver-shaped;  finspines  short  and  stout,  heavily  tuberculated 
with  a longitudinal  ridge  along  the  leading  edge  of  the  crown,  not  or  only  slightly  recurved, 
posterior  wall  concave  without  denticulate  ornamentation,  central  cavity  displaced  posteriorly; 
placoid  scales  with  a lanceolate  crown  ornamented  with  three  or  five  longitudinal  striae. 

Acronemus  tuberculatus  (Bassani  1886) 

Selected  synonymy 

1886  Nemacanthus  tuberculatus,  Bassani,  p.  30. 

1886  Acrodus  bicarinatus,  Bassani,  p.  31. 

1891  Nemacanthus  tuberculatus.  Woodward,  p.  117. 

1910  Acrodus  bicarenatus,  Alessandri,  p.  34,  pi.  7,  figs.  6-9. 

1910  Nemacanthus  tuber colatus,  Alessandri,  p.  36,  pi.  7,  fig.  10. 

Neotype.  Palaontologisches  Institut  und  Museum  der  Universitat  Zurich  T 1548,  Monte  San  Giorgio,  point 
902,  layer  118,  collected  20.9.1957. 

Diagnosis.  Same  as  for  genus. 

Referred  material.  All  specimens  with  T-numbers  belong  to  the  Tessin  collection,  Palaontologisches  Institut 
und  Museum  der  Universitat  Zurich.  T 1177,  dentition  and  two  finspines;  T 1178,  dentition,  two  finspines, 
skull  fragments;  T 1 181,  isolated  tooth;  T 1289,  dentition,  anterior  finspine;  T 1291,  dentition,  anterior  finspine; 
T 1292,  isolated  tooth;  T 1426,  isolated  tooth;  T 1427,  isolated  tooth;  T 1431,  sectioned  finspine;  T 1434, 
isolated  tooth;  T 1448,  dentition;  T 1457,  teeth,  sectioned  finspine;  T 1487,  dentition;  T 1531,  dentition  and 
two  finspines;  T 1551,  incomplete  finspine;  T 2465,  dentition,  upper  and  lower  jaws,  two  finspines  (see  Kuhn 
1945,  fig.  4);  T 3297,  isolated  tooth;  T 3812,  dentition;  T 3818,  two  finspines;  T 3819,  dentition;  T 3820, 
dentition;  T 3821,  dentition  and  jaw  fragments;  T 3822,  dentition  and  anterior  finspine;  T 3827,  dentition; 
T 3825,  dentition  and  two  finspines;  T 3826,  dentition;  T 3828,  posterior  finspine;  T 3829,  dentition  and  anterior 
finspine;  T 3831,  dentition;  T 3833,  dentition;  T 3834,  dentition;  T 3835,  dentition  and  finspine;  T 3836, 
dentition;  T 3837,  dentition;  T 3840,  sectioned  finspine;  T 3841,  anterior  finspine;  T 3843,  sectioned  finspine; 
T 3844,  finspine;  T 3845,  finspine;  T 3846,  sectioned  finspine;  T 3847,  sectioned  finspine;  T 3849,  fifty-two 
isolated  teeth;  British  Museum  (Natural  History)  P 19450,  complete  dentition. 

Distribution.  Middle  Triassic  of  Southern  Alps  (Grenzbitumenzone  of  Monte  San  Giorgio,  Kt.  Tessin, 
Switzerland,  and  the  same  beds  near  Besano,  Lombardy,  Italy). 

Description 

Neurocranium  and  jaws.  Virtually  nothing  is  known  of  the  neurocranium.  Remains  are  observed  in  T 1548 
(lateral  view,  text-fig.  1)  and  in  T 2465  (dorsal  view,  text-fig.  2).  There  is  a very  prominent  postorbital  process 
which  lies  just  in  front  of  the  otic  process  of  the  palatoquadrate.  The  rostrum  appears  to  have  been  short 
and  blunt,  projecting  little  beyond  the  suborbital  ramus  of  the  palatoquadrate. 

The  palatoquadrate  (text-figs.  1 and  2)  shows  a large  postorbital  ramus  (slightly  more  than  half  the  total 
length  of  the  palatoquadrate)  with  a prominent  otic  process.  The  latter  articulates  with  the  postorbital  process 
of  the  neurocranium.  The  anterior  edge  of  the  otic  process  is  steeply  inclined.  The  posterior  edge  is  thickened 
and  thus  forms  a rim  which  limits  the  area  of  origin  of  the  m.  adductor  mandibulae  externus.  The  lower  end 
of  the  thickened  posterior  edge  forms  the  articular  condyle  of  the  palatoquadrate  which  fits  into  a facet  on 
the  posterior  end  of  Meckel’s  cartilage.  The  narrow,  tapering  suborbital  ramus  makes  up  slightly  less  than 
half  of  the  total  length  of  the  palatoquadrate.  A weak  elevation  of  its  dorsal  edge  in  its  anterior  portion 
(text-fig.  2)  forms  the  orbital  process  which  articulates  with  the  suborbital  shelf  of  the  neurocranium. 


RIEPPEL:  NEW  TRIASSIC  SHARK 


401 


text-fig.  1.  Skull  remains  in  Acronemus  tuberculatus  T 1548.  Abbreviations:  crt, 
ceratohyal;  hy,  hyomandibula;  lc,  labial  cartilages;  Me,  Meckel’s  cartilage;  nc,  neuro- 
cranium; po,  postorbital  process;  pq,  palatoquadrate.  Scale  equals  10  mm. 


Meckel’s  cartilages  (text-figs.  1 and  2)  are  elongated  and  moderately  deep.  The  ventral  margin  is  convex 
and  slightly  thickened.  The  thickened  ventral  rim  limits  the  site  of  insertion  of  the  m.  adductor  mandibulae 
externus.  The  dorsal  edge  is  concave  except  for  the  posterior  portion  where  the  dorsal  edge  is  straight. 

All  the  Monte  San  Giorgio  material  is  strongly  compressed,  but  from  a thickening  it  appears  that  both 
the  dorsal  edge  of  Meckel’s  cartilage  as  well  as  the  ventral  edge  of  the  palatoquadrate  formed  an  outwardly 
turned  shelf  for  the  support  of  the  teeth. 

Labial  cartilages.  Only  broken  fragments  are  observed  in  T 1548  (text-fig.  1)  along  the  ventral  edge  of  Meckel’s 
cartilage. 

Hyoid  arch.  The  articulation  of  the  hyomandibula  with  the  ceratohyal  is  preserved  just  posterior  to  the  jaw 
articulation  in  T 1548  (text-fig.  1).  The  jaw  suspension  in  a shark  with  a cleaver-shaped  palatoquadrate  may 
be  assumed  to  have  been  amphistylic  (Schaeffer  1967). 

Cephalic  spines.  These  are  not  recorded  in  any  specimen. 

Teeth.  The  teeth  of  Acronemus  tuberculatus  have  long  been  known  under  the  name  of  Acrodus  bicarinatus 
(Bassani  1886;  Alessandri  1910).  The  teeth  are  characterized  by  a single,  blunt  main  cusp.  No  accessory  lateral 
cusps  are  developed  (PI.  43,  fig.  1).  There  is  a marked  longitudinal  and  a marked  transverse  ridge  across  the 
crown.  The  ridges  meet  at  right  angles  on  the  apex  of  the  main  cusp  (text-fig.  3 and  PI.  43,  fig.  3).  From  the 
longitudinal  and  transverse  ridges  fine  striae  radiate  towards  the  edges  of  the  crown.  The  main  cusp  forms 
a bulbous  lingual  projection  which  overlaps  the  next  inner  tooth  of  the  same  tooth  family.  This  results  in  a 
supporting  mechanism  for  the  outer  functional  tooth  very  similar  to  the  one  observed  in  A.  lateralis. 

The  enameloid  layer  covering  the  crown  is  of  the  single-crystallite  type  (PI.  43,  fig.  4)  as  defined  by  Reif 
(1973),  without  a superficial  shiny  layer.  The  root  bears  no  expanded  lingual  torus.  Root  foramina  were  very 
difficult  to  observe  because  the  bituminous  matrix  does  not  allow  chemical  preparation  to  expose  these  foramina. 


402 


PALAEONTOLOGY,  VOLUME  25 


The  large  lateral  teeth  are  symmetrical,  with  the  main  cusp  in  a fairly  central  position.  The  crown  is  high 
and  the  lower  edge  of  the  root  is  distinctly  concave  (text-fig.  3).  Towards  the  symphysis  the  teeth  become 
progressively  smaller  and  the  crown  lower  and  asymmetrical  in  that  the  main  cusp  is  shifted  towards  the 
distal  side  of  the  crown.  The  lower  edge  of  the  root  becomes  less  concave  or  even  straight  (text-fig.  3).  Distal 
to  the  large  lateral  teeth  there  is  a series  of  very  low  but  very  elongated  and  distinctly  asymmetrical  teeth 
(text-fig.  3)  with  the  main  cusp  displaced  towards  the  mesial  side  of  the  tooth.  The  lower  edge  of  the  crown 
is  almost  straight.  Again,  the  teeth  diminish  in  size  towards  the  distal  end  of  the  jaw. 

Pectoral  girdle.  The  left  pectoral  girdle  is  preserved  in  T 1 548  (text-fig.  4).  Its  structure  clearly  resembles  that 
of  Acrodus  and  Hybodus  (Koken  1907).  The  suprascapular  portion  is  an  elongated,  spinous  structure  curved 
in  an  anterior  direction.  The  scapulocoracoid  and  the  ventral  coracoid  bar  are  poorly  preserved.  Below  the 
pectoral  girdle,  a single  basal  element  of  the  pectoral  fin  is  preserved  (text-fig.  4).  The  incompleteness  of  the 
specimen  does  not  allow  the  reconstruction  of  the  basal  skeleton  of  the  pectoral  fin. 

Finspines.  The  finspines  of  Acronemus  (syn.  Nemacanthus  tuberculatus)  are  relatively  short  and  broad  which 
results  in  characteristic  stout  proportions.  The  anterior  finspines  are  relatively  longer  and  relatively  broader 
than  the  posterior  finspines  (text-fig.  5;  PI.  43,  figs.  6,  7).  The  crown  of  the  anterior  finspine  is  frequently  of 
a broad-based,  upright  triangular  shape  (text-fig.  6a).  However,  it  may  become  somewhat  elongated  and 
recurved  to  a variable  degree  (text-fig.  5;  PI.  43,  figs.  6,  8).  The  posterior  finspines  are  relatively  smaller  and 
narrower  with  an  upright  crown  that  is  never  distinctly  curved  in  a posterior  direction. 

The  most  characteristic  feature  of  the  spines  is  their  conspicuous  tuberculation.  Large  and  rounded  tubercles 
are  arranged  in  a regular  pattern  on  which  it  is  possible  to  superimpose  straight  longitudinal  lines.  It  is  equally 
possible  to  superimpose  on  the  pattern  of  tuberculation  curved  and  obliquely  oriented  lines  which  run  in  a 
posterodorsal  direction  across  the  lateral  surface  of  the  crown  (text-fig.  6a).  There  is  in  some  spines  the  tendency 
of  the  tubercles  to  fuse  into  segments  of  ridges  running  in  a curved  posterodorsal  direction  across  the  lateral 
surface  of  the  crown  (text-fig.  6;  PI.  43,  figs.  6,  7).  The  arrangement  of  the  tubercles  along  curved,  obliquely 


RIEPPEL:  NEW  TRIASSIC  SHARK 


403 


text-fig.  3.  The  teeth  of  Acronemus  tuberculatus.  a:  T 3821,  articulated  part  of  dentition  (surface  abraded). 
Scale  equals  5 mm.  b-g:  T 3849,  isolated  teeth.  B-c,  large  lateral  teeth  in  lingual,  lateral,  and  occlusal  views; 
d-e,  successive  mesial  teeth  in  lingual  view;  f-g,  successive  distal  teeth  in  lingual  view. 


oriented  lines  reflects  the  growth  pattern  of  the  spine.  The  completely  closed  enameloid  mantle  of  euselachian 
finspines  frequently  shows  similarly  oriented  growth  lines  (Maisey  1977). 

In  some  specimens  the  tubercles  are  small  and  scattered  irregularly  (text-fig.  6b).  The  scattered  arrangement 
is  probably  due  to  wear,  but  the  smaller  size  of  the  tubercles  results  in  a larger  number  of  longitudinal  rows 
counted  across  the  base  of  the  crown.  Also,  the  anterior  finspines  have  a larger  number  of  longitudinal  rows 
of  tubercles  than  the  posterior  finspines  since  the  anterior  spines  are  relatively  broader.  On  the  anterior 
finspines  there  are  from  10  to  23  longitudinal  rows  of  tubercles  counted  at  the  base  of  the  crown,  the  average 
ranging  from  13  to  15  rows.  On  the  posterior  finspines  there  are  usually  13  rows  of  tubercles,  variation  ranging 
from  9 to  13  rows. 

Along  the  leading  edge  of  the  crown  the  finspines  bear  a more  or  less  conspicuously  developed  longitudinal 
ridge.  In  T 1289  this  ridge,  as  far  as  preserved,  retains  an  enameloid  covering.  The  anterior  ridge  results  from 
the  fusion  of  a single  anteromesial  row  of  tubercles.  This  is  demonstrated  by  the  finspines  of  T 3825  in  which 
this  fusion  is  incomplete.  The  anterior  ridge  usually  extends  ventrally  along  the  leading  edge  of  the  finspine 
beyond  the  lowermost  anterior  tubercles  up  to  close  to  an  anterior  projection  formed  by  the  upper  part  of 
the  root.  This  anterior  projection  indicates  the  level  to  which  the  finspine  was  inserted  in  the  epaxial  trunk 
musculature. 

The  posterior  wall  of  the  finspine  is  concave.  The  posterior  opening  of  the  central  cavity  extends  upwards 
to  the  level  of  the  lowermost  posterior  tubercles.  A longitudinal  row  of  tubercles  runs  along  the  posterolateral 
edge  of  the  crown,  but  these  tubercles  do  not  differ  in  size  or  shape  from  those  covering  the  lateral  surface 
of  the  crown. 

The  line  which  connects  the  lowermost  tubercles  along  the  anterior  and  posterior  edges  of  the  finspine  is 
chosen  to  represent  the  base  of  the  crown.  The  angle  6 between  the  base  of  the  crown  and  the  normal  on 
the  longitudinal  axis  of  the  finspine  (text-fig.  5)  ranges  from  8 to  24  in  the  available  spines. 

Five  finspines  of  Acronemus  were  sectioned  at  successive  levels  along  their  long  axis.  A composite  and 
slightly  diagrammatic  representation  of  finspine  histology  shown  in  text-fig.  7a  may  be  compared  to  the 
section  figured  in  text-fig.  7b.  There  is  a lamellar  inner  trunk  layer  which  is  only  apically  distributed,  above 
the  posterior  opening  of  the  central  cavity.  The  central  cavity  is  generally  small  at  this  level,  sometimes  even 
(ontogenetically?)  reduced  to  the  size  of  an  ordinary  vascular  canal  (T  1431,  section  Nr.  8).  The  lamellar 
inner  trunk  layer  is  surrounded  by  a well  vascularized  outer  trunk  layer  which  also  forms  the  entire  root  of 
the  finspine.  The  central  cavity  is  displaced  backwards  which  results  in  a thick  anterior  wall  largely  made  up 
by  the  vascularized  outer  trunk  layer.  The  vascular  canals  and  their  surrounding  denteons  are  rather  small 
and  widely  spaced  which  gives  the  outer  trunk  layer  a relatively  compact  appearance.  Towards  the  leading 
edge  of  the  spine  a distinctly  larger,  longitudinally  running  vascular  canal  is  observed  lying  straight  in  front 
of  the  central  cavity  (text-fig.  7).  The  tubercles  are  formed  by  the  mantle  component,  but  no  clearcut  boundary 
separates  the  mantle  component  from  the  outer  trunk  layer  (for  the  terminology  of  finspine  histology  see 
Maisey  1979). 


404 


PALAEONTOLOGY,  VOLUME  25 


text-fig.  4.  Tentative  reconstruction  of  Acronemus  tuberculatus,  based  on  T 1548.  Approx.  xO-4. 


text-fig.  5.  Variation  in  size  and  shape  of  the  anterior  (left)  and  posterior  (right)  finspines  of  Acronemus 
tuberculatus.  a,  T 1178;  B,  T 2465;  c,  T 1548.  Scale  equals  10  mm.  For  further  explanations  see  text. 


text-fig.  6.  Ornamentation  of  the  anterior  (left)  and  posterior  (right)  finspines  of  Acronemus  tuberculatus. 
a,  T 3818;  b,  T 2465.  Scale  equals  10  mm. 


Scales.  Flank  scales  of  the  specimen  T 1548  are  shown  on  PI.  43,  fig.  5.  They  are  of  the  non-growing,  placoid 
type.  They  bear  a recurved,  lanceolate  crown  which  is  ornamented  with  three  or  five  widely  spaced  longitudinal 
striae. 


COMPARISON  WITH  THE  GENUS  NEM ACANTHUS 

The  finspines  of  Acronemus  tuberculatus  were  originally  referred  to  the  genus  Nemacanthus  Ag. 
(Bellotti  1873,  in  Bassani  1886).  Type  species  of  the  latter  genus  is  Nemacanthus  monilifer  Ag.,  with 
which  comparison  thus  must  proceed.  N.  monilifer  is  represented  by  finspines  from  the  Rhaetic  of 
England  which  differ  in  several  respects  from  those  of  A.  tuberculatus.  Assuming  a relatively 


406 


PALAEONTOLOGY,  VOLUME  25 


text-fig.  7.  Finspine  histology  of  Acronemus  tuberculatus.  a.  Semidiagrammatic  representation  of 
a transverse  section  through  the  crown  of  a finspine,  based  mainly  on  section  T 3847/7.  Abbreviations: 
cc,  central  cavity;  lvc,  longitudinal  vascular  canal,  b.  Ground  section  Nr.  T 3847/7.  Approx,  x 30. 


constant  structure  of  the  finspines  within  shark  genera,  this  justifies  the  erection  of  a new  genus 
to  include  the  spines  of  tuberculatus  (Maisey,  pers.  comm.). 

The  finspines  of  N.  monilifer  (comparative  material  used  for  this  study  is  housed  in  the  British 
Museum  (Natural  History)  and  is  listed  in  the  Appendix)  are  long  and  slender  in  comparison  to 
Acronemus.  They  are  regularly  curved  in  a posterior  direction.  If  the  total  height  of  the  finspines 
is  divided  by  their  maximum  width,  the  values  obtained  for  two  complete  specimens  of  N.  monilifer 
are  8-56  (BM(NH)  P.8328)  and  9-3  (BM(NH)  P.46830).  Nine  anterior  finspines  of  A.  tuberculatus 
have  a corresponding  mean  value  of  3T5,  while  six  posterior  finspines  have  a mean  value  of  4T. 
In  both  genera  the  finspines  bear  an  anterior  enamelled  ridge  as  well  as  a tuberculate  mantle 
ornament,  but  the  tubercles  are  fewer  in  number  in  N.  monilifer.  In  the  latter  genus  one  usually 


RIEPPEL:  NEW  TRIASSIC  SHARK 


407 


text-fig.  8.  Semidiagrammatic  sections  through  the  finspines  of  chimaeroids  and  selachians,  a,  Leptacanthus 
longissimus,  BM(NH)  Egerton  coll.;  b,  Ctenacanthus  angustus,  BM(NH)  P.9581;  c,  Nemacanthus  monilifer, 
BM(NH)  P.15497,  section  10;  d,  Palaeobates  keuperinus,  BM(NH)  P.7604;  e,  Hybodus  lawsoni,  BM(NH) 
P.2174a;  f,  Hybodus  acutus,  BM(NH)  P.6157.  Scale  equals  5 mm. 


counts  less  than  ten  rows  of  tubercles  across  the  base  of  the  crown,  variation  ranging  from  7 rows 
in  BM(NH)  P.46830  to  1 1 rows  in  BM(NH)  P.2852,  whereas  in  A.  tuberculatus  one  usually  counts 
more  than  10  rows  (9-23)  of  tubercles  across  the  base  of  the  crown.  The  connection  of  the  lowermost 
tubercles  at  the  anterior  and  posterior  margins  of  the  finspine  results  in  a line  which  slants  much 
more  steeply  in  a posterodorsal  direction  in  Nemacanthus  as  compared  to  Acronemus.  This  line 
intersects  the  normal  on  the  long  axis  of  the  spine  at  an  angle  9 (text-fig.  5)  which  in  N.  monilifer 
ranges  from  60°  (BM(NH)  P.46830)  to  78°  (BM(NH)  P.51433),  but  from  8-24°  in  Acronemus.  A 
function  of  this  angle  is  the  fact  that  the  tubercles  approach  the  posterior  edge  of  the  spine  only 
towards  its  apex  in  N.  monilifer.  The  extent  of  tuberculation  in  N.  monilifer  may  either  be  due  to 
ontogenetic  changes  or  to  wear.  In  the  specimen  BM(NH)  P.8328  isolated  rudiments  of  a few  worn 
or  resorbed  tubercles  can  be  observed  on  the  posteroventral  part  of  the  lateral  surface  of  the  crown, 
an  area  which  lies  well  below  the  regularly  tuberculated  part  of  the  crown  but  above  the  ventral 
end  of  the  enamelled  anterior  ridge.  The  angle  9 might  be  an  indication  of  the  degree  of  posterior 
inclination  of  the  finspine  insertion  relative  to  the  long  axis  of  the  body.  Specimen  BM(NH)  P.8328 
suggests  that  the  degree  of  posterior  inclination  of  finspine  insertion  increased  during  ontogeny  in 
N.  monilifer. 

In  both  genera  the  posterior  wall  of  the  finspine  is  concave.  In  Acronemus  longitudinal  rows  of 
unmodified  tubercles  run  along  the  posterolateral  edges  of  the  crown.  In  some  specimens  of  N. 
monilifer  rows  of  small  but  pointed  denticles  are  observed  running  along  the  posterolateral  edges 
of  the  crown  (Maisey  1977,  PI.  1.  fig.  d,  and  specimens  BM(NH)  P.1882,  P.51433).  Such 
posterolateral  rows  of  denticles  are  also  observed  in  some  specimens  of  the  genus  Ctenacanthus 
(BM(NH)  P.2525,  P.2529). 

The  histology  of  the  finspines  is  basically  similar  in  Acronemus  and  Nemacanthus , except  for  the 
detailed  histology  of  the  thick  anterior  wall.  In  N.  monilifer  the  vascular  canals  are  large  and  closely 
juxtaposed,  which  results  in  the  characteristic  open  spongy  texture  of  the  anterior  wall  (Maisey 
1977,  pi.  2,  fig.  c;  Stromer  1927,  text-fig.  12).  This  contrasts  with  the  much  more  compact  texture 
of  the  anterior  finspine  wall  in  Acronemus. 


DISCUSSION:  THE  CLASSIFICATION  OF  ACRONEMUS 

Maisey  (1975)  subdivided  the  phalacanthous  sharks  ( sensu  Zangerl  1973)  into  three  groups  of 
ordinal  rank,  the  Hybodontiformes,  the  Ctenacanthiformes,  and  the  Euselachiformes.  The 
distinction  of  these  three  groups  is  based  mainly  on  a detailed  study  of  finspine  structure.  On  the 
basis  of  finspine  structure  alone,  Acronemus  clearly  has  to  be  classified  with  the  Ctenacanthiformes 
as  defined  by  Maisey  (1975).  The  structure  of  the  palatoquadrate  ties  in  well  with  such  a conclusion. 


408 


PALAEONTOLOGY,  VOLUME  25 


It  is  cleaver-shaped  in  Acronemus  as  in  some  Palaeozoic  ctenacanthiformes  such  as  Goodrichthys 
(Moy-Thomas  1936),  whereas  Hybodus  and  its  allies  show  a reduction  of  the  otic  process.  The 
cleaver-shaped  palatoquadrate  probably  is  a primitive  feature,  however  (Schaeffer  1967).  Cephalic 
spines  are  not  reported  in  ctenacanthiform  sharks  or  in  Acronemus. 

However,  other  features  do  not  support  the  classification  of  Acronemus  with  the  Ctenacanthi- 
formes. The  teeth  of  Acronemus  are  very  similar  to  those  of  Acrodus,  and  they  lack  an  expanded 
lingual  torus  on  the  root.  The  teeth  of  Palaeozoic  ctenacanthiforms  are  of  the  multicuspid  cladodont 
type  with  a lingual  torus  on  the  root  (Maisey  1975).  Which  type  of  tooth  was  possessed  by  the 
Lower  Triassic  (Stensio  1921,  1932)  or  Rhaetic  Nemacanthus  is  controversial  (Maisey  1977).  No 
associated  Nemacanthus  specimen  has  yet  been  found. 

All  Palaeozoic  ctenacanthiform  sharks  have  scales  of  a composite,  growing  type  (Reif  1978). 
Acronemus  has  placoid  scales  which  are  typical  of  pre-Rhaetian  hybodontids  (Reif  1978,  p.  126), 
as  well  as  of  euselachians. 

These  latter  features  of  Acronemus  might  indicate  that  the  ctenacanthiform  and  hybodontiform 
sharks  can  only  be  recognized  by  their  finspine  structure.  However,  they  might  also  indicate  that 
the  distinction  between  the  two  orders  is  not  as  clearcut  as  it  would  appear  (Schaeffer  and  Williams 
1977).  In  fact,  some  mixture  of  hybodontiform  and  ctenacanthiform  features  was  noted  by  Dick 
(1978)  in  his  description  of  Tristychius. 

Among  all  those  finspines  examined  by  me  (see  Appendix)  there  is  enough  variability  to 
substantiate  this  point.  The  only  ctenacanthiform  features  which  are  really  constant  throughout 
the  material  examined  are  a flat  or  concave  posterior  wall  and  a posteriorly  displaced  central  cavity 
which  results  in  a relatively  thick  anterior  wall.  In  hybodontiform  finspines  the  posterior  wall  may 
be  flat  or  convex  to  a variable  degree,  and  the  anterior  wall  of  the  finspine  is  never  much  thicker 
than  the  posterior  wall.  This  latter  feature  is  a function  of  the  posteriorly  displaced  central  cavity 
of  ctenacanthiform  spines,  but  the  position  of  the  central  cavity  may  again  be  correlated  at  least 
to  some  extent  with  the  degree  of  convexity  of  concavity  of  the  posterior  wall.  If  the  central  cavity 
is  held  at  a constant  distance  from  the  leading  edge  of  the  spine  it  will  appear  in  a central  position 
if  the  posterior  wall  is  strongly  convex,  but  it  will  appear  in  a posteriorly  displaced  position  if  the 
posterior  wall  is  strongly  concave.  The  crown  of  the  Hybodus  finspine  BM(NH)  P.57794  is  broken 
at  successive  levels.  Examination  of  the  four  levels  of  breakage  showed  that  the  anterior  wall  gets 
relatively  thicker  towards  the  apex  of  the  spine  as  the  diameter  of  the  central  cavity  diminishes. 
A finspine  of  Nemacanthus  monilifer  (BM(NH)  P.  1 5497)  which  was  sectioned  at  successive  levels 
from  top  to  bottom  shows  that  the  relative  thickness  of  the  anterior  wall  remains  constant  throughout 
the  length  of  the  spine. 

The  anterior  wall  of  the  ctenacanthiform  finspine  is  said  frequently  to  be  of  an  open  spongy 
texture  (Maisey  1975).  This  in  fact  is  only  true  for  N.  monilifer  (and  for  the  euselachian  genus 
Breviacanthus  Maisey  1976).  The  anterior  wall  of  the  Ctenacanthus  finspine  has  exactly  the  same 


EXPLANATION  OF  PLATE  43 
Acronemus  tuberculatus 

Fig.  1.  Tooth  T 3849  in  lingual  view,  x 10. 

Fig.  2.  Tooth  T 3849  in  lateral  view,  x 10. 

Fig.  3.  Tooth  T 3849  in  occludal  view,  x 10. 

Fig.  4.  Tooth  T 3849,  cross-section  etched  with  2n-HCl  for  5 sec.  to  show  the  single-cristallite  enamel,  approx, 
x 150. 

Fig.  5.  Placoid  scale  from  specimen  T 1548,  approx,  x 130. 

Fig.  6.  Anterior  finspine  from  specimen  T 3818,  x 1-8. 

Fig.  7.  Posterior  finspine  from  specimen  T 3818,  x 1-8. 

Fig.  8.  Anterior  finspine  T 1289,  x 1-25. 


PLATE  43 


RIEPPEL,  Middle  Triassic  shark 


410 


PALAEONTOLOGY,  VOLUME  25 


vermiculate  texture  produced  by  relatively  small  vascular  canals  as  is  observed  in  typical 
hybodontiform  finspines.  Acronemus  has  an  even  more  compact  finspine  histology. 

Hybodontiform  finspines  are  said  to  show  a concentric  ring  of  longitudinally  running  vascular 
canals  around  the  central  cavity  (Maisey  1975,  1978).  This  is  not  always  very  clearly  differentiated. 

In  addition,  a comparison  with  chimaeroid  finspines  (see  Appendix)  as  well  as  with  acanthodian 
finspines  (Krebs  1960)  indicates  that  concentrically  arranged  vascular  canals  are  a primitive 
character-state  in  selachians.  However,  in  all  the  hybodontiform  and  ctenacanthiform  finspines 
examined  by  me  there  is  a conspicuously  larger  longitudinal  vascular  canal  which  lies  straight  in 
front  of  the  central  cavity. 

An  important  difference  between  hybodontiform  and  ctenacanthiform  finspines  is  the  ornamenta- 
tion of  the  posterior  wall  in  hybodontiforms,  with  one  or  two  longitudinal  rows  of  denticles,  said 
to  be  absent  in  ctenacanthiforms  (Maisey  1975).  Hybodontiform  finspines  typically  carry  a double 
or  a single  row  of  rather  large  denticles  close  to,  or  on,  the  midline  of  the  posterior  wall.  Such 
denticles  indeed  are  absent  in  ctenacanthiforms.  But  in  genera  such  as  Ctenacanthus  and 
Nemacanthus  (Maisey  1977,  PI.  1,  fig.  d,  and  quotations  above)  rows  of  small  denticles  are  arranged 
along  the  posterolateral  edges  of  the  crown  which  has  a concave  posterior  wall.  Maisey  (1975) 
considers  the  denticles  on  the  posterior  wall  of  hybodontiform  finspines  as  either  highly  modified 
mantle  components  or  as  specialized  scales  secondarily  fused  to  the  finspine.  The  posterolateral 
denticles  of  ctenacanthiform  finspines  are  merely  considered  to  represent  ‘posterolateral  rows  of 
pointed  tubercles’  (Maisey  1977,  p.  265).  This  distinction  appears  meaningless  since  it  is  impossible 
to  determine  the  morphogenetic  status  of  the  hybodontiform  versus  the  ctenacanthiform  denticles. 

If  the  denticles  are  considered  homologous,  it  is  possible  to  claim  that  posterolaterally  placed 
denticles  are  a primitive  character  state  while  denticles  which  have  shifted  to  a posteromesial 
position  represent  an  advanced  character  state.  Outgroup  comparison  with  chimaeroid  finspines 
(see  Appendix)  demonstrates  that  in  this  group  there  are  usually  two  longitudinal  rows  of 
posterolaterally  placed  denticles,  arranged  along  the  edges  of  a concave  posterior  wall  (text-fig. 
8a).  One  undescribed  chimaeroid  finspine  (BM(NH)  P.2850)  shows  a single  row  of  denticles  along 
the  midline  of  the  concave  posterior  wall.  This  single  row  of  denticles  may  well  have  originated 
from  double  rows  of  posterolaterally  placed  denticles  that  have  shifted  towards  the  midline  of  the 
posterior  wall.  A similar  phenomenon  has  been  noted  for  Hybodus  finspines  (Maisey  1978).  A 
single  row  of  posteromesially  arranged  denticles  appears  to  originate  from  the  superposition  of 
two  collateral  rows  of  denticles  on  the  midline  of  the  posterior  wall  (text-fig.  8e-f).  This  point  is 
further  corroborated  by  a consideration  of  the  finspines  of  Palaeobates  keuperinus  (text-fig.  8d).  In 
fact,  the  structure  of  these  finspines  is  perfectly  intermediate  between  the  ctenacanthiform  and 
hybodontiform  types  and  thus  illustrates  the  potential  difficulties  in  recognizing  the  two  orders. 

In  P.  keuperinus  the  posterior  wall  of  the  finspine  is  weakly  convex  (as  in  some  Hybodus  finspines), 
but  it  carries  two  posterolaterally  placed  rows  of  denticles  (as  in  ctenacanthiforms).  The  central 
cavity  is  small  and  somewhat  displaced  backwards  as  in  ctenacanthiforms,  but  the  posterior  wall 
still  retains  a considerable  thickness  although  it  is  thinner  than  the  anterior  wall. 

In  summary  it  can  be  stated  that  hybodontiform  and  ctenacanthiform  sharks  together  form  a 
group  of  selachians  with  a finspine  structure  involving  an  apically  distributed  inner  trunk  layer 
surrounded  by  a trabecular  outer  trunk  layer.  The  details  of  the  histology  of  the  outer  trunk  layer 
are  quite  variable  and  only  characteristic  at  the  generic  level.  There  is  no  clearcut  boundary  between  I 
the  outer  trunk  layer  and  the  mantle  component.  The  mantle  ornamentation  may  be  tuberculate 
or  costate.  As  evidenced  by  the  genus  Acronemus,  the  orders  Hybodontiformes  and  Ctenacanthi- 
formes  can  at  present  be  recognized  on  the  basis  of  finspine  structure  alone.  Ctenacanthiform 
finspines  show  a flat  or  concave  posterior  wall  with  two  rows  of  posterolaterally  placed  denticles 
(primitive)  and  a posteriorly  displaced  central  cavity.  Hybodontiform  finspines  show  a flat  or 
convex  posterior  wall  with  one  or  two  rows  of  denticles  close  to  or  on  the  midline  of  the  posterior 
wall  (derived).  The  central  cavity  is  rather  centrally  positioned.  However,  intermediate  forms  such 
as  P.  keuperinus  and  Tristy chius  (Dick  1978)  do  occur. 

On  the  basis  of  finspine  structure,  Acronemus  can  be  classified  as  a ctenacanthiform  shark.  The 


RIEPPEL:  NEW  TRIASSIC  SHARK 


411 

posterior  wall  of  the  finspine  is  concave,  and  the  central  cavity  displaced  backwards.  If  such  an 
arrangement  is  accepted  it  must  be  concluded  that  neither  the  ‘hybodontiform’  tooth  structure  nor 
the  presence  of  placoid  scales  can  be  used  to  distinguish  hybodontiform  and  ctenacanthiform 
sharks  from  the  Triassic. 

Acknowledgements.  I thank  Dr.  C.  Patterson  who  received  me  at  the  British  Museum  (Natural  History), 
granting  me  free  access  to  the  collection  of  fossil  fishes.  Drs.  C.  Patterson,  Chr.  Duffin,  and  W.-E.  Reif  all 
provided  some  opportunity  to  discuss  the  material  presented  in  this  study.  Preparative  and  photographic 
work  was  done  by  H.  Lanz.  The  SEM-micrographs  were  taken  at  the  Institut  fur  Pflanzenbiologie  der 
Universitat  Zurich  by  U.  Jauch.  Text-fig.  5,  6,  and  7a  were  prepared  by  O.  Garraux. 

REFERENCES 

alessandri,  d.  de.  1910.  Studii  sui  pesci  triasici  della  Lombardia.  Mem.  Soc.  ital.  Sci.  Nat.,  Milano,  7,  1-145. 
bassani,  fr.  1886.  Sui  fossili  e sulfeta  degli  schisti  bituminosi  triasici  di  Besano  in  Lombardia.  Atti  Soc.  ital.  Sci. 
Nat.,  Milano,  29,  15-72. 

dick,  J.  R.  F.  1978.  On  the  Carboniferous  shark  Tristychius  arcuatus  Agassiz  from  Scotland.  Trans.  R.  Soc. 
Edinburgh,  70,  63-109. 

koken,  e.  1907.  Ueber  Hybodus.  Geol.  Palaeont.  Abh.  (n.f.),  5,  261-276. 

krebs,  b.  1960.  Ueber  einen  Flossenstachel  von  Gyracanthus  (Acanthodii)  aus  dem  Oberkarbon  Englands. 
Eclogue  geol.  Helv.  53,  811-827. 

kuhn,  E.  1945.  Ueber  Acrodus- Funde  aus  dem  Grenzbitumenhorizont  der  anisischen  Stufe  der  Trias  des  Monte 
San  Giorgio  (Kt.  Tessin).  Eclogae  geol.  Helv.  38,  662-673. 
maisey,  j.  G.  1975.  The  interrelationships  of  phalacanthous  selachians.  N.  Jb.  Geol.  Palaeont.  Mh.  1975  (9), 
553-567. 

— 1976.  The  Middle  Jurassic  selachian  fish  Breviacanthus  n.g.  N.  Jb.  Geol.  Palaeont.  Mh.  1976  (7),  432-438. 

— 1977.  The  fossil  selachian  fishes  Palaeospinax  Egerton,  1872  and  Nemacanthus  Agassiz,  1837.  Zool.  J.  Linn. 
Soc.  60,  259-273. 

1978.  Growth  and  form  of  finspines  in  hybodont  sharks.  Palaeontology,  21,  657-666. 

— 1979.  Finspine  morphogenesis  in  squalid  and  heterodontid  sharks.  Zool.  J.  Linn.  Soc.  66,  161-183. 
moy-thomas,  J.  a.  1936.  The  structure  and  affinities  of  the  fossil  elasmobranch  fishes  from  the  Lower 

Carboniferous  rocks  of  Glencartholm  Eskdale.  Proc.  zool.  Soc.,  Lond.  1936,  761-788. 
peyer,  b.  1957.  Ueber  die  morphologische  Deutung  der  Flossenstacheln  einiger  Haifische.  Mitt,  naturf.  Ges.  Bern 
(n.f.),  14,  159-176. 

reif,  w.-e.  1973.  Morphologie  und  Ultrastruktur  des  Hai-‘Schmelzes’.  Zoologica  Scripta,  2,  231-250. 

— 1978.  Types  of  morphogenesis  of  the  dermal  skeleton  in  fossil  sharks.  Palaeont.  Z.  52,  110-128. 
Schaeffer,  b.  1967.  Comments  on  elasmobranch  evolution.  In  gilbert,  p.  w.,  mathewson,  r.  f.  and  rall,  d.  p. 

(eds.).  Sharks,  Skates  and  Rays.  John  Hopkins  Press,  Baltimore.  Pp.  3-35. 

— and  williams,  m.  1977.  Relationships  of  fossil  and  living  elasmobranchs.  Amer.  Zool.  17,  293-302. 
stensio,  e.  1921.  Triassic  fishes  from  Spitzbergen,  Pt.  1.  Adolf  Holzhausen,  Vienna,  i-xxviii,  1-307,  pis.  1-35. 
1932.  Triassic  fishes  from  East  Greenland,  collected  by  the  Danish  expeditions  in  1929-1931.  Medd.  om 

Grenl.  83,  1-305. 

stromer,  e.  1927.  Ergebnisse  der  Forschungsreisen  Prof.  E.  Stromers  in  den  Wiisten  Aegyptens.  2.  Wirbeltier- 
Reste  der  Baharije-Stufe  (unterstes  Cenoman).  9.  Die  Plagiostomen,  mit  einem  Anhang  fiber  kano-  und 
mesozoische  Rfickenflossenstacheln  von  Elasmobranchiern.  Abh.  Bayr.  Akad.  Wiss.,  math.-natw.  Abt.  31, 
1-64. 

woodward,  a.  s.  1891.  Catalogue  of  the  fossil  fishes  in  the  British  Museum  ( Natural  History),  Vol.  3.  British 
Museum  (Natural  History),  London.  Pp.  i-xliv,  1-567,  pis.  1-16. 
zangerl,  r.  1973.  Interrelationships  of  early  chondrichthyans.  In  greenwood,  p.  h.,  miles,  r.  s.  and  Patterson, 
c.  (eds.).  Interrelationships  of  fishes.  Zool.  J.  Linn.  Soc.  53,  Suppl.  1,  1 14. 

O.  RIEPPEL 

Palaeontologisches  Institut  u.  Museum 
der  Universitat  Zfirich 
Kfinstlergasse  16 
CH-8006  Zfirich 

Typescript  received  7 January  1981  Switzerland 


412 


PALAEONTOLOGY,  VOLUME  25 


APPENDIX 

All  the  comparative  material  used  to  study  finspine  structure  is  housed  in  the  British  Museum  (Natural 
History).  The  specimen  numbers  are  as  follows. 

Sharks:  Hybodus  sp.,  P.47145,  P.57794;  acutus , P.6157;  ensatus , P.2162;  lawsoni , P. 2174a;  minor , P.51443; 
reticulatus , one  uncatalogued  specimen;  Ctenacanthus  sp.,  P.2525,  P.28928-9.  P.34865;  angustus,  P.9581, 
P.9262;  pustulatus,  P.2529;  Nemacanthus  monilifer,  P.1882,  P.2852,  P.2854,  P.8328,  P.  1 5497,  P.46830,  P.51433; 
Palaeobates  keuperinus,  P.2161,  P.7604. 

Chimaeroids:  Edaphodon  sp.,  P.3097;  Leptacanthus  longissimus , Egerton  coll.;  Myriacanthus  paradoxus, 
P.  1736a,  one  uncatalogued  specimen;  undescribed  genus,  P.2850. 

Ground  sections — Sharks:  Breviacanthus  brevis,  P.2851;  Ctenacanthus  angustus,  P.9581;  Lonchidion  sp., 
CP.  12.9.64;  Nemacanthus  monilifer,  P.2217,  P.15497;  Wracanthus,  P.45768.1. 

Chimaeroids:  Callorhynchus  callorhynchus,  Zool.  Dept.  1935.4.23.17;  Deltoptychius  armigurus,  P.11358; 
Edaphodon  sp.,  P.11807;  ? Erismacanthus  sp.,  P.6257;  Metopacanthus  granulatus,  P.43065;  Myriacanthus 
paradoxus,  P.1736. 


A NEW  SPECIES  OF  THE  FISH  AMIA  FROM  THE 
MIDDLE  EOCENE  OF  BRITISH  COLUMBIA 

by  MARK  V.  H.  WILSON 


Abstract.  A new  species  of  amiid  fish  is  described  from  a semi-articulated  skeleton  found  in  Middle  Eocene 
freshwater  shales  of  the  Allenby  Formation,  Princeton  Group,  south-central  British  Columbia.  The  new 
species  is  assigned  to  Amia  because  it  lacks  Kindleia  specializations  such  as  styliform  teeth,  and  it  shares  skull 
specializations  with  A.  calva,  A.  scutata,  and  A.  uintaensis.  The  new  species  is  reconstructed  as  a deep-bodied 
piscivore  with  large  jaws  and  strong,  sharp  teeth.  The  holotype  is  the  first  identifiable  skeleton  to  be  found 
among  many  amiid  scales  recovered  from  numerous  fossil-fish  assemblages  in  southern  British  Columbia  and 
northern  Washington  State. 

This  paper  presents  a description  and  partial  reconstruction  of  a new  species  of  amiid  fish,  based 
on  a single  partially  articulated  skeleton  and  several  disarticulated  skull  bones  from  Middle  Eocene 
freshwater  shales  in  British  Columbia.  The  significance  of  the  discovery  lies  in  the  relatively  complete 
information  obtainable  about  the  anatomy  of  this  fish,  and  the  resulting  implications  for  the 
taxonomy  of  fossils  of  Amia  elsewhere  in  North  America. 

Amia  calva,  the  only  living  species  of  amiid,  is  confined  to  the  fresh  waters  of  south-eastern 
North  America,  but  Late  Cretaceous  and  Tertiary  records  of  amiids  are  widespread  in  central  and 
western  parts  of  the  continent  (text-fig.  1),  as  well  as  in  Europe  and  Asia.  North  American  fossil 
amiids  were  reviewed  by  Boreske  (1974),  who  recognized  three  valid  fossil  species  in  a single  genus 
from  the  twenty-three  species  and  seven  genera  previously  described.  Many  of  the  rejected  names 
were  based  on  poorly  preserved  material  or  on  single  vertebrae,  and  were  considered  to  be  junior 
synonyms  or  nomina  dubia.  The  three  fossil  species  considered  valid  by  Boreske  are  the  Late 
Cretaceous  to  Middle  Eocene  Kindleia  fragosa,  the  Palaeocene  to  Early  Oligocene  A.  uitaensis,  and 
the  Oligocene  A.  scutata. 

None  of  the  previously  known  Eocene  occurrences  of  Amia  from  British  Columbia,  Washington, 
and  Oregon  had  been  identified  to  species,  because  they  consisted  primarily  of  undiagnostic  scales 
(Cavender  1968;  Wilson  1977a,  1978a,  1979,  1980)  which  were  nevertheless  very  common  fossils. 
These  westernmost  records  seemed  to  reflect  habitats  similar  to  those  preferred  by  other  Amia 
species  including  A.  calva  (Scott  and  Crossman  1973):  warm,  shallow,  swampy  conditions,  as 
evidenced  by  lithology  and  the  associated  fishes,  insects,  and  plants  (Wilson  1980). 

The  discovery  of  the  Amia  specimens  described  here  resulted  from  work  by  palaeobotanists  on 
silicified  plant  fossils,  preserved  in  an  outcrop  of  alternating  chert  and  coal  layers  in  the  Allenby 
Formation  (Boneham  1968;  Miller  1973;  Robison  and  Person  1973;  Basinger  1976;  Basinger  and 
Rothwell  1977).  James  Basinger  discovered  a trionychid  turtle  in  shales  immediately  overlying  the 
chert.  When  I revisited  the  site  in  1977  and  1978,  I obtained  the  fish  specimens  described  here, 
along  with  coprolites  containing  fish  bones,  and  disarticulated  remains  of  small  suckers 
(Catostomidae:  Amyzon  sp.),  and  trout-perches  (Percopsidae:  Libotonius  sp.).  The  fish  occur  in  the 
shales  with  carbonized  plant  fossils  which  include  stems  and  twigs,  dicotyledonous  leaves, 
taxodiaceous  leafy  shoots,  seeds,  ferns,  and  amber  (Wilson  1980). 

GEOLOGY  AND  AGE 

The  specimens  were  found  in  a hard  black  siliceous  shale  immediately  overlying  a 10-m-thick 
outcrop  of  interbedded  carbonaceous  chert  and  coal  which  extends  into  the  Similkameen  River 

IPalaeontology,  Vol.  25,  Part  2,  1982,  pp.  413-424.| 


414 


PALAEONTOLOGY,  VOLUME  25 


text-fig.  1.  Map  of  North  America  showing  the  fossil  locality.  Also 
shown  are  Late  Cretaceous  through  Oligocene  sites  from  which 
remains  of  Amia  and/or  Kindleia  have  been  reported  (modified  after 
Boreske  1974),  together  with  the  distribution  of  Recent  Amia  calva 
(stipple,  after  Scott  and  Crossman  1973). 

from  its  east  bank,  8 -4  km  south  of  Princeton,  British  Columbia  (text-fig.  1).  According  to  Boneham 
(1968)  the  fossiliferous  layers  are  550  m above  the  Princeton -Black  Coal  of  the  Allenby  Formation, 
Princeton  Group.  The  Allenby  Formation  consists  of  sedimentary  rocks,  deposited  in  fresh  water, 
and  yields  an  abundant  assemblage  of  fishes  and  insects  at  several  localities  in  the  Princeton  Basin 
(Wilson  1977a,  19776,  19786,  1980). 

The  age  of  the  fossils  is  Middle  Eocene,  based  on  potassium -argon  ages  of  approximately  47-50 
million  years  for  the  Allenby  Formation  (Hills  and  Baadsgaard  1967),  and  on  the  occurrence  of 
the  mammalian  genus  Trogosus  elsewhere  in  the  formation  (Russell  1935).  Other  occurrences  of 
fossil  amiid  scales  in  British  Columbia  and  in  the  Klondike  Mountain  Formation  of  Washington 
State  are  of  a similar  geologic  age  (Wilson  1977a,  1978a,  1979). 

MATERIALS  AND  METHODS 

The  Amia  fossils  occur  as  fractured  bone,  spread  parallel  to  bedding  planes.  The  hard  and  tough 
matrix  contains  a faithful  imprint  of  the  original  shape  and  ornamentation  of  the  bones  and, 
therefore,  the  specimens  were  prepared  by  removing  the  fossilized  bone  with  mechanical  tools  and 
an  ultrasonic  probe.  Casts  of  the  resulting  impressions  were  made  in  black  latex.  Measurements 
were  taken  from  the  original  shale  moulds.  For  photographic  purposes  the  black-latex  casts  were 
coated  with  ammonium  chloride.  The  reconstructions  were  made  by  drawing  the  outlines  of  the 
bones  at  uniform  scale,  using  a camera  lucida  attached  to  a Wild  M8  stereomicroscope,  and  then 
graphically  assembling  the  bone  outlines  with  regard  to  perspective  distortions  and  suture  outlines. 
For  dermal  roofing  bones  of  the  skull,  the  external  surfaces  of  the  bones  correspond  to  their 


WILSON:  NEW  SPECIES  OF  THE  FISH  AMIA 


415 


ornamented  portions.  Osteological  terminology  follows  Boreske  (1974)  except  that,  following  Janot 
(1967),  the  term  ‘posttemporal’  is  used  in  place  of  ‘suprascapular’  (Table  1).  The  abbreviation 
UAVP  designates  that  specimens  are  deposited  in  the  Vertebrate  Paleontology  Collections, 
Department  of  Geology,  The  University  of  Alberta. 


table  1 . Abbreviations  used  in  the  figures 


a 

angular 

es 

extrascapular 

P 

pectoral  fin 

sc 

supracleithrum 

ao 

antorbital 

fr 

frontal 

pa 

parietal 

sm 

supramaxilla 

br 

branchiostegal 

h 

hyomandibular 

pop 

preopercle 

so 

subopercle 

c 

cleithrum 

io 

infraorbital 

ps 

parasphenoid 

sy 

symplectic 

ch 

ceratohyal 

la 

lachrymal 

pst 

parasphenoid  teeth 

V 

vomer 

cot 

coronoid  teeth 

m 

maxilla 

Pt 

(dermo)pterotic 

Vt 

vomerine  teeth 

d 

dentary 

me 

metacleithrum 

r 

rostral 

vtb 

vomerine  tooth  base 

ds 

dermosphenotic 

n 

nasal 

s 

posttemporal 

e 

endopterygoid 

op 

opercle 

sa 

surangular 

SYSTEMATIC  DESCRIPTION 
Class  OSTEICHTHYES 
Order  amiiformes 
Family  amiidae  Bonaparte,  1837 
Genus  amia  Linnaeus,  1766 

Type  species.  Amia  calva  Linnaeus,  1766. 


Amia  hesperia  sp.  nov. 

Text-figs.  2-6 

Diagnosis.  Deep-bodied  Amia  having  square  parietals;  long  frontals  with  shallow  orbital  excavation; 
large  nasals  without  anterior  notch;  large  fourth  infraorbital;  deep  maxilla,  mandible,  and  opercle; 
dentary  with  large  teeth;  lachrymal  with  posterior  notch;  parasphenoid  with  long  posterior  ramus 
and  short  tooth  patch;  vomers  with  short  tooth  patch;  vomers  and  coronoids  with  sharp,  conical 
teeth;  and  cleithrum  with  arms  at  obtuse  angle. 

Holotype.  UAVP  14758  (text-figs.  2-5  and  6a,  b),  an  almost  complete,  partially  articulated  fish  in  part  and 
counterpart  with  estimated  total  length  about  55  cm,  preserved  as  a mould  in  hard  siliceous  shale,  and 
collected  by  the  author’s  party  in  1977. 

Etymology.  The  specific  epithet  is  from  the  Latin  hesperius  meaning  ‘western’. 

Locality  and  age.  Ashnola  chert  site,  8-4  km  S of  Princeton,  British  Columbia  (U.T.M.  Grid  Reference 
10UFK783724),  from  the  Middle  Eocene  Allenby  Formation,  Princeton  Group. 

Non-type  material.  UAVP  13801,  a patch  of  scales  and  two  branchial  tooth  plates;  UAVP  13804  (text-fig.  6c), 
a right  dentary  and  maxilla;  UAVP  13805,  a right  extrascapular  (text-fig.  6f),  right  fourth  infraorbital,  and 
left  fifth  infraorbital;  UAVP  13806  (text-fig.  6d),  a branchial  toothplate;  and  UAVP  13812  (text-fig.  6e),  a 
right  opercle.  All  of  the  above  specimens  were  collected  at  the  type  locality  in  1977,  within  a few  centimetres 
of  the  holotype. 

Description.  Unless  otherwise  indicated,  the  following  description  is  based  on  the  holotype,  which  shows  the 
dermal  investing  bones  of  the  skull  in  the  part  (text-fig.  3)  and  many  ventral  and  internal  bones  of  the  skull 
in  the  counterpart  (text-fig.  4).  The  postcranial  skeleton  is  present  but  not  well  preserved  (text-fig.  2).  Non-type 
material  is  limited,  but  corroborates  conclusions  based  on  the  holotype. 

Parietals  are  approximately  square  (text-fig.  3),  and  frontals  are  relatively  long  and  narrow  (width  to  length 
ratio  0-41),  with  shallow  orbital  excavations  (‘orbital  concavity  ratio’  of  Boreske  1974  is  0117;  ‘dermosphenotic 
angle’  is  136°).  The  ‘parietal/frontal’  ratio  is  0-37.  Extrascapulars  are  decidedly  wider  laterally  than  medially 


416 


PALAEONTOLOGY,  VOLUME  25 


(text-figs.  3,  6f),  dermopterotics  substantially  overlap  the  frontals  laterally,  and  are  tapered  both  anteriorly 
and  posteriorly.  Nasals  are  large,  not  notched  anteriorly,  and  fit  the  anterior  outline  of  the  frontals  posteriorly. 
The  rostral  is  stout,  and  the  ornamented  area  of  the  antorbitals  is  small. 

Laterally,  the  lachrymal  has  a prominent  posterior  notch  which  fits  the  anterior  end  of  the  small  second 
infraorbital  (text-fig.  3).  The  large  fourth  infraorbital  is  deep  posteriorly,  where  it  has  an  angular  margin.  It 
tapers  to  the  orbit,  but  forms  less  of  the  orbital  rim  than  does  the  fifth  infraorbital.  The  latter  is  about  as 
deep  anteriorly  as  posteriorly,  and  is  considerably  smaller  than  the  fourth  infraorbital. 


text-fig.  2.  Amia  hesperia  sp.  nov.,  holotype,  part,  UAVP  14758a,  latex  peel  of  whole  fish. 


The  premaxilla  is  not  preserved.  The  maxilla  is  a stout,  deep  bone  with  a marginal  row  of  numerous,  very 
small  teeth  (text-figs.  3,  4).  The  posterodorsal  margin  of  the  maxilla  has  an  elongate  excavation  which  fits  a 
long,  deep  supramaxilla. 

The  mandible  is  deep  and  stout,  with  a steeply  inclined  posterior  border  formed  by  the  angular  and 
surangular  (text-figs.  3,  4),  a broad  coronoid  process,  and  a dentary  with  a long  row  of  large,  pointed  teeth 
(text-figs.  3,  4,  6b).  Proportions  and  shape  of  the  dentary  are  more  clearly  shown  in  UAVP  13804  (text-fig. 
6c),  where,  toward  the  front  of  the  bone,  the  ventral  margin  is  apparently  angled  ventrally.  This  is  interpreted 
as  a medial  deflection  of  the  mandibular  ramus  as  seen  most  noticeably  in  Kindleia  fragosa  (Boreske  1974, 
figs.  16b,  18).  The  coronoids  are  not  visible,  but  small,  sharp  teeth  located  internal  to  the  dentary  teeth,  where 
coronoid  teeth  would  be  expected,  are  visible  in  the  holotype  (text-fig.  6b). 

The  parasphenoid  (text-fig.  4)  is  partially  obscured  by  a branchiostegal,  a ceratohyal,  an  angular,  and  the 
vomers,  but  its  proportions  are  evident.  It  has  a relatively  long  posterior  ramus,  0-84  times  the  length  of  its 
anterior  ramus.  The  ascending  process  of  the  right  side  is  partially  covered  in  the  holotype  by  an  angular,  i 
but  is  approximately  perpendicular  to  the  long  axis  of  the  parasphenoid.  The  posterior  portion  of  the 
parasphenoid  tooth  patch  consists  of  many  tiny  denticles.  Where  it  is  not  obscured  by  the  ceratohyal  (text-fig. 
4),  the  anterior  ramus  of  the  parasphenoid  appears  devoid  of  denticles. 

Vomers  are  elongate  (text-fig.  4)  but  have  short-toothed  portions.  Most  teeth  are  broken,  but  unbroken 
teeth  together  with  some  broken  tooth  tips  indicate  that  the  vomerine  teeth  are  conical  and  sharply  pointed 
(text-fig.  6a).  The  endopterygoid  (text-figs.  3,  4)  is  plate-like  and  bears  numerous  tiny  denticles. 

The  opercular  bones  are  deep  relative  to  their  length.  The  opercle  in  the  holotype  has  a length  to  height 
ratio  of  0-90,  and  these  proportions  are  also  seen  in  UAVP  13812  (text-fig.  6e).  The  subopercle  has  an  elongate 
anterodorsal  ramus.  The  preopercle  is  not  well  preserved  (text-fig.  4),  but  seems  similar  to  those  of  other 
species  of  Amia. 

The  hyomandibular,  preserved  in  lateral  view  (text-fig.  3),  has  a posteroventrally  directed  opercular  process 
and  a moderately  developed  posterodorsal  notch.  The  symplectic  is  partially  visible  in  the  holotype  (text-fig. 


WILSON:  NEW  SPECIES  OF  THE  FISH  AMIA 


417 


text-fig.  3.  Amia  hesperia  sp.  nov.,  holotype,  part,  UAVP  14758a,  latex  peel  of  skull.  Abbreviations  Table  1. 

3).  The  ceratohyal  (text-fig.  4)  is  typical  of  the  genus,  and  bears  branchiostegals  which  vary  from  narrow  to 
broad.  Branchiostegal  ornamentation  is  extensive  and  posterior  margins  of  preserved  branchiostegals  are 
rounded.  Branchial  arches  are  not  preserved,  but  several  tiny  bones  bearing  numerous  sharp,  hollow,  conical, 
and  slightly  curved  teeth  (text-fig.  6d)  were  found  in  the  shales  close  to  the  holotype.  In  view  of  the  similarity 


418 


PALAEONTOLOGY,  VOLUME  25 


text-fig.  4.  Amia  hesperia  sp.  nov.,  holotype,  counterpart,  UAVP  14758b,  latex  peel  of  skull.  Abbreviations 
Table  1.  The  two  areas  outlined  in  white  are  enlarged  in  text-fig.  6a,  b. 

of  teeth  of  these  bones  to  teeth  of  the  holotype,  the  bones  are  interpreted  as  branchial  tooth  plates  of  the 
same  species. 

The  dermal  pectoral  girdle  is  well  preserved  in  the  holotype  (text-figs.  3-5).  The  cleithrum  is  like  that  of 
other  Amia  species,  but  the  ornamented  area  is  more  extensive  than  in  A.  calva.  As  well,  the  posterodorsal 


WILSON:  NEW  SPECIES  OF  THE  FISH  A MIA 


text-fig.  5.  Amia  hesperia  sp.  nov.,  holotype  UAVP  14758.  a.  Reconstruction  of  para- 
sphenoid  and  vomers  in  central  view.  b.  Left  hyomandibular,  lateral  view.  c.  Left  ceratohyal. 
D.  Left  dermal  pectoral  girdle  in  lateral  view.  Abbreviations  Table  1. 


ramus  of  the  cleithrum  forms  an  obtuse  angle  of  about  120°  with  the  an tero ventral  ramus,  as  in  K.  fragosa 
but  not  in  A.  calva  or  A.  uintaensis  where  the  rami  are  at  right  angles  to  each  other  (Boreske  1974,  fig.  21). 
The  plate-like  posterior  portion  of  the  cleithrum  is  broadly  rounded.  The  posttemporal  (text-fig.  3)  is  elongate, 
and  the  posterolateral  portion  is  ornamented.  The  pectoral  fin  has  about  sixteen  rays  and  originates  medial 
to  the  postero ventral  angle  of  the  cleithrum  (text-fig.  3).  Only  the  general  position  of  the  pelvic  fin  is  apparent 
in  the  holotype. 

The  postcranial  portion  of  the  axial  skeleton  is  poorly  preserved  (text-fig.  2),  but  a few  general  statements 
are  possible.  Anterior  trunk  vertebrae  are  about  four  times  as  broad  as  they  are  long,  and  bear  dorsal  neural 
facets  and  ventral  aortal  facets  as  in  other  Amia  species  (Boreske  1974,  fig.  11).  Ribs,  neural  and  haemal 
spines,  hypurals,  and  pterygiophores  are  all  slender,  rod-like  elements.  Posterior  caudal  vertebrae  are 
diplospondylous.  The  dorsal  fin  originates  approximately  above  the  seventeenth  trunk  centrum,  and  the  anal 
fin  consists  of  about  nine  rays. 

Scales  are  thinner  than  those  of  Cretaceous  and  Palaeocene  K.  fragosa  from  Alberta  (O’Brien  1969),  thicker 
than  scales  of  A.  calva,  but  similar  to  other  Eocene  Amia  scales  from  British  Columbia  (Wilson  1977a).  They 
are  mostly  rounded  apically  and  truncate  basally,  about  two-thirds  as  wide  as  long,  with  an  apical  (posterior) 
focus,  a thick,  horseshoe-shaped  rim  of  smooth  lamellar  bone  around  the  lateral  and  apical  margins,  and  a 
central  area  of  somewhat  thinner  bone  ornamented  on  the  internal  surface  by  small  bumps.  Externally  the 
circuli  or  ridges  radiate  to  the  margins  from  the  focus,  which  is  approximately  triangular  and  has  its  apex 
directed  basally  (anteriorly).  In  the  focal  area  the  circuli  form  a vermiculate  pattern. 


Reconstruction 

The  parasphenoid  and  vomers  of  A.  hesperia  are  reconstructed  in  text-fig.  5a.  Notable  features 
are  the  inferred  short  parasphenoid  tooth  patch,  the  ascending  processes  perpendicular  to  the 


CO 


420 


PALAEONTOLOGY,  VOLUME  25 


cot 


1 cm 


text-fig.  6.  Amia  hesperia  sp.  nov.  a.  Detail  of  anterior  portion  of  vomers  of  holotype,  UAVP  14758b, 
showing  pointed  vomerine  teeth,  b.  Detail  of  dentary  tooth  row  of  holotype,  UAVP  14758b,  showing  pointed 
coronoid  teeth,  c.  Right  dentary  and  maxilla,  lateral  view,  UAVP  13804.  D.  Branchial  tooth  plate,  UAVP 
13806.  e.  Right  opercle,  lateral  view,  UAVP  13812b.  f.  Right  extrascapular,  dorsal  view,  UAVP  13805c. 

Abbreviations  Table  1. 


long  axis  of  the  parasphenoid,  and  the  elongate  vomers  with  their  anterior  one-third  occupied  by 
sharp  teeth. 

Text-fig.  7 shows  the  skull,  reconstructed  in  dorsal  view,  compared  with  partial  reconstructions 
of  the  skulls  of  A.  calva,  K.  fragosa , A.  scutata,  and  A.  uintaensis.  Notable  are  the  elongate  frontals 
with  shallow  orbital  excavations,  relatively  short  parietals,  tapered  extrascapulars,  long 


WILSON:  NEW  SPECIES  OF  THE  FISH  AMIA 


421 


text-fig.  7.  Comparison  of  dorsal  skull  reconstructions  of  North  American  amiids.  a.  Amia  hesperia,  sp. 
nov.,  based  on  the  holotype,  UAVP  14758.  b.  A.  calva.  c.  Kindleia  fragosa.  d.  A.  scutata.  e.  A.  uintaensis.  b-e 
after  Boreske  (1974),  not  to  uniform  scale.  Abbreviations  Table  1. 


dermopterotics,  large  nasals  shown  interdigitating  with  the  frontals,  and  the  over-all  shape  of  the 
skull  roof,  which  is  elongate  but  only  slightly  narrower  at  the  rear  of  the  orbit  than  it  is  at  the 
rear  of  the  extrascapulars. 

The  skull  of  A.  hesperia  is  reconstructed  in  lateral  view  in  text-fig.  8.  In  this  view  the  skull 
appears  relatively  deep  and  large-jawed.  Features  that  contribute  to  the  impression  of  skull  depth 
are  the  large  lachrymal,  deep  opercle  and  subopercle,  deep  maxilla  and  supramaxilla,  deep  mandible, 
and  large  teeth  in  the  dentary.  The  obtusely  angled  cleithrum  is  additional  evidence  of  a deep 
body.  In  view  of  the  poor  preservation  of  the  postcranial  skeleton,  a reconstruction  of  the  entire 
fish  is  not  presented  here.  However,  the  evidence  of  the  dorsal  fin  apparently  originating  over  the 
seventeenth  trunk  centrum  suggests  that  A.  hesperia  was  long-bodied,  rather  than  short  and  stout 
as  was  K.  fragosa  (Boreske  1974,  fig.  21). 


DISCUSSION 

In  addition  to  the  one  Recent  and  three  fossil  species  of  Amia  from  North  America  that  were 
recognized  by  Boreske  (1974),  there  are  a number  of  nominal  European  and  Asian  fossil  species 
(Boreske  1974;  Janot  1967).  Fairly  complete  information  is  available  about  several  of  these,  including 
some  which  are  most  similar  to  the  North  American  K.  fragosa  (A.  kehreri  Andreae,  A.  valenciennesi 
Agassiz,  A.  munieri  Priem,  and  A.  russelli  Janot),  and  others  which  are  similar  to  A.  uintaensis  (A. 
robusta  Priem  and  possibly  A.  mongoliensis  Hussakof).  Janot  (1967),  Estes  and  Berberian  (1969), 


422 


PALAEONTOLOGY,  VOLUME  25 


text-fig.  8.  Comparison  of  lateral  skull  reconstructions  of  North  American  amiids.  A.  Amia  hesperia,  sp. 
nov.,  based  on  the  holotype,  UAVP  14758.  B.  A.  calva.  c.  Kindleia  fragosa.  D.  A.  scutata.  E.  A.  uintaensis.  b-e  j 
after  Boreske  (1974),  not  to  uniform  scale.  Abbreviations  Table  1. 


WILSON:  NEW  SPECIES  OF  THE  FISH  AMIA 


423 


Boreske  (1974),  and  Gaudant  (1980)  advocate  synonymizing  Kindleia  Jordan  with  Amia  Linnaeus. 
The  main  distinguishing  features  of  Kindleia  are  the  presence  of  styliform  teeth  on  the  coronoids, 
dermopalatines,  and  vomers;  differences  in  proportions  of  some  skull  bones;  and  a shorter  body 
with  less  separation  between  the  skull  and  the  origin  of  the  dorsal  fin  and  between  the  insertion 
of  the  dorsal  fin  and  the  caudal  fin,  and  about  twelve  fewer  trunk  centra  and  eight  fewer 
monospondylous  caudal  centra  (Boreske  1974).  These  authors  cited  intra-  and  interspecific 
variability  in  Amia,  and  found  the  above  differences  insufficient  grounds  for  separation  into  two 
genera,  especially  considering  that  Estes  and  Berberian  (1969)  believe  most  features  of  Kindleia  to 
be  primitive.  Gaudant  (1980),  however,  treats  Kindleia  as  a subgenus  of  Amia,  and  cites  several 
derived  features  shared  by  K.  fragosa  and  A.  kehreri  in  support  of  this  view.  Gaudant’s  list  of 
shared  derived  characters  is  very  similar  to  the  list  of  features  (above)  that  Estes  and  Berberian 
(1969)  interpreted  as  mostly  primitive.  For  example,  Gaudant  cites  a relatively  short,  deep  body; 
an  enlarged  fourth  infraorbital;  short  parietals;  and  styliform  coronoid,  palatal,  and  vomerine  teeth. 

I agree  with  Gaudant  that  these  characters  are  mostly  derived,  but  go  further  in  favouring  the 
recognition  of  Kindleia  as  a genus  distinct  from  Amia.  To  Gaudant’s  list  of  shared  derived  characters 
can  probably  be  added  the  relatively  short,  wide  frontal:  the  frontals  of  the  geologically  older  but 
closely  related  genera  Urocles  (Lange  1968)  and  Enneles  (Santos  1960)  are  long  and  narrow.  An 
enlarged  fourth  infraorbital  found  also  in  A.  hesperia  and  in  Enneles  (Santos  1960)  should  probably 
be  removed  from  the  list  of  derived  features  of  Kindleia. 

Retention  of  Kindleia  as  a separate  genus  is  also  useful  because  most  species  share  a number  of 
other  similarities,  some  or  all  of  which  might  be  primitive.  These  include  the  deep  orbital  notch 
in  the  frontal;  the  short  vomers;  the  small  supramaxilla;  the  narrow  maxilla  and  mandible;  the 
short,  truncated  gular  plate;  and  fewer  than  seventy-five  vertebrae. 

Boreske  (1974)  suggested  that  molluscs  were  a more  important  part  of  the  diet  for  Kindleia  than 
for  Amia.  At  some  localities  in  the  Palaeocene  Paskapoo  Formation  of  Alberta,  Kindleia  remains 
are  found  with  finely  crushed  mollusc  shells  concentrated  in  patches  on  bedding  planes.  To  my 
knowledge  gut  contents  have  not  been  observed  in  Kindleia  specimens,  but  the  explanation  of  the 
styliform  teeth  as  an  adaptation  to  molluscivorous  habits  is  a reasonable  one.  Perhaps  the  fish  did 
not  completely  swallow  the  shells  with  their  contents,  or  perhaps  the  patches  of  broken  shells 
represent  regurgitated  gastric  residues. 

The  new  species  described  here  clearly  belongs  with  Amia.  A.  hesperia  has  none  of  the  Kindleia 
specializations  listed  above.  It  shares  with  some  or  all  of  the  other  well-known  species  of  Amia 
derived  features  such  as  a shallow  orbital  excavation  in  the  frontal,  a deep  maxilla  with  elongate 
supramaxilla,  the  lack  of  an  anterior  notch  in  the  nasal,  the  presence  of  a posterior  notch  in  the 
lachrymal,  and  a tooth  patch  restricted  to  the  front  end  of  the  vomer. 

A.  hesperia  differs  from  valid  North  American  species  of  Amia  (Boreske  1974)  in  the  following 
features.  Compared  with  A.  uintaensis  it  has  slightly  shorter  frontals,  shorter  parietals,  unnotched 
nasals,  interdigitating  nasals  and  frontals,  posteriorly  notched  lachrymals,  more  angular  fourth 
infraorbitals,  a relatively  longer  posterior  parasphenoid  ramus,  parasphenoid  ascending  processes 
more  nearly  at  right  angles  to  the  long  axis,  shorter  parasphenoid  and  vomerine  tooth  patches, 
and  more  obtusely  angled  cleithra.  Compared  with  A.  scutata  and  A.  calva  it  has  longer  frontals, 
shorter  parietals,  interdigitating  nasals  and  frontals,  a shorter  parasphenoid  tooth  patch,  more 
rectangular  opercles,  and  more  obtusely  angled  cleithra.  It  differs  further  from  A.  calva  in  having 
longer  pterotics,  larger  fourth  infraorbitals,  a relatively  shorter  posterior  parasphenoid  ramus, 
ascending  processes  more  nearly  at  right  angles  to  the  long  axis,  and  more  extensive  ornament  on 
the  cleithra. 

A.  robusta,  a Palaeocene  to  Oligocene  European  species,  is  very  similar  to  A.  uintaensis.  Compared 
with  A.  hesperia  it  has  more  elongate  frontals,  more  rectangular  parietals,  a longer  parasphenoid 
tooth  patch,  and  more  right-angled  cleithra,  judging  by  bones  illustrated  in  Janot  (1967). 

Piscivorous  habits  for  A.  hesperia  are  implied  by  the  large  mouth  and  the  large,  sharp  teeth. 
This  supports  previous  conclusions  (Wilson  1980)  based  on  association  of  amiid  scales  with 
coprolites  containing  fish  bones  at  many  Eocene  localities  in  British  Columbia  and  Washington 


424 


PALAEONTOLOGY,  VOLUME  25 


State.  I would  predict  that  the  amiid(s)  occurring  at  these  other  localities  will  prove  to  be  closely 
related  to  or  conspecific  with  A . hesperia.  The  geographic  distribution  of  Eocene  amiid  species  in 
North  America,  with  A.  hesperia  in  the  extreme  west  and  A.  uintaensis  and  K.  fragosa  in  the 
mid-west,  is  suggestive  of  geographic  ranges  separated  by  a north-south  barrier  such  as  a continental 
divide  similar  to  the  one  separating  western  and  eastern  species  of  fishes  today. 

Acknowledgements.  I thank  Drs.  J.  F.  Basinger  and  W.  N.  Stewart  for  bringing  the  locality  to  my  attention, 
and  L.  A.  Lindoe  for  field  and  laboratory  assistance.  Dr.  B.  G.  Naylor  made  suggestions  that  resulted  in 
improvements  in  the  manuscript.  This  research  was  supported  by  Natural  Sciences  and  Engineering  Research 
Council  of  Canada  operating  grant  A9180. 


REFERENCES 

basinger,  j.  f.  1976.  Paleorosa  similkameenensis,  gen.  et  sp.  nov.,  permineralized  flowers  (Rosaceae)  from  the 
Eocene  of  British  Columbia.  Can.  J.  Bot.  54,  2293-2305. 

and  rothwell,  G.  w.  1977.  Anatomically  preserved  plants  from  the  Middle  Eocene  (Allenby  Formation)  of 

British  Columbia.  Ibid.  55,  1984-1990. 

boneham,  R.  F.  1968.  Palynology  of  three  Tertiary  coal  basins  in  south-central  British  Columbia.  Ph.D.  thesis, 
University  of  Michigan,  105  pp. 

bores ke,  j.  r.  1974.  A review  of  the  North  American  fossil  amiid  fishes.  Bull.  Mus.  Comp.  Zool.  Harv.  146,  1-87. 
CA vender,  t.  m.  1968.  Freshwater  fish  remains  from  the  Clarno  Formation  Ochoco  Mountains  of  north-central 
Oregon.  The  Ore  Bin,  30,  125-141. 

estes,  R.  and  berberian,  p.  1969.  Amia  (=  Kindleia)  fragosa  (Jordan),  a Cretaceous  amiid  fish,  with  notes  on 
related  European  forms.  Breviora,  329,  1-13. 

gaudant,  j.  1980.  Sur  Amia  kehreri  Andreae  (poisson  Amiidae  du  Lutetien  de  Messel,  Allemagne)  et  sa 
signification  paleogeographique.  C.R.  Acad.  Sc.  Paris,  290D,  1107-1110. 
hills,  l.  v.  and  baadsgaard,  h.  1967.  Potassium-argon  dating  of  some  Lower  Tertiary  strata  in  British 
Columbia.  Bull.  Can.  Petrol.  Geol.  15,  138-149. 

janot,  c.  1967.  A propos  des  amiides  actuels  et  fossiles.  Colloques  Int.  Centr.  Nat.  Rech.  Scient.  Paris,  163, 
139-153. 

lange,  s.  p.  1968.  Zur  Morphologie  und  Taxonomie  der  Fischgattung  Urocles  aus  Jura  und  Kreide  Europas. 
Palaeontographica,  ser.  A,  131a,  1-78. 

miller,  c.  n.  1973.  Silicified  cones  and  vegetative  remains  of  Pinus  from  the  Eocene  of  British  Columbia.  Contr. 
Mus.  Paleontol.  Univ.  Mich.  24,  101-118. 

o’brien,  d.  e.  1969.  Osteology  of  Kindleia  fragosa  Jordan  (Holostei:  Amiidae),  from  the  Edmonton  Formation 
(Maestrichtian)  of  Alberta.  M.Sc.  thesis.  University  of  Alberta,  118  pp. 
robison,  c.  r.  and  person,  c.  p.  1973.  A silicified  semiaquatic  dicotyledon  from  the  Eocene  Allenby  Formation  of 
British  Columbia.  Can.  J.  Bot.  51,  1373-1377. 

russell,  l.  s.  1935.  A Middle  Eocene  mammal  from  British  Columbia.  Am.  J.  Sci.  29,  54-55. 
scott,  w.  b.  and  crossman,  e.  j.  1973.  Freshwater  fishes  of  Canada.  Fish  Res.  Bd.  Can.,  Bull.  184. 
santos,  r.  da  s.  1960.  A posujao  sistematica  de  Enneles  audax  Jordan  e Branner  da  Chapada  do  Araripe,  Brasil. 
Monografias  Div.  Geol.  Miner.  Bras.  17,  1-25. 

wilson,  m.  v.  h.  1977a.  Middle  Eocene  freshwater  fishes  from  British  Columbia.  Life  Sci.  Contr.,  Roy.  Ont.  Mus. 
113,  1-61. 

— 1977ft.  New  records  of  insect  families  from  the  freshwater  Middle  Eocene  of  British  Columbia.  Can.  J. 
Earth  Sci.  14,  1139-1155. 

— 1978a.  Eohiodon  woodruffi  n.sp.  (Teleostei,  Hiodontidae),  from  the  Middle  Eocene  Klondike  Mountain 
Formation  near  Republic,  Washington.  Ibid.  15,  679-686. 

— 1978ft.  Paleogene  insect  faunas  of  western  North  America.  Quaest.  Entomol.  14,  13-34. 

— 1979.  A second  species  of  Libotonius  (Pisces:  Percopsidae)  from  the  Eocene  of  Washington  State.  Copeia, 
1979,  400-405. 

— 1980.  Eocene  lake  environments:  depth  and  distance-from-shore  variation  in  fish,  insect,  and  plant 
assemblages.  Palaeogeogr.,  Palaeoclimat.,  Palaeoecol.  32,  21-44. 

MARK  V.  H.  WILSON 
Department  of  Zoology 
The  University  of  Alberta 
Edmonton,  Alberta  T6G  2E9  Canada 


Typescript  received  19  February  1981 


A FUSED  CLUSTER  OF  CONIFORM  CONODONT 
ELEMENTS  FROM  THE  LATE  ORDOVICIAN 
OF  WASHINGTON  LAND,  WESTERN 
NORTH  GREENLAND 

by  R.  J.  ALDRIDGE 


Abstract.  A fused  conodont  cluster,  comprising  six  distacodontiform  elements  and  one  oistodontiform 
element,  from  upper  Ordovician  limestones  of  the  Aleqatsiaq  Fjord  Formation  of  Washington  Land,  north 
Greenland,  is  described.  A notable  microstructural  feature  of  all  the  elements  is  the  presence  of  oblique 
striations  along  the  anterior  margins  of  the  lateral  faces.  The  distacodontiform  elements  are  similar  to  elements 
included  in  Acodusl  mutatus  (Branson  and  Mehl)  and  Dapsilodus  obliquicostatus  (Branson  and  Mehl),  but 
the  apparatus  structure  appears  to  be  different  and  the  cluster  is  assigned  to  Besselodus  arcticus  gen.  et  sp.  nov. 

Conodont  elements  are  normally  found  as  isolated,  discrete  specimens,  and  until  the  mid  1960s 
conodont  taxonomy  and  nomenclature  were  almost  entirely  based  on  the  morphology  of  single 
elements.  A multielement  concept  of  conodont  species  was  first  proposed  over  a hundred  years 
ago,  by  Hinde  (1879),  and  the  discovery  of  ‘natural  assemblages’  of  conodonts  on  shale  bedding 
surfaces  by  Schmidt  (1934)  and  Scott  (1934)  provided  direct  evidence  that  each  animal  possessed 
a skeletal  apparatus  consisting  of  different  element  types.  Subsequently,  several  hundred  similar 
associations,  each  representing  the  skeletal  remains  of  an  individual  animal,  have  been  found, 
mostly  in  Carboniferous  shales  (e.g.  Scott  1942;  DuBois  1943;  Rhodes  1952).  Descriptions  of  many 
of  these  naturally  occurring  apparatuses  have  yet  to  be  published.  Additional  evidence  of  apparatus 
structures  is  provided  by  fused  conodont  clusters  recovered  from  acid-insoluble  residues  (e.g. 
Rexroad  and  Nicoll  1964;  Austin  and  Rhodes  1969;  Pollock  1969;  Ramovs  1977,  1978),  most  of 
which  represent  complete  or  partial  apparatuses. 

Once  the  multielement  structure  of  conodont  apparatuses  was  appreciated  it  was  only  a matter 
of  time  before  workers  began  to  reconstruct  apparatuses  from  collections  of  isolated  specimens. 
Walliser  (1964),  for  example,  suggested  reconstructions  of  nine  Silurian  conodont  apparatuses 
(informally  named  ‘Conodonten-Apparat’  A-J),  although  no  naturally  occurring  apparatuses  or 
clusters  were  at  that  time  known  from  the  Silurian.  A transition  of  emphasis  in  conodont  taxonomy 
from  a single  element  to  a multielement  basis  has  followed  as  more  and  more  reconstructions, 
based  on  statistical,  distributional,  and  morphological  evidence,  have  been  published.  Huddle  (1972) 
has  documented  the  development  of  the  early  phases  of  this  transition.  Of  importance  in  this 
taxonomic  revolution  was  the  realization  that  apparatuses  conform  to  a limited  number  of  basic 
plans  (Klapper  and  Philip  1971 ; Barnes,  Kennedy,  McCracken,  Nowlan,  and  Tarrant  1979).  Hence, 
the  structures  exhibited  by  naturally  occurring  apparatuses  and  clusters,  together  with  those  of 
well-established  reconstructions,  can  serve  as  models  in  the  analysis  of  new  collections. 

Naturally  occurring  apparatuses  and  clusters  are  rare,  and  it  is  probable  that  the  apparatuses 
of  most  species  will  be  recognized  entirely  from  collections  of  isolated  elements.  However,  the 
direct  evidence  furnished  by  natural  associations  is  of  paramount  importance  to  multielement 
taxonomy;  not  only  do  they  provide  templates  for  reconstructions,  but  new  finds  allow  testing 
and  evaluation  of  current  hypotheses.  Additionally,  it  is  only  from  these  associations  that  we  can 
confidently  assess  the  relative  numbers  and  dispositions  of  the  component  elements. 

The  only  natural  associations  of  late  Ordovician  conodonts  recorded  to  date  are  three  clusters 

[Palaeontology,  Vol.  25,  Part  2,  1982,  pp.  425-430,  pi.  44.| 


426 


PALAEONTOLOGY,  VOLUME  25 


of  Belodina  compressa  (Branson  and  Mehl)  from  Canada  (Barnes  1967;  Nowlan  1979).  Another 
cluster,  also  of  coniform  elements,  has  recently  been  isolated  from  late  Ordovician  limestones  of 
northern  Greenland.  Although  the  material  is  limited,  the  apparatus  differs  in  structure  from  those 
reconstructed  for  similar  Ordovician  and  Silurian  elements,  and  the  cluster  is  assigned  to  a new 
taxon,  Besselodus  arcticus. 


SAMPLE  LOCALITY  AND  HORIZON 

The  cluster  was  recovered  from  Geological  Survey  of  Greenland  sample  no.  GGU  242821,  collected  by  Dr. 
J.  M.  Hurst  from  the  Aleqatsiaq  Fjord  Formation  of  Aleqatsiaq  Fjord,  Washington  Land,  western  North 
Greenland  (see  Peel  and  Hurst  1980;  Hurst  1980).  The  sample  of  fissile,  argillaceous  calcilutite  is  from  132  m 
above  the  base  of  the  formation  and  approximately  140  m below  the  top.  The  conodont  fauna  includes 
Amorphognathus  aff.  ordovicicus  Branson  and  Mehl,  which  is  also  present  in  samples  GGU  242820  and  GGU 
242822,  from  20  m below  and  above  the  horizon.  A late  Ordovician,  Ashgill,  age  is  indicated. 

THE  CONODONT  FAUNA 

The  730  gm  sample  available  for  processing  yielded  only  a small  number  of  conodont  specimens.  Some  of 
these  are  fragmentary,  but  the  following  are  identifiable: 


Amorphognathus  aff.  ordovicicus  (Branson  and  Mehl) 

Pa  element  1 

Sa  element  2 

Belodina  compressa  (Branson  and  Mehl) 

compressed  rastrate  element  1 

eobelodiniform  element  1? 

Besselodus  arcticus  gen.  et  sp.  nov. 

in  cluster:  distacodontiform  element  6 

oistodontiform  element  1 

isolated:  distacodontiform  elements  8? 

oistodontiform  elements  5? 

Coelocerodontus  trigonius  Ethington  1 

Ouludusl  sp.  Sc  element  4 

Pander odus  sp.  2 

Pseudobelodindl  sp.  rastrate  element  1 

Pseudooneotodus  sp.  1 


Isolated  specimens  similar  to  those  of  Besselodus  arcticus  occur  sporadically  and  in  small  numbers  throughout 
the  lower  and  middle  Aleqatsiaq  Fjord  Formation.  Small  acodontiform  elements  are  occasionally  found  in 
the  same  samples,  but  it  is  not  at  present  possible  to  ascertain  whether  these  specimens  belong  in  Besselodus 
or  represent  another  apparatus. 


SYSTEMATIC  DESCRIPTION 
Genus  besselodus  gen.  nov. 

Type  species.  Besselodus  arcticus  sp.  nov. ; from  the  Aleqatsiaq  Fjord  Formation  of  Washington  Land. 

Diagnosis.  The  apparatus  contains  at  least  two  members,  distacodontiform  and  oistodontiform. 
All  known  elements  are  laterally  compressed  with  sharp  anterior  and  posterior  edges. 


EXPLANATION  OF  PLATE  44 

Figs.  1-8.  Besselodus  arcticus  gen.  et  sp.  nov.  1-2,  lateral  views  of  sub-cluster  ‘a’,  partial  holotype,  MGUH 
15071,  x 250.  3-4,  lateral  views  of  sub-cluster  ‘b’,  partial  holotype,  MGUH  15072,  x 250.  5-6,  lateral 

views  of  isolated  specimen,  MGUH  15073,  x 250.  7,  detail  of  partial  holotype  MGUH  15071,  x 660.  8, 

detail  of  partial  holotype  MGUH  15072,  x 660. 


PLATE  44 


ALDRIDGE,  fused  conodont  cluster 


428 


PALAEONTOLOGY,  VOLUME  25 


Discussion.  Distacodontiform  ( = acontiodontiform)  elements  similar  to  those  of  Besselodus  are 
included  in  the  reconstructions  of  the  Silurian  genus  Dapsilodus  (Cooper  1976,  p.  211)  and  of  the 
Ordovician  species  Acodusl  mutatus  (Branson  and  Mehl).  The  generic  assignment  of  A.?  mutatus 
was  discussed  fully  by  Lofgren  (1978,  pp.  43-45),  who  pointed  out  that  the  apparatuses  of  the 
type  species  of  Acodus,  Distacodus,  and  Acontiodus  are  all  unknown. 

Dapsilodus  has  an  apparatus  of  distacodontiform,  modified  distacodontiform  and  acodontiform 
elements  (Barrick  1977,  p.  50);  no  oistodontiform  element  is  included,  nor  are  there  any 
oistodontiform  or  modified  oistodontiform  elements  known  from  the  Silurian  that  might  be 
considered  as  homologous  to  the  oistodontiform  of  Besselodus.  A A mutatus  also  has  an  apparatus 
that  includes  distacodontiform  and  acodontiform  elements  (Bergstrom  and  Sweet  1966,  pp. 
303-305),  although  there  have  been  suggestions  that  an  oistodontiform  element  should  also  be 
included  here  (Barnes  and  Poplawski  1973,  p.  779;  Sweet,  in  Cooper  1976,  p.  211).  This  has  been 
contested  by  Lofgren  (1978,  pp.  45,  57),  who  found  no  distributional  relationship  between 
oistodontiforms  and  elements  of  A.?  mutatus  in  her  early  Ordovician  samples  from  northern  Sweden. 
The  oistodontiform  included  by  Barnes  and  Poplawski  (1973)  is  Oistodus  venustus  Stauffer,  which 
is  not  closely  similar  to  the  element  in  the  cluster  of  Besselodus. 

There  is  currently  little  evidence  for  the  inclusion  of  an  acodontiform  element  in  Besselodus. 
None  occurs  in  the  cluster,  nor  are  there  any  isolated  specimens  in  the  remainder  of  the,  admittedly 
small,  sample  from  the  same  horizon. 

Besselodus  arcticus  sp.  nov. 

Plate  44,  figs.  1-8 

Diagnosis.  The  apparatus  contains  at  least  two  members,  distacodontiform  and  oistodontiform. 
All  known  elements  are  laterally  compressed  with  sharp  anterior  and  posterior  edges;  at  the  anterior 
margins,  the  lateral  faces  of  all  elements  display  prominent  oblique  striations. 

Description.  The  material  to  hand  comprises  a cluster  and  a few  isolated  specimens,  all  from  the  same  sample. 
The  cluster  consists  of  very  small,  delicate  coniform  elements,  all  of  which  are  broken  at  the  tips.  The  elements 
are  all  very  thin  and  translucent  and  white  matter  cannot  be  distinguished.  Unfortunately,  the  cluster  fell 
apart  into  two  sub-clusters  on  initial  picking,  and  subsequent  handling  has  been  kept  to  a minimum. 

Sub-cluster  ‘a’  (PI.  44,  figs.  1,  2,  7)  is  0-25  mm  in  maximum  dimension  and  consists  of  four  distacodontiform 
elements  in  subparallel  orientation.  Each  component  measures  about  0-2  mm  from  base  to  broken  tip  and  is 
laterally  compressed  with  sharp  anterior  and  posterior  edges.  The  basal  outline  is  a highly  compressed  ellipse. 
Each  lateral  face  bears  a longitudinal  costa,  slightly  to  the  posterior  of  the  mid-line;  on  those  specimens 
where  the  basal  portion  is  visible,  the  costae  terminate  sharply  a short  distance  from  the  basal  margin.  The 
posterior  edge  of  each  element  is  gently  curved;  the  anterior  edge  is  more  sharply  curved  near  the  basal 
margin,  producing  a slight  geniculation.  As  far  as  can  be  ascertained  all  the  elements  are  bilaterally  symmetrical, 
or  very  nearly  so,  and  there  is  no  apparent  gradation  in  size  or  curvature.  Each  element  displays  well-developed 
fine,  parallel  striae  on  the  lateral  faces  at  the  anterior  margin;  at  the  broken  tips  these  striae  are  at  an  angle 
of  less  than  10°  to  the  anterior  edge,  towards  the  base  this  angle  steadily  increases  to  greater  than  20°.  The 
striae  fade  towards  the  base  and  terminate  at  the  point  of  geniculation  (PI.  44,  fig.  7). 

Eight  isolated  distacodontiform  specimens  from  the  same  sample  show  similar  features,  with  some  variation 
in  curvature  apparent.  Preservation  is  variable;  the  best-preserved  is  illustrated  in  PI.  44,  figs.  5,  6.  This 
specimen  differs  from  those  in  the  cluster  in  lacking  the  geniculation  of  the  anterior  edge,  producing  a more 
triangular  outline  for  the  basal  portion  of  the  unit. 

Sub-cluster  ‘b’  (PI.  44,  figs.  3,  4,  8)  is  0-24  mm  in  maximum  dimension.  Two  distacodontiform  elements 
comparable  with  those  in  sub-cluster  ‘a’  are  fused  in  subparallel  orientation;  a third,  broken  and  cracked, 
oistodontiform  element  is  orientated  so  that  its  cusp  is  parallel  to  those  of  the  distacodontiform  elements. 
The  strong  geniculation  of  the  oistodontiform  results  in  the  basal  margin  of  the  unit  lying  at  an  angle  of  30° 
to  the  basal  margins  of  the  distacodontiforms.  The  visible  lateral  face  of  the  oistodontiform  shows  a longitudinal 
costa  on  the  posterior  portion  of  the  cusp;  oblique  striae  at  the  anterior  margin  of  the  cusp  are  also  apparent, 
ranging  in  angle  from  12°  at  the  broken  tip  to  20°  near  the  geniculation,  where  they  terminate.  These  striae 
compare  closely  with  those  displayed  by  the  distacodontiform  elements. 

There  are  five  small,  isolated  oistodontiform  specimens  in  the  sample,  but  those  examined  under  the  scanning 


ALDRIDGE:  FUSED  CONODONT  CLUSTER 


429 


electron  microscope  do  not  exhibit  clear  oblique  marginal  striations;  they  may  or  may  not  be  referable  to 
the  same  species  as  the  cluster. 

Holotype.  MGUH  15071  and  MGUH  15072,  separated  portions  of  a single  cluster  from  sample  GGU  242821; 
deposited  in  the  type  collection  of  the  Geologisk  Museum,  Copenhagen. 

Discussion.  The  cluster  probably  represents  a partial,  rather  than  a complete,  apparatus.  The 
presence  of  a single  oistodontiform  may  indicate  that  it  is  a partial  or  complete  half-apparatus. 
In  addition  to  the  general  similarity  of  the  morphology  of  the  distacodontiform  elements,  the 
oblique  striae  on  the  anterior  margins  of  the  elements  may  indicate  a relationship  to  Dapsilodus. 
Similar  striae  are  apparent  on  specimens  of  D.  obliquicostatus  figured  by  Serpagli  (1970,  pi.  23, 
figs.  1-10,  pi.  24,  figs.  1-6),  Cooper  (1976,  pi.  2,  figs.  11-12,  18-20),  and  Barrick  (1977,  pi.  2,  figs. 
6,  10).  Serpagli  (1970)  noted  the  absence  of  similar  striations  on  specimens  referred  to  the  older, 
possibly  ancestral,  species  Acodus ? mutatus.  In  the  material  referred  to  that  species  by  Lofgren 
(1978,  p.  44,  pi.  2,  figs.  9-21)  longitudinal  striations  are  developed  and  a single  specimen  (pi.  2, 
fig.  11a,  b)  also  possesses  anterior  striae  at  an  angle  of  5-10°  to  the  anterior  margin.  The  presence 
of  anterior  striae  may  not  be  an  important  character  in  determining  relationships,  as  they  are 
displayed  by  other  specimens,  and  are  apparent  in  the  clusters  of  Belodina  compressa  figured  by 
Nowlan  (1979,  especially  fig.  35.2). 

Acknowledgements.  This  contribution  arises  from  a major  co-operative  project  with  the  Geological  Survey  of 
Greenland;  I am  grateful  to  them  for  providing  the  sample,  especially  to  Drs.  J.  M.  Hurst  and  J.  S.  Peel  for 
collecting  the  material  and  arranging  the  loan.  Mr.  A.  Swift  processed  the  sample  in  the  laboratory  and  first 
recognized  the  cluster.  I also  thank  Mr.  L.  Green  and  Mr.  D.  J.  Jones  for  photographic  assistance.  Dr.  Anita 
Lofgren  (University  of  Lund)  kindly  read  and  commented  on  the  typescript.  This  paper  is  published  with 
the  permission  of  the  Director  of  the  Geological  Survey  of  Greenland. 


REFERENCES 

Austin,  R.  l.  and  Rhodes,  f.  h.  t.  1969.  A conodont  assemblage  from  the  Carboniferous  of  the  Avon  Gorge, 
Bristol.  Palaeontology,  12,  400-405. 

barnes,  c.  R.  1967.  A questionable  natural  conodont  assemblage  from  Middle  Ordovician  limestone,  Ottawa, 
Canada.  J.  Paleont.  41,  1557-1560. 

— Kennedy,  D.  J.,  mccracken,  a.  d.,  nowlan,  G.  s.  and  tarrant,  g.  a.  1979.  The  structure  and  evolution 
of  Ordovician  conodont  apparatuses.  Lethaia,  12,  125-151. 

— and  poplawski,  m.  l.  s.  1973.  Lower  and  Middle  Ordovician  conodonts  from  the  Mystic  Formation, 
Quebec,  Canada.  J.  Paleont.  47,  760-790,  pis.  1-5. 

barrick,  j.  e.  1977.  Multielement  simple-cone  conodonts  from  the  Clarita  Formation  (Silurian),  Arbuckle 
Mountains,  Oklahoma.  Geol.  and  Palaeontol.  11,  47-68,  pis.  1-3. 

Bergstrom,  s.  M.  and  sweet,  w.  c.  1966.  Conodonts  from  the  Lexington  Limestone  (Middle  Ordovician)  of 
Kentucky  and  its  lateral  equivalents  in  Ohio  and  Indiana.  Bull.  Am.  Paleont.  50,  269-441,  pis.  28-35. 
cooper,  b.  j.  1976.  Multielement  conodonts  from  the  St.  Clair  Limestone  (Silurian)  of  Southern  Illinois. 
J.  Paleont.  50,  205-217,  pis.  1-2. 

dubois,  E.  p.  1943.  Evidence  on  the  nature  of  conodonts.  Ibid.  17,  155-159,  pi.  25. 

hinde,  G.  J.  1879.  On  conodonts  from  the  Chazy  and  Cincinnati  Group  of  the  Cambro-Silurian,  and  from  the 
Hamilton  and  Genesee-Shale  divisions  of  the  Devonian,  in  Canada  and  the  United  States.  Q.  Jl  geol.  Soc. 
Lond.  35,  351-369,  pis.  15-17. 

huddle,  j.  w.  1972.  Historical  introduction  to  the  problem  of  conodont  taxonomy.  Geol.  and  Palaeontol.  SB  1, 
3-16,  pi.  1. 

hurst,  j.  m.  1980.  Silurian  stratigraphy  and  facies  distribution  in  Washington  Land  and  Western  Hall  Land, 
North  Greenland.  Bull.  Gronlands  geol.  Unders.  138,  1-95. 
klapper,  g.  and  philip,  G.  m.  1971.  Devonian  conodont  apparatuses  and  their  vicarious  skeletal  elements. 
Lethaia,  4,  429-452. 

lofgren,  a.  1978.  Arenigian  and  Llanvirnian  conodonts  from  Jamtland,  northern  Sweden.  Foss,  and  Strata, 
13,  1-129,  pis.  1-16. 

nowlan,  G.  s.  1979.  Fused  clusters  of  the  conodont  genus  Belodina  Ethington  from  the  Thumb  Mountain 


430 


PALAEONTOLOGY,  VOLUME  25 


Formation  (Ordovician),  Ellesmere  Island,  District  of  Franklin.  Curr.  Res.  Part  A,  Geol.  Surv.  Can.,  Pap. 
79-1A,  213-218. 

peel,  j.  s.  and  hurst,  J.  M.  1980.  Late  Ordovician  and  early  Silurian  stratigraphy  of  Washington  Land,  western 
North  Greenland.  Rapp.  Gronlands  geol.  Unders.  100,  18-24. 
pollock,  c.  a.  1969.  Fused  Silurian  conodont  clusters  from  Indiana.  J.  Paleont.  43,  929-935,  pis.  110-112. 
ramovs,  a.  1 977.  Skelettapparat  von  Pseudo furnishius  murcianus  (Conodontophorida)  im  Mitteltrias  Sloweniens 
(NW  Jugoslawien).  Neues  Jb.  Geol.  Palaont.  Abh.  153,  361-399. 

— 1978.  Mitteltriassische  conodonten-clusters  in  Slowenien,  NW  Jugoslawien.  Palaont.  Z.  52,  129-137. 
rexroad,  c.  b.  and  nicoll,  R.  s.  1964.  A Silurian  conodont  with  tetanus?  J.  Paleont.  38,  771-773. 

Rhodes,  F.  h.  T.  1952.  A classification  of  Pennsylvanian  conodont  assemblages.  Ibid.  26,  886-901,  pis.  126-129. 
schmidt,  H.  1934.  Conodonten-Funde  in  urspriinglichen  Zusammenhang.  Palaont.  Z.  16,  76-85,  pi.  6. 
scott,  h.  w.  1934.  The  zoological  relationships  of  the  conodonts.  J.  Paleont.  8,  448-455,  pis.  58-59. 

— 1942.  Conodont  assemblages  from  the  Heath  formation,  Montana.  Ibid.  16,  293-300,  pis.  37-40. 
serpagli,  E.  1970.  Uppermost  Wenlockian-upper  Ludlovian  (Silurian)  conodonts  from  Western  Sardinia. 

Boll.  Soc.  Paleontol.  Ital.  9,  76-96,  pis.  21-24. 

walliser,  o.  h.  1964.  Conodonten  des  Silurs.  Abh.  hess.  Landesamt  Bodenforsch.  41,  1-106,  pis.  1-32. 


Typescript  received  27  November  1980 
Revised  typescript  received  26  January  1981 


R.  J.  ALDRIDGE 

Department  of  Geology 
University  of  Nottingham 
Nottingham  NG7  2RD 


A NEW  CALCAREOUS  GREEN  ALGA  FROM  THE 
MIDDLE  JURASSIC  OF  ENGLAND:  ITS 
RELATIONSHIPS  AND  EVOLUTIONARY 
POSITION 

by  GRAHAM  F.  ELLIOTT 


Abstract.  Leckhamptonella  llewellyae  gen.  et  sp.  nov.,  is  described  from  the  Middle  Jurassic  (Inferior  Oolite; 
Aalenian)  of  the  Cotswold  district,  England.  Although  known  only  from  fragmentary  material,  it  is  recognizable 
as  the  remains  of  a new  serial-segmented  member  of  the  Udoteaceae  (Chlorophyta).  It  is  compared  with  Boueina 
and  Arabicodium  (both  Mesozoic)  and  Halimeda  (Cretaceous-Recent),  and  shows  certain  affinities  with  the 
latter  especially.  The  evolution  of  serial-segmented  Udoteaceae  from  Lower  Palaeozoic  to  Recent  is  briefly 
reviewed. 

The  fossil  described  below  occurs  as  pieces  and  fragments  of  original  small  calcareous-walled  hollow 
ovoids,  near-circular  in  transverse  section  and  curved-elongate  in  vertical  section.  The  walls  show  a 
very  distinctive  structure  formed  by  irregular  but  characteristic  branching  pores,  with  possible 
vestiges  of  a structure  interior  to  the  wall.  The  material  studied  occurs  in  an  oomicrite  with 
subordinate  echinoderm,  molluscan,  brachiopod,  and  other  organic  fragments,  often  worn  and 
encrusted  and  presumed  current-swept  and  not  on  the  site  of  growth. 

The  new  fossil  does  not  show  the  optical  extinction  of  echinoderm-calcite  pieces,  nor  the  laminar 
structure  of  molluscs  or  brachiopods.  The  wall-structure  is  not  that  of  stromatoporoids  or  hydrozoa, 
and  has  not  the  spicular  mesh  of  any  known  sponge.  It  does,  however,  have  the  characteristics  of 
certain  green  calcareous  algae,  notably  the  Order  Dasycladales  and  the  family  Udoteaceae  of  the 
Order  Caulerpales.  The  resemblance  is  very  much  with  the  cortical  structure  of  the  ‘serial-segmented’ 
members  of  the  Udoteaceae  ( Halimeda , and  various  extinct  genera  discussed  below).  It  possesses  the 
marked  irregularity  in  detail  which  is  so  characteristic,  and  the  general  pore-plan  is  that  of  yet  another 
variant  of  the  branching  utricles  of  the  family.  It  does  not,  in  most  specimens,  show  clearly  the 
longitudinal  medullary  threads  appropriate  to  this  interpretation,  but  two  well-preserved  pieces  show 
traces  of  this  structure,  which  has  to  be  carefully  distinguished  from  peripheral  diagenesis  and 
staining  of  the  clasts.  The  absence  is  probably  due  to  the  rolled  state  of  the  fragments  and  the  usual 
originally  weak  medullary  calcification  in  the  Udoteaceae.  In  the  Permian  Tauridium,  normally  found 
fragmented,  very  much  more  abundant  material  than  is  available  with  the  present  fossil  still  leaves 
reconstruction  a difficult  task. 

If  comparison  is  made  with  the  Udoteacean  Ovulites  (Cretaceous,  Cenozoic),  the  external 
dimensions  and  shape  of  the  Cotswold  fossil  as  so  far  known  are  comparable  with  those  of  O. 
margaritula  (Lmk.)  Munier-Chalmas  (cf.  Massieux  1966).  In  these,  however,  and  other  Ovulites  spp., 
the  calcareous  wall  is  thinner,  the  pores  near-uniformly  straight,  thin,  and  radial,  and  no  trace 
survives  of  medullary  structures.  Ovulites  is  usually  regarded  as  remains  of  a plant  comparable  with 
the  living  Penicillus,  the  ‘Neptune’s  Shaving  Brush’,  where  the  thallus  is  of  different  morphology  to 
that  of  Halimeda.  Ovulites  and  Halimeda  both  appear  in  the  Cretaceous,  and  it  is  interesting  that  the 
new  Jurassic  fossil  brings  both  to  mind,  even  if  apparently  much  more  similar  to  the  latter  in 
structure. 

If  the  present  remains  are  envisaged  as  those  of  a dasycladalean,  they  are  anomalous  in  showing 
udoteacean  irregularity  and  not  a verticillate  arrangement.  The  thin  calcification  of  the  swollen  heads 
of  the  Jurassic  Petrascula  and  Coniporella  belongs  to  very  much  larger  individuals. 


IPalaeontology,  Vol.  25,  Part  2,  1982,  pp.  431-437,  pi.  45.| 


432 


PALAEONTOLOGY,  VOLUME  25 


It  would  thus  seem  that  the  fragments  can  be  interpreted  as  those  of  ovoid  serial-segments  of  a new 
udoteacean,  the  medullary  part  originally  weakly  calcified  and  now  largely  missing  in  the  worn 
fragments  available. 

SYSTEMATIC  PALAEONTOLOGY 

Division  chlorophyta  (Green  Algae) 

Order  caulerpales  Feldmann  1946 
Family  udoteaceae  Feldmann  1946 
Genus  leckhamptonella  gen.  nov. 

Udoteacean  segments  with  calcified  cortical  zone  showing  swollen  branching  utricles  each  dividing  into  several 
thinner  outer  parallel  utricles  which  again  divide  peripherally.  Type-species  Leckhamptonella  llewellyae  sp.  nov., 
dedicated  to  Dr.  Llewellya  Hillis-Colinvaux  in  recognition  of  her  extensive  studies  of  the  living  Halimeda. 

Leckhamptonella  llewellyae  sp.  nov. 

Plate  45,  figs.  1-6 

Description.  Leckhamptonella  with  presumed  ovoid  segments  of  observed  length  up  to  2-70  mm  and  estimated 
matching  diameter  of  approximately  1 -80  mm  (but  a section  of  2-40  diameter  may  indicate  larger  segments). 
Medullary  zone  mostly  missing  in  fossil  pieces:  thickness  of  presumed  medullary  filaments  where  preserved 
about  0 030  mm.  Thickness  of  calcified  cortical  zone  about  0-36  mm;  inner  half  of  this  occupied  by  a zone  of 
outwardly  directed  waisted  and  swollen  branching  utricles  of  0-030-0  090  mm  diameter.  These  each  divide  into 
several  straight,  thin,  near-parallel  outer  utricles  of  about  0-015  mm  diameter.  These  are  at  right  angles  to  the 
longitudinal  axis  of  the  segment,  and  divide  again  just  below  the  segment-surface,  to  short  adjacent  peripheral 
utricles  of  0-010  mm  or  less  diameter,  which  expand  up  to  0-020  mm  terminal  diameter,  almost  touching. 

Holotype.  The  specimen  figured  in  PI.  45,  fig.  1;  British  Museum  (Natural  History),  Dept.  Palaeontology 
registered  number  V.60703.  Middle  Jurassic,  Lower  Inferior  Oolite  (Aalenian  bradfordensis  subzone);  Upper 
Freestone  facies  of  Scotsquar  Hill  Limestone  (Mudge  1978):  Leckhampton,  Cheltenham,  Gloucestershire. 

Paratypes.  The  specimens  figured  in  PI.  45,  figs.  2-6;  registered  numbers  V.60704-60708  inch;  same  locality  and 
horizon. 

Other  material.  About  thirty-five  fragments  in  thin-sections:  same  locality  and  horizon. 


THE  EVOLUTION  OF  THE  ‘SERIAL-SEGMENTED’  UDOTEACEAE 

The  members  of  the  Udoteaceae,  to  which  Leckhamptonella  is  referred,  show  a structure  of  repeatedly 
branching  threads,  much  intertangled  and  interwoven.  In  Udotea  itself  the  whole  thallus  is  a single 


EXPLANATION  OF  PLATE  45 

Figs.  1-6.  Leckhamptonella  llewellyae  gen.  et  sp.  nov.  All  pieces  in  thin-section  from  the  Middle  Jurassic, 
Aalenian  bradfordensis  subzone,  Lower  Inferior  Oolite;  Leckhampton,  Gloucestershire,  England.  1 , portion 
of  vertical  section  of  calcareous  wall,  showing  presumed  medullary  thread  at  base,  swollen  branching  cortical 
utricles,  radial  utricles,  and  terminal  peripheral  utricles  (shown  a little  left  of  top  centre);  x 80;  Holotype, 
British  Museum  (Natural  History),  Department  of  Palaeontology,  registered  number  V.60703.  2,  portion  of 

transverse  section  of  large  unit,  with  external  dark  crust,  x 30;  V.60704.  3,  portion  of  vertical  section, 

showing  two  presumed  medullary  threads  (bottom  right),  swollen  utricles  and  radial  utricles,  x 80; 
V. 60705.  4,  tangential  subsurface  section,  showing  close-set  peripheral  utricles,  x 80;  V.60706.  5,  slightly 

oblique  vertical  section  of  whole  side  of  unit,  x 40;  V. 60707.  6,  tangential-longitudinal  section  of  wall 

showing  swollen  and  radial  utricles,  heavy  dark  crusting  externally,  x 40;  V.60708. 

Fig.  7.  Oblique  cut,  piece  of  Tauridium  sp.  for  comparison;  Upper  Permian,  Southern  Tunisia,  x 40;  V. 54052. 


PLATE  45 


ELLIOTT,  calcareous  green  alga 


434 


PALAEONTOLOGY,  VOLUME  25 


fan-shaped  structure,  but  another  common  growth-form,  of  which  Halimeda  is  the  modern 
representative,  is  shown  by  what  may  be  termed  ‘the  serial-segmented  udoteaceans’.  In  these,  known 
since  the  Lower  Palaeozoic,  the  plant  is  (or  is  presumed  to  have  been)  a clump  of  branching 
successions  of  numerous  individual  calcified  segments,  connected  by  uncalcified  threads  or  filaments. 
Each  segment  shows  internal  coarse  longitudinal  medullary  filaments  from  end  to  end,  and  an  outer 
cortical  zone  of  lateral,  branching,  and  often  swollen  utricles  which  terminate  in  a surface  layer  of 
closely  packed  peripheral  utricles.  The  medullary  filaments  give  rise  distally  during  growth  to  new 
segments  beyond  the  old,  the  connecting  filaments  being  either  separate  or  partially  or  completely 
fused  at  the  nodes,  so  uniting  the  whole  segmented  plant  as  a flexible  structure.  The  peripheral  utricles 
are  concerned  with  the  assimilatory  functions,  and  the  swollen  inner  cortical  utricles  with  eventual 
transference  of  their  content  to  deciduous  reproductive  outgrowths  (gametangia)  at  the  times  of 
sexual  reproduction  (Hillis-Colinvaux  1980).  These  last  are  only  really  known  in  the  living  Halimeda 
(one  fossil  record  in  Pfender  1940,  p.  245)  and  our  knowledge  is  incomplete. 

Calcification  of  the  segments  is  usually  heavy  cortically  (interutricular),  but  much  less  so  in  the 
medullary  zone.  In  the  fossils  the  segments  occur  dissociated,  mixed,  and  often  broken.  They  are 
preserved  because  of  the  calcification.  The  behaviour  of  the  medullary  filaments  at  the  nodes,  so 
important  in  classification  of  the  living  Halimeda  (Hillis-Colinvaux  1980)  is  thus  not  normally  known 
in  the  extinct  genera.  A solitary  example  in  the  collections  of  the  British  Museum  (Natural  History)  of 
an  undescribed  udoteacean  from  the  Upper  Permian  of  Tunisia,  reg.  no.  V. 54065,  shows  the  nodal 
filaments  parallel  and  in  contact  along  their  length,  but  apparently  not  fused  (if  this  is  the  correct 
interpretation  of  their  slightly  different  mineralizations).  The  variation  in  detail  and  degree  of 
structure  and  calcification  between  older  and  younger  segments  of  individual  plants,  between  those  of 
different  individuals  of  the  same  species,  and  between  those  typical  of  different  species,  as  known  in 
living  Halimeda  (Hillis  1959;  Hillis-Colinvaux  1980)  can  make  precise  specific  evaluation  of  the 
fragmentary  resorted  fossil  remains  difficult,  though  often  occurrences  seem  to  be  of  a single  species. 
Ideally,  one  should  have  a real  abundance  of  material  to  describe  (Conard  and  Rioult  1977),  but  this 
is  not  always  available.  The  state  of  geological  preservation  may  add  to  these  difficulties,  explaining 
such  records  as  Boueina/ Arabicodium,  Boueina/ Halimeda  (Bismuth,  Bonnefous,  and  Dufaure  1967). 

The  evolution  of  serial-segmented  udoteaceans,  as  preserved  fossil,  appears  to  have  consisted  of 
variation  and  different  combinations  of  the  basic  structures  outlined  above,  from  the  Lower 
Palaeozoic  onwards  (text-fig.  1).  Thus  Dimorphosiphon  (Ordovician)  shows  a coarse  thread-structure 
and  Palaeoporella  (Ordovician-Devonian)  a fine  one,  while  Maslovina  (Silurian)  has  the  dense  layer 
of  small  peripheral  utricles  typical  of  some  later  genera  (Obrhel  1968),  but  not  present  in  the  other 
two.  Text-fig.  1 shows  a selection  of  patterns  in  different  genera  from  Dimorphosiphon  to  Halimeda.  Is 
this  evolution  more  or  less  random,  or  does  it  show  some  progression  in  time  with  the  surviving 
Halimeda  as  the  most  advanced  as  well  as  the  latest  of  its  kind?  And  does  the  new  Jurassic 
Leckhamptonella  throw  any  light  on  this  problem? 

It  can  be  seen  from  text-fig.  1 that  the  central  medullary  filaments  can  be  thick  or  thin,  straight  or 
tangled  in  varying  degree  in  different  genera.  Similarly,  the  cortical  utricles,  while  often  branching, 
vary  much  in  spacing  or  crowding,  and  in  degree  of  swelling  between  genera.  The  examples  figured 
show  various  combinations,  typical  of  those  genera.  On  mechanical  grounds,  strong  straight 
medullary  filaments,  with  some  fusion  and  flexibility  at  the  nodes,  would  equip  such  a plant  to 
withstand  a moderate  degree  of  water-movement  and  enable  it  to  colonize  moderate-energy 
environments,  other  conditions  being  suitable.  In  addition,  swollen  cortical  utricles  and  a layer  of 
close-set  peripheral  utricles  would  provide  for  quick  segment-growth  and  eventually  for  rapid 
production  of  gametangia  at  times  of  sexual  reproduction.  This  combination,  familiar  in  Halimeda , is 
apparently  first  achieved  in  the  Permian  Tauridium,  though  with  different  proportions  in  the  utricles, 
but  these  were  extremely  fragile  plants  post-mortem,  almost  invariably  found  as  debris  or 
comminuted.  Apparently  calcification  was  thin  and  mostly  outer-cortical,  and  in  life  the  plant  was 
probably  confined  to  quiet  waters.  In  Halimeda,  however,  calcification  is  heavy,  and  its  success  as 
witnessed  by  abundance,  wide  distribution,  and  occurrence  in  a range  of  low  to  moderate  energy- 
environments  (Hillis-Colinvaux  1980)  is  well  known. 


ELLIOTT:  CALCAREOUS  GREEN  ALGA 


435 


text-fig.  1 . Diagrammatic  representations  of  filament  and  utricle  structure  in  various 
udoteacean  genera,  mostly  based  on  materials  in  the  collections  of  the  British  Museum 
(Natural  History):  (a)  Dimorphosiphon  (Ordovician);  ( b ) Aphroditicodium  (Permian; 
BM(NH)  Dept.  Palaeont.  reg.  no.  V.59461);  (c)  Tauridium  (Permian);  (d)  Arabicodium 
(Jurassic-Cretaceous);  (e)  Boueina  (Triassic-Cretaceous);  (/)  Halimeda  incrassata 
(Ellis)  Lamx  (Recent)  (Hillis  1 959);  (g)  Leckhamptonella  llewellyae  Elliott  (Jurassic); 
( h ) diagrammatic  growth-plan  of  serial-segmented  udoteacean.  See  also  comparisons 
for  some  other  Palaeozoic  genera  in  Obrhel  (1968,  fig.  1)  and  Guilbault  and  Mamet 
(1976,  fig.  2). 


436 


PALAEONTOLOGY,  VOLUME  25 


Leckhamptonella  llewellyae  shows  similarities  in  cortical  utricle-structure  to  typical  modern 
Halimeda  spp.  It  differs  in  that  the  third  layer  of  cortical  branches  are  straight,  thin,  and  parallel 
before  dividing  into  peripheral  utricles,  whereas  in  the  modern  Halimeda  spp.  swollen  branches  and 
branchlets  usually  continue  outwards  to  the  peripheral  utricles.  In  the  light  of  the  functional 
reasoning  above  this  difference  would  be  a primitive  character  is  Leckhamptonella.  The  general 
appearance  of  the  medullary  zone  is  not  known,  and  so  cannot  be  compared  with  those  of  other 
genera. 

Halimeda  is  itself  known  rarely  from  the  Lower  Cretaceous  (Dragastan  and  Bucur  1979;  possibly 
Wells  1944),  becomes  more  common  in  the  Upper  Cretaceous,  and  is  abundant  throughout  the 
Cenozoic  to  the  present  day.  It  has  been  considered  as  arising  by  hybridization  from  the  earlier 
Mesozoic  Boueina  and  Arabicodium  (Elliott  1965)  or  as  more  closely  related  to  Boueina  as  evidenced 
by  comparable  intrageneric  species  groupings  (Conard  and  Rioult  1977).  The  fragmentary  condition 
of  the  Middle  Jurassic  Leckhamptonella , as  described  above,  precludes  a detailed  comparison  of  these 
four  genera,  but  the  cortical  structure  of  Leckhamptonella  appears  closer  to  that  of  the  later  Halimeda 
than  to  those  of  the  earlier  genera. 

Arabicodium , Boueina , and  Halimeda  were  all  Tethyan  in  origin:  Boueina  appears  in  the  Upper 
Triassic  of  Central  Europe  and  of  Thailand  (Flrigel  1975;  Kemper,  Maronde,  and  Stoppel  1976)  and 
its  pan-tropical  distribution  in  the  Lower  Cretaceous  has  been  plotted  by  Elliott  (1981).  Halimeda 
may  have  appeared  in  the  Lower  Cretaceous  of  both  hemispheres  (Dragastan  and  Bucur  1979;  Wells 
1944)  and  was  certainly  widely  distributed  in  the  Upper  Cretaceous  (Elliott  1981).  Arabicodium  from 
the  Jurassic-Cretaceous  of  the  western  Tethys  (Mediterranean-Middle  East)  had  a Jurassic  straggler 
as  far  north  as  southern  England  (Elliott  1975)  in  the  same  area  where  Leckhamptonella  occurs. 

Acknowledgements.  My  thanks  are  due  to  Mr.  Mark  Crawley  and  Mr.  P.  V.  York,  both  British  Museum 
(Natural  History),  for  drawings  and  photography  respectively. 


REFERENCES 

bismuth,  h.,  bonnefous,  j.  and  dufaure,  ph.  1967.  Mesozoic  microfacies  of  Tunisia.  In  Guidebk  Ann.  Fid  Conf. 

Petroleum  Expl.  Soc.  Libya:  9th  Guidebk  to  Geology:  A history  of  Tunisia.  Pp.  159-214. 
conard,  m.  and  rioult,  m.  1977.  Halimeda  elliotti  nov.  sp.,  algue  calcaire  (Chlorophyceae)  du  Turonien  des 
Alpes-maritimes  (SE  France).  Geol.  mediterran.  4,  83-98. 
dragastan,  o.  and  bucur,  I.  1979.  Upper  Aptian  Microfossils  from  the  Camenita  Valley-Sasca  Romana 
(Resita-Moldova  Nova  Zone,  Banat),  Revue  roum.  Geol.  Geophys.  Geogr.  ( ser . Geol),  23,  111-115. 
elliott,  g.  f.  1965.  The  interrelationships  of  some  Cretaceous  Codiaceae  (calcareous  algae).  Palaeontology,  8, 
199-203. 

— 1975.  Transported  algae  as  indicators  of  different  marine  habitats  in  the  English  middle  Jurassic.  Ibid.  18, 
351-366. 

— 1981.  The  Tethyan  dispersal  of  some  chlorophyte  algae  subsequent  to  the  Palaeozoic.  Palaeogeogr., 
Palaeoclimat.,  Palaeoecol.,  32,  341-358. 

flugel,  E.  1975.  Kalkalgen  aus  Riffkomplexen  der  alpin-mediterranen  Obertrias.  Verh.  Geol.  B.-A.  Wien,  Jhg. 
1974,  h.  2-3,  297-346. 

guilbault,  j.  p.  and  mamet,  b.  l.  1976.  Codiacees  (Algues)  ordoviciennes  des  Basses-Terres  du  Saint-Laurent. 
Can.  J.  Earth  Sci.  13,  636-660. 

hillis,  l.  w.  1959.  A revision  of  the  genus  Halimeda  (Order  Siphonales).  Publ.  Inst.  mar.  Sci.  Univ.  Tex.  6, 
321-403. 

hillis-colinvaux,  l.  1980.  Ecology  and  taxonomy  of  Halimeda-,  primary  producer  of  coral  reefs.  Adv.  mar.  Biol. 
17,  1-327. 

kemper,  e.,  maronde,  h.  d.  and  stoppel,  d.  1976.  Triassic  and  Jurassic  Limestone  in  the  region  Northwest  and 
West  of  Si  Sawat  (Kanchaburi  Province,  Western  Thailand).  Geol.  Jb.  B21,  93-127. 
massieux,  m.  1966.  Presence  d 'Ovulites  dans  le  ‘Calcaire  Ypresien’  des  Corbieres  septentrionales  et  discussion  sur 
la  nature  de  l’algue  Griphoporel/a  arabica  Pfender.  Rev.  Micropaleont.  8,  240-248. 
mudge,  d.  c.  1978.  Stratigraphy  and  sedimentation  of  the  Lower  Inferior  Oolite  of  the  Cotswolds.  Jl  geol.  Soc. 
Lond.  135,  611-627. 


437 


ELLIOTT:  CALCAREOUS  GREEN  ALGA 


obrhel,  j.  1968.  Maslovina  meyenii  n.g.  et  sp.— neue  Codiacea  aus  dem  Silur  Bohmens.  Vest,  ustred  Ust  seo! 

43,  367-370.  5 

PFENDER,  j.  1940.  Les  algues  du  Nummulitique  egyptien  et  des  terrains  Cretaces- Eocenes  de  quelques  regions 
mesogeennes.  Bull.  Inst.  Egypte,  22,  225-250. 

WELLS,  J.  w.  1944.  Cretaceous,  Tertiary,  and  Recent  Corals,  a sponge,  and  an  alga  from  Venezuela.  J.  Paleont 
18,  429-447. 


After  completion  of  this  paper,  I read  the  memoir  of  Dr.  D.  Vachard  (Doc.  Trav.  IGAL  Paris  no.  2, 
dated  1980,  issued  1981)  in  which  he  suggests  that  various  genera,  including  Aphroditicodium  Elliott 
and  Tauridium  Guveng  should  be  considered  as  synonyms  of  Permocalculus  (Vachard  1981,  382). 
Aphroditicodium  he  considers  to  be  a well-preserved  Permocalculus  fragilis.  The  point  is  of  some 
importance  because  it  would  transfer  these  two  genera  from  Udoteaceae  (green  algae)  to 
Gymnocodiaceae  (red  algae).  I have  examined  material  from  Burma,  much  better  preserved  than  the 
type  of  Aphroditicodium , and  more  extensive  material  of  Tauridium  from  north  Italy,  in  both  of  which 
the  medullary  structures  and  their  relations  to  the  cortical  structures  are  well  shown,  better  than  in  the 
types.  However,  in  none  of  these  specimens  are  reproductive  structures  to  be  seen  (an  absence  to  be 
expected  in  Udoteaceae),  whereas  in  Permocalculus  they  are  conspicuous  in  many  individuals,  even 
when  preservation  is  poor.  Whilst  agreeing  with  Dr.  Vachard  as  to  the  similarities  of  structural  plan  in 
these  fossil  genera,  and  as  to  the  various  changes  in  taxonomic  allocation  (to  which  I contributed)  in 
the  past,  I consider  that  Aphroditicodium  and  Tauridium  are  Udoteacean  genera,  however  wide  a 
variation  the  Gymnocodiacean  Permocalculus  shows. 


Typescript  received  20  February  1981 
Revised  typescript  received  8 May  1981 


G.  F.  ELLIOTT 

Department  of  Palaeontology 
British  Museum  (Natural  History) 
Cromwell  Road 
London  SW7  5BD 


ADDENDUM 


A RARE  LYTOCERATID  AMMONITE  FROM  THE 
LOWER  LIAS  OF  RADSTOCK 

by  D.  T.  DONOVAN  and  M.  K.  howarth 


Abstract.  A single  lytoceratid  ammonite  from  the  Jamesoni  or  Ibex  Zone  of  the  Lower  Lias  of  Clandown 
Colliery  Quarry,  Radstock,  Avon,  is  made  the  holotype  of  Derolytoceras  (£>.)  radstockense  sp.  nov.  It  is  the 
first  record  in  Britain  of  a genus  that  is  largely  restricted  to  Tethyan  areas  of  south  and  central  Europe. 

The  partial  or  complete  restriction  of  certain  groups  of  ammonites  to  the  Tethys  has  long  been 
known,  and  was  documented  in  some  detail  by  Donovan  (1967)  for  the  Lower  Jurassic.  At  this 
time,  the  Suborders  Phylloceratina  and  Lytoceratina  are  so  restricted  except  for  certain  limited 
spans  of  time  when  certain  genera  are  found  in  northern  Europe.  Thus,  Phylloceratina  are 
represented  in  Britain  by  Tragophylloceras  in  the  Pliensbachian  and  Phylloceras  in  the  Upper 
Pliensbachian  and  the  Toarcian,  and  Lytoceratina  by  Lytoceras  in  the  Upper  Pliensbachian  and 
the  Toarcian  and  by  genera  of  the  Alocolytoceratinae  in  the  Upper  Toarcian.  A second  class  of 
occurrence  comprises  rare  finds,  usually  made  from  a single  horizon.  The  following  fall  into  this 
category: 

? a member  of  Juraphyllitidae  from  the  Planorbis  Zone  of  the  Stowell  Park  Borehole, 
Gloucestershire  (Spath  1956,  p.  158).  One  of  us  (DTD)  has  re-examined  this  specimen,  which 
was  preserved  in  soft  mudstone,  but  it  is  now  in  such  an  abraded  state  that  further  comment  is 
not  possible. 

Galaticeras  jacksoni  Howarth  and  Donovan,  a member  of  Juraphyllitidae,  known  from  nine 
examples  from  the  Flatstones  (Obtusum  Subzone)  of  the  Dorset  coast;  Galaticeras  sp.  has  been 
recorded  from  two  boreholes  in  southern  England  (Howarth  and  Donovan  1964). 
Meneghiniceras  lariense  (Meneghini),  another  juraphyllitid,  a single  specimen  from  the  Grey 
Shales  (Semicelatum  Subzone)  in  the  Toarcian  near  Whitby,  Yorkshire  (Howarth  1976). 
Aegolytoceras  rotundicosta  (Tutcher  and  Trueman)  from  the  Jamesoni  Limestone  of  Radstock, 
discussed  in  detail  below. 

The  present  paper  records  an  example  of  a lytoceratid  genus  new  to  Britain,  from  the  Lower  Lias 
of  Radstock,  collected  sixty  years  ago  but  only  recently  recognized  for  what  it  is. 

SYSTEMATIC  PALAEONTOLOGY 

Superfamily  lytocerataceae  Neumayr  1875 
Family  derolytoceratidae  Spath  1927 
Genus  Derolytoceras  Rosenberg  1909 

Type  species.  Ammonites  lineatus  tortus  Quenstedt  1885,  subsequently  designated  by  Spath  1924  (p.  4). 

Subgenus  derolytoceras 
Derolytoceras  ( Derolytoceras ) radstockense  sp.  nov. 

Text-fig.  1 

Diagnosis.  Derolytoceras  attaining  at  least  80  mm  diameter,  with  evolute,  slowly  expanding  whorls, 
and  an  evenly  rounded  elliptical  whorl  section.  Regular  radial  ribs  are  strong  from  13  mm  diameter, 

(Palaeontology,  Vol.  25,  Part  2,  1982,  pp.  439-442.] 


440 


PALAEONTOLOGY,  VOLUME  25 


text-fig.  1.  Derolytoceras  ( D .)  radstockense  sp.  nov.  Holotype.  Jamesoni  Limestone,  Lower  Pliensbachian, 
Jamesoni  or  Ibex  Zones,  Clandown  Colliery  Quarry,  Radstock,  Avon.  Bristol  University  Geology  Museum 

no.  2877,  x 0-95. 

and  curve  forwards  to  the  middle  of  the  venter.  Three  or  four  constrictions  per  whorl  are  similar 
in  shape  to  the  ribs,  and  are  accentuated  by  an  enlarged  rib  immediately  in  front.  The  suture-line 
is  Lytoceratid  with  highly  indented  and  undercut  saddles. 

Material.  The  holotype  only,  Bristol  University  Geology  Museum  no.  2877,  from  the  Jamesoni  Limestone, 
Lower  Lias  (Lower  Pliensbachian:  Jamesoni  or  Ibex  Zone)  at  Clandown  Colliery  Quarry  (ST  679  558),  near 
Radstock,  Avon.  For  the  section  at  this  quarry  see  Tutcher  and  Trueman  1925,  p.  600.  Collected  in  1922  or 
1923  by  Mr.  T.  R.  Fry. 

Measurements.  Maximum  diameter  80-5  mm.  At  78T  mm  diameter,  whorl  height  = 23-8  mm  (0-30),  whorl 
breadth  =181  mm  (0-23),  umbilical  width  = 38  0 mm  (0-49).  There  are  thirty-three  ribs  and  four  constrictions 
on  the  outer  whorl  ending  at  80  mm  diameter. 

Description.  The  specimen  consists  of  about  four  visible  whorls  ending  at  a diameter  of  80  mm.  The  mouth 
border  is  not  present,  for  at  least  the  final  one-eighth  of  a whorl  is  missing.  Suture-lines  can  be  clearly  seen 
just  over  half  a whorl  behind  the  aperture  at  a diameter  of  57  mm.  From  the  rough  state  of  preservation  it 
is  not  possible  to  see  whether  there  are  more  suture-lines  at  a larger  size,  but  the  phragmocone  may  have 
extended  about  one-eighth  of  a whorl  further  forwards.  This  makes  the  length  of  the  body  chamber  that  is 
preserved  between  three-eighths  and  half  a whorl  in  length.  The  whorls  are  evolute  but  not  completely  so, 
for  each  whorl  overlaps  about  20%  of  the  next  inner  whorl.  The  whorl  section  is  compressed  elliptical  and 
has  an  evenly  rounded  venter  and  umbilical  wall.  Capricorn  ribs  are  well  developed  on  all  visible  whorls  from 
the  smallest  size  seen,  about  13  mm  diameter,  up  to  the  aperture,  though  they  diminish  in  strength  on  the 
last  half  whorl.  They  are  approximately  radial  on  the  side  of  the  whorl,  then  curve  forwards  towards  the 
middle  of  the  venter.  There  are  four  well-marked  constrictions  on  the  outer  whorl  that  follow  the  line  of  the 
ribs  exactly.  They  are  made  more  prominent  by  an  enlarged  rib  immediately  in  front  of  each  one.  Three 
constrictions  can  be  seen  on  the  next  inner  whorl,  but  at  smaller  sizes  the  preservation  is  not  good  enough 
for  constrictions  to  be  visible. 

The  suture-line  (text-fig.  2)  is  typically  Lytoceratid.  The  large  first  and  second  lateral  saddles  are  deeply 
indented  and  undercut,  and  have  complicated  moss-like  endings.  The  large  first  lateral  lobe  is  divided  into  I 
three  prongs. 

Remarks.  Ammonites  of  the  superfamily  Lytocerataceae,  other  than  Lytoceras  itself,  are  almost 
unknown  in  the  Sinemurian  and  Pliensbachian  in  Britain.  The  only  one  known  previously  is  the 


DONOVAN  AND  HOWARTH:  LYTOCERATID  AMMONITE 


441 


single  specimen  from  the  Jamesoni  Limestone  at  Radstock  that  was  the  holotype  and  sole  basis 
of  the  name  ‘ Peripleuroceras'  rotundicosta  Tutcher  and  Trueman  (1925,  p.  646,  pi.  41,  fig.  1) 
(BM(NH)  C. 41 760).  That  specimen  is  smooth  up  to  about  30  mm  diameter,  and  then  develops 
capricorn  ribs  mainly  on  the  venter,  that  have  become  fairly  strong  by  its  maximum  size  of  49 
mm  diameter.  It  is  wholly  septate.  It  belongs  to  the  Derolyoceratidae  genus  Aegolytoceras  Spath 
1924,  of  which  Peripleuroceras  Tutcher  and  Trueman  1925  is  a junior  synonym.  Other  species  of 
Aegolytoceras  develop  the  characteristic  coarse  capricorn  ribs  immediately  behind  constrictions, 
reverting  to  fine  ribbing  in  front  of  constrictions.  Derolytoceras  ( D .)  radstockensis  differs  in  having 
regularly  strong  capricorn  ribs  that  develop  at  an  earlier  growth  stage,  and  the  only  irregularity 
on  the  phragmocone  is  the  single  slightly  stronger  rib  that  follows  in  front  of  each  constriction. 
It  is  difficult  to  tell  whether  the  decrease  in  rib  strength  on  the  last  half  whorl  is  due  to  poor 
preservation  or  is  the  onset  of  adult  ornamentation.  These  are  features  of  the  genus  Derolytoceras , 
of  which,  at  80  mm  diameter,  it  is  one  of  the  largest  known  examples.  The  subgenus  D. 
(Derolytoceras)  has  ribs  from  an  early  growth  stage,  certainly  by  13  mm  diameter  which  is  the 
smallest  size  visible  in  this  Radstock  specimen,  while  the  subgenus  D.  ( Tragolytoceras ) remains 
smooth  up  to  about  25  mm  diameter.  So  the  Radstock  specimen  belongs  to  the  nominal  subgenus. 
The  type  species  is  D.  (D.)  tortum  (Quenstedt),  of  which  the  lectotype  (Quenstedt,  1885,  p.  309, 
pi.  39,  fig.  15),  refigured  by  Wiedmann  (1970,  p.  995,  text-fig.  8e,  pi.  6,  fig.  3),  is  from  the  Upper 
Pliensbachian  of  south-west  Germany.  It  has  quickly  expanding  whorls  and  very  strong  ribs 
commence  at  about  11  mm  diameter.  Another  Sinemurian  or  Pliensbachian  species  is  D.  (D.) 
haueri  Rosenberg  (1909,  p.  251,  pi.  11,  figs.  31,  32),  which  is  based  on  specimens  of  up  to  15  mm 
diameter  that  have  slowly  expanding  and  finely  ribbed  evolute  whorls.  Though  it  is  difficult  to 
compare  the  much  larger  Radstock  specimen  with  either  of  these  species,  it  appears  to  differ  from 
both  of  them  in  its  slowly  expanding  whorls  and  coarse  ribs. 

At  first  sight  the  Radstock  specimen  seems  to  have  considerable  resemblances  to  the  Polymorphitid 
genus  Platypleuroceras,  and  many  examples  of  P.  brevispina  (J.  de  C.  Sowerby)  or  closely  allied 
species  have  been  found  in  the  Jamesoni  Limestone  at  Radstock  (e.g.  Tutcher  and  Trueman  1925, 
p.  650,  pi.  39,  fig.  3;  pi.  40,  fig.  2).  The  Derolytoceras  differs,  however,  in  lacking  the  ventro-lateral 
tubercle  that  occurs  in  all  the  specimens  of  Platypleuroceras , and  in  having  more  strongly  projected 
ribs  on  the  venter.  It  also  has  constrictions  that  are  only  rarely  present  in  Radstock  Platypleuroceras 
(though  constrictions  are  better  developed  in  similar  Polymorphitidae  from  Germany,  e.g.  Quenstedt 
1885,  pi.  32,  fig.  6;  pi.  33,  figs.  11,  12).  Finally,  the  suture-line  of  the  Derolytoceratid  is  distinctively 
Lytoceratid,  and  suture-lines  in  Platypleuroceras  do  not  have  such  deeply  indented  or  undercut 
saddles. 

The  family  and  generic  classification  used  here  is  that  of  the  Treatise  (Arkell,  1957,  p.  LI 94). 
Two  more  recently  proposed  generic  names  are  Adnethiceras  Wiedmann  (1970,  p.  997)  for  species 
that  have  ventro-lateral  tubercles,  and  Lytoconites  Wiedmann  (1970,  p.  1004)  which  is  considered 
here  to  be  a synonym  of  Aegolytoceras.  Peripleuroceras  Tutcher  and  Trueman  1925,  is  also  a 
synonym  of  the  latter  genus.  Wiedmann  (1970,  p.  988)  did  not  admit  a separate  family 
Derolytoceratidae,  but  placed  it  and  its  included  genera  in  the  Lytoceratidae.  The  distinctive 


442 


PALAEONTOLOGY:  VOLUME  25 


capricorn  ornament  of  Derolytoceratidae  seems  to  us,  however,  to  be  sufficiently  different  from 
the  ornament  of  normal  Lytoceratidae  to  warrant  separation  of  this  Sinemurian  to  Toarcian  group 
as  a family.  Most  genera  of  Derolytoceratidae  were  also  discussed  by  Fantini  Sestini  (1973)  who 
reinterpreted  the  genus  Audaxlytoceras  Fucini  1923  (type  species  Ammonites  audax  Meneghini  1881 
(non  Oppel,  1863))  on  the  basis  of  specimens  that  were  larger  than  the  very  small  originals,  and 
concluded  that  it  was  the  oldest  name  for  the  group  of  species  that  have  usually  been  referred  to 
Aegolytoceras  Spath  1924. 

The  best  collection  of  Derolytoceratidae,  which  was  not  discussed  by  either  Wiedmann  (1970) 
or  Fantini  Sestini  (1973),  comes  from  the  Pliensbachian  at  Monte  di  Cetona,  central  Italy.  Nine 
species  of  "Deroceras'  were  described  by  Fucini  (1903,  pp.  166-185):  all  of  them  have  Lytoceratid 
suture-lines,  and  amongst  them  are  the  largest  known  Derolytoceratidae,  some  specimens  attaining 
165  mm  diameter.  Several  develop  ventro-lateral  tubercles  and  belong  to  Adnethiceras  (e.g.  A. 
olenoptychum  Fucini,  A.  mutans  Fucini  and  A.  instabile  Fucini),  and  about  eight  examples  were 
figured  of  Aegolytoceras  pecchiolii  (Meneghini),  which  is  the  most  highly  developed  species  of  its 
genus.  Nothing  in  the  collection  is  exactly  like  the  Radstock  specimen,  though  the  rather  poorly 
preserved  examples  of  Derolytoceras  connexum  Fucini  (1903,  p.  176,  pi.  26,  figs.  7,  8)  are  the 
nearest,  differing  mainly  in  lacking  clear  constrictions. 

REFERENCES 

arkell,  w.  J.  1957.  In  R.  c.  MOORE,  Treatise  on  invertebrate  paleontology.  Part  L,  Mollusca  4,  Cephalopoda, 
Ammonoidea.  Geol.  Soc.  Amer. 

donovan,  d.  t.  1967.  The  geographical  distribution  of  Lower  Jurassic  ammonites  in  Europe  and  adjacent  areas. 
Pubis  Syst.  Ass.  7,  111-134. 

fantini  sestini,  n.  1973.  Revisione  del  genere  Audaxlytoceras  Fucini,  1923  (Ammonoidea).  Riv.  ital.  Paleont. 
Stratigr.  79,  479-502,  pi.  49. 

fucini,  a.  1903.  Cefalopodi  liassici  del  Monte  di  Cetona.  Part  3.  Palaeontogr.  ital.  9,  125-185,  pis.  19-27. 
howarth,  m.  k.  1976.  An  occurrence  of  the  Tethyan  ammonite  Meneghiniceras  in  the  Upper  Lias  of  the 
Yorkshire  coast.  Palaeontology , 19,  773-777. 

— and  donovan,  d.  t.  1964.  Ammonites  of  the  Liassic  family  Juraphyllitidae  in  Britain.  Ibid.,  7, 286-305,  pis. 
48,  49. 

quenstedt,  F.  A.  1885.  Die  Ammoniten  des  Schwdbischen  Jura.  1,  der  Schwarze  Jura  (Lias),  241-440,  pis.  31-54. 
Tubingen. 

rosenberg,  p.  1909.  Die  Liasische  Cephalopodenfauna  der  Kratzalpe  im  Hagengebirge.  Beitr.  Palaont.  Geol. 
Ost.-Ung.  22,  193-345,  pis.  10-16. 

spath,  l.  f.  1924.  On  the  Blake  Collection  of  Ammonites  from  Kachh,  India.  Mem.  geol.  Surv.  India  Palaeont. 
indica,  9,  mem.  1,  29  pp. 

— 1956.  The  Liassic  ammonite  faunas  of  the  Stowell  Park  Borehole.  Bull.  geol.  Surv.  Gt.  Br.  11,  140-164,  pis. 
9,  10. 

tutcher,  j.  w.  and  trueman,  a.  e.  1925.  The  Liassic  rocks  of  the  Radstock  district,  Somerset.  Q.  Jl.  geol.  Soc. 
Lond.  81,  595-666,  pis.  38-41. 

wiedmann,  j.  1970.  Uber  den  Ursprung  der  Neoammonoideen— Das  Problem  einer  Typogenese.  Eclog.  geol. 
Helv.  63,  923-1020,  pis.  1-10. 

D.  T.  DONOVAN 
Department  of  Geology 
University  College 
Gower  Street 
London  WC1E  6BT 

M.  K.  HOWARTH 
Department  of  Palaeontology 
British  Museum  (Natural  History) 
Cromwell  Road 

Typescript  received  13  October  1980  London  SW7  5BD 


THE  PALAEONTOLOGICAL  ASSOCIATION 


The  Association  was  founded  in  1957  to  further  the  study  of  palaeontology.  It  holds  meetings  and  demonstrations  as  well 
as  publishing  Palaeontology  and  Special  Papers  in  Palaeontology.  Membership  is  open  to  individuals  and  to  institutions  on 
payment  of  the  appropriate  annual  subscription.  Rates  l or  1982  arc: 

Institutional  membership . . . £32  (U.S.  $71) 

Ordinary  membership.  . . . £15  (U.S.  $33) 

Student  membership.  . £9-50  (U.S.  $21) 

There  is  no  admission  fee.  Correspondence  concerned  with  Institutional  Membership  should  be  addressed  to  Dr.  C.  H.  C. 
Brunton,  Department  of  Palaeontology,  British  Museum  (Natural  History),  Cromwell  Road,  London  SW7  5BD,  England. 
Student  members  are  persons  receiving  full-time  instruction  at  educational  institutions  recognized  by  the  Council.  On  first 
applying  for  membership,  an  application  form  should  be  obtained  from  the  Membership  Treasurer.  Subscriptions  cover  one 
calendar  year  and  are  due  each  January;  they  should  be  sent  to  the  Membership  Treasurer  (for  name  and  address  see 
Council  list  below). 


PALAEONTOLOGY 

All  members  who  join  for  1982  will  receive  Volume  25,  Parts  1-4.  All  back  numbers  are  still  in  print  and  may  be  ordered  from 
Marstori  Book  Services,  P.O.  Box  87,  Oxford  0X4  1LB,  England,  at  £15  (U.S.  $33)  per  part  (post  free). 

SPECIAL  PAPERS  IN  PALAEONTOLOGY 

Members  may  subscribe  to  this  by  writing  to  the  Membership  Treasurer  (for  name  and  address  see  Council  list  below): 
the  subscription  rate  for  1982  is  £25  (U.S.  $55)  for  Institutional  Members,  and  £12-50  (U.S.  $27-50)  for  Ordinary  and  Student 
Members.  A single  copy  of  each  Special  Paper  is  available  to  Ordinary  and  Student  Members  only,  for  their  personal  use, 
at  a discount  of  25%  below  the  prices  shown  in  the  list  inside  the  front  cover.  Non-members  may  obtain  copies,  at  the 
listed  prices,  from  Marston  Book  Services,  P.O.  Box  87,  Oxford  0X4  1LB,  England. 

COUNCIL  1982-1983 

President : Professor  A.  Hallam,  Department  of  Geological  Sciences,  The  University,  Birmingham  B15  2TT 
Vice-Presidents : Professor  J.  W.  Murray,  Department  of  Geology,  The  University,  Exeter  EX4  4QE 
Dr.  R.  A.  Fortey,  Department  of  Palaeontology,  British  Museum  (Natural  History),  Cromwell  Road,  London  SW7  5BD 
Treasurer.  Dr.  M.  Romano,  Department  of  Geology,  The  University,  Sheffield  SI  3JD 
Membership  Treasurer.  Dr.  J.  C.  W.  Cope,  Department  of  Geology,  University  College,  Swansea  SA2  8PP 
Secretary : Dr.  R.  Riding,  Department  of  Geology,  University  College,  Cardiff  CF1  1XL 
Marketing  Manager:  Dr.  R.  J.  Aldridge,  Department  of  Geology,  The  University,  Nottingham  NG7  2RD 


Editors 

Dr.  K.  C.  Allen,  Department  of  Botany,  The  University,  Bristol  BS8  1UG 
Dr.  M.  G.  Bassett,  Department  of  Geology,  National  Museum  of  Wales,  Cardiff  CF1  3NP 
Dr.  D.  E.  G.  Briggs,  Department  of  Geology,  Goldsmiths’  College,  London  SE8  3BU 
Dr.  L.  B.  Halstead,  Department  of  Geology,  The  University,  Reading  RG6  2AB 


Other  Members  of  Council 


Dr.  C.  H.  C.  Brunton,  London 

Dr.  E.  N.  K.  Clarkson,  Edinburgh 

Dr.  D.  Edwards,  Cardiff 

Dr.  R.  Harland,  Leeds 

Dr.  P.  D.  Lane,  Keele 

Dr.  A.  R.  Lord,  London 

Dr.  J.  Miller,  Edinburgh 


Dr.  T.  J.  Palmer,  Aberystwyth 

Dr.  D.  J.  Siveter,  Hull 

Dr.  P.  W.  Skelton,  Open  University 

Dr.  P.  D.  Taylor,  London 

Dr.  A.  Thomas,  Birmingham 

Dr.  H.  S.  Torrens,  Keele 

Dr.  N.  H.  Trewin,  Aberdeen 


Overseas  Representatives 

Australia : Professor  B.  D.  Webby,  Department  of  Geology,  Sydney  University,  Sydney,  N.S.W.,  2006 
Canada : Dr.  B.  S.  Norford,  Institute  of  Sedimentary  and  Petroleum  Geology,  3303-33rd  Street  NW.,  Calgary,  Alberta 
New  Zealand : Dr.  G.  R.  Stevens,  New  Zealand  Geological  Survey,  P.O.  Box  30368,  Lower  Hutt 
West  Indies  and  Central  America:  Mr.  John  B.  Saunders,  Geological  Laboratory,  Texaco  Trinidad,  Inc.,  Pointe-a-Pierre, 
Trinidad,  West  Indies 

U.S. A:  Dr  R.  Cuffey,  Department  of  Geology,  Pennsylvania  State  University,Pennsylvania,  U.S. A. 

South  America:  Dr.  O.  A.  Reig,  Departmento  de  Ecologia,  Universidad  Simon  Bolivar,  Caracas  108,  Venezuela 


Palaeontology 

VOLUME  25  • PART  2 

CONTENTS 


Ecology  and  population  structure  of  the  Recent  brachiopod 
Terebratulina  from  Scotland 

GORDON  B.  CURRY  227 

A new  zosterophyll  from  the  Lower  Devonian  of  Poland 

DANUTA  Z DEB  SKA  247 

Two  salenioid  echinoids  in  the  Danian  of  the  Maastricht  area 

j.  F.  geys  265 

Devonian  miospore  assemblages  from  Fair  Isle,  Shetland 

J.  E.  A.  MARSHALL  and  K.  C.  ALLEN  277 

Conodonts,  goniatites  and  biostratigraphy  of  the  earlier  Carboni- 
ferous from  the  Cantabrian  Mountains,  Spain 

A.  C.  HIGGINS  and  C.  H.  T.  WAGNER-GENTIS  313 

Liassic  plesiosaur  embryos  reinterpreted  as  shrimp  burrows 
RICHARD  A.  THULBORN  351 

Limpet  grazing  on  Cretaceous  algal-bored  ammonites 

ETIE  BEN  AKPAN,  GEORGE  E.  FARROW,  and  NOEL  MORRIS  361 

A review  of  Recent  and  Quaternary  organic-walled  dinoflagellate 
cysts  of  the  genus  Protoperidinium 

REX  HARLAND  369 

A new  genus  of  shark  from  the  Middle  Triassic  of  Monte  San 
Giorgio,  Switzerland 

o.  rieppel  399 

A new  species  of  the  fish  Amia  from  the  Middle  Eocene  of  British 
Columbia 

MARK  V.  H.  WILSON  413 

A fused  cluster  of  coniform  conodont  elements  from  the  late 

Ordovician  of  Washington  Land,  Western  North  Greenland 

R.  J.  ALDRIDGE  425 

A new  calcareous  green  alga  from  the  Middle  Jurassic  of  England: 
its  relationships  and  evolutionary  position 

GRAHAM  F.  ELLIOTT  431 

A rare  lytoceratid  ammonite  from  the  Lower  Lias  of  Radstock 

D.  T.  DONOVAN  and  M.  K.  HOWARTH  439 


Printed  in  Great  Britain  at  the  University  Press,  Oxford 
by  Eric  Buckley,  Printer  to  the  University 


issn  003 1-0239 


ITUTtON  NOliniliSNi  NVINOSHJLI1NS  S3IHVHen  LIBRARIES  SMITHSONIAN  INSTITUTION  NOI1 
z ~ \«  z.  r-  z r~  2 


m 2 m ~ * ' m ^ wj-J^  *2  >s 


CO  f , t , 

~ W '*  E tO  ~ CO  ” 

ivaan  libraries  Smithsonian  institution  NoiiniusNi  nvinoshjjws  ssiavaan  libi 

z w 2 CO  2 w . « V w a I I 

< ^ S S S < 


tution  NoiinniSNi_ nvinoshiiws  saiavaan  libraries  Smithsonian  institution  noii 

<2  „ — ^ z \ 40  ^ ^ 5 oo  ...  = oo 


jvaan  libraries  Smithsonian  institution  NoiiniusNi  nvinoshliws  saiavaan  lib 

“ z r-  . z r-  z “ 


ITUTION  NOlinilJLSNI  NVINOSHJLIWS  S3iavaail  LIBRARIES  SMITHSONIAN  INSTITUTION  MOIJ 
in  2 to  2 ....  to  2 ...  co 


O 
z 

> "4^  S v.  > 

avaan  libraries  Smithsonian  institution  NoiiniusNi  NviNOSHiiws^saiavaan  lib 


5 ” v 5 NSjypcjy  " 5 “ ^oIu%^  O 

itution2  NoiiniusNi  "VjviNOSHims^sa  lavaail^LIBRARI  eszsmithsonian-,institution2noij 

Z r-  Z r-  2 f“_z 


yosy^X  E?  m ^ m S£  * m to 

jvaan  i 

(fe  I 


7 ' 1 ' \V  rr  VL,rt3L>^  I ' I -r  ' * ' ~r 

~ CO  £ CO  ? CO  £ 

jvaan  libraries  Smithsonian  institution  NoiiniusNi  nvinoshiiws  S3iavaan  libi 

^ <£  ^ . z > to  z to  z 

s i s i •*  ' ^ ^ i I 

OO  A-  2 ^ •-•«  2 to  Z w 

ITUTION  NOliniliSNi  _ NVIN0SH1IWS  S3iaV08n  LIBRARIES  SMITHSONIAN  INSTITUTION  NOII 

CO  ~ CO  — CO 


&JP  s iOP  I s I1&JP I lOp  IImP  « if 


O _ O 

z — _i  z 

jvaan  libraries  Smithsonian  institution  NoiiniusNi  nvinoshiiws  saiavaan  lib 

z •”  — z r-  z r_ 


Oin s2iz  m ^ m *2  m 


sjviNOSHims  S3iavaan  libraries  Smithsonian  institution  NOimillSNI  nvinoshuins. 

n,  Z r~  Z r-  z ^ > ■ 

O — Z-iau7N  O O 


SMITHSONIAN  INSTITUTION  NOimillSNI  NVIN0SH1MS  S3IBVHan  LI  B R AR  1 ES  SMITHSONIAN 

z c/>  z , <2  z 


7 z 

>'  S x-vosj-ix  2 ‘".W  > 2 

(«  •••  2 CO  Z CO  - Z </) 

nvinoshiiws  saiavaan  libraries  Smithsonian  institution  NouniiiSNi  nvinoshhws 

- w 5 ' 


J < 

;Qva5*S>''/  O 5 ’ O ~ '^viVA5J^/ 

z 

SMITHSONIAN  INSTITUTION  NOIlfUllSNI  NVIN0SH1I1NS  S3  I U VH  3 H_LI  B R AR  I ES  SMITHSONIAN^ 

P o I 


rn  «;  rn  I"  m xg\  pcz  ^ 'f^p"  m 

co  _ c/>  \ fE  co  — w 

NVIN0SH1IINS  S3iavnan  LIBRARIES  SMITHSONIAN  INSTITUTION  NOIlfUllSNI  nvinoshiiws 

CO  Z CO  2 v.„  CO  z 


5 v > 2 ^ > 

Z CO  * Z £/)  v z 

SMITHSONIAN  INSTITUTION  NOIlfUllSNI  NVINOSH1IIMS  S3IHVyail  LIBRARIES  SMITHSONIAN 


O X^AI  pi>-  _ O ~ Q _ •*??«"  O 

NVIN0SHimSZS3  I y vy  a n^LI  B RAR  I ESZSMITHSONIAN~,|NSTiTUTIONZNOIlfUllSNl“VlNOSHimS. 
z r-  Z r-  2 r- 

o zCstit/T^x  ~ O _ /oa  VnX  O 


SMITHSONIAN  INSTITUTION  NOIlfUllSNI  NVIN0SH1MS  S3iyvyail  LI  B RAR  I ES  SMITHSONIAN 
z > co  z co  z > 52  ^ — . z 


2 >•  2 >’  2 s 

CO  ' Z CO  z to  z </ 

nvinoshiiws  S3iyvyan  libraries  smithsonian  institution  noiidiiisni_nvinoshiiws 

W =:  co  - CO  2 ' 

Kk  h am  - /#w\  g "%%  D 


o 

-1  z _J  z 

SMITHcnNIAN  INSTITUTION  NOimillSNI  NVIN0SH1IWS  S3iyvyaiT  LIBRARIES  SMITHSONIAN 
Z - Z _ H g [ 


a