Skip to main content

Full text of "Psyche"

See other formats


IlfllllSNI  ""'nVINOSHJLI  3 I H V d a !1  "*  L I B R A R I ES^  SMITHSONIAN^INSTITUTli 
r-  , z r*  2 r- 

o 


CO 


- St  I — 

z \ m 

CO  « CO  H CO 

3 R ARIES  SMITHSONIAN  INSTITUTION  NOlinilXSNi  NVINOSHillNS  S3IHVtfa 
<2  ^ ^ z:  co  z 2 

1 %%  1 I 

v z t t \3X4AKsv  = > 

linillSNI  ~NVIN0SHJLIWS^$3  I dVd  a h2li  b RAR  I ES^SMITHSONIAN^INSTITUTI 

■ 00  - CO  — 


5 W 2 1 

z 'X'  -I  z — _|  z 

3 R ARIES  SMITHSONIAN  INSTITUTION  NOlinilJLSNI  NVIN0SH1IWS  S3  I H Vd  € 
5 „ ^ z r-  z 


Oi 
2/ 
of 
o! 

m “ Vt' 

IXfUIXSNI  NVINOSHXIWS  S3  I U Vd  3 n~LI  B RAR  I ES^SMITHSONIAN^NSTITUTI 

I ^SL 

MR  z >wv\'  _i  ^ 


> 5 S 

O)  z co  * z — ^ *• 

SRARIES  SM ITHSON IAN _ INSTITUTION  NOIXfUIXSNI_NVINOSHXIWS  S3ldVdf 


O ><1«XI XV>^  Q 

linillSNI  ^NVINQSHillNS^S  3 I d VH  QIl^LIBRARI  ES^SMITHSONIAN^INSTITUTI 
r*  * z r-  z r- 

2 


m • Z p-j 

3RAR  I ES  SMITHSON  IAN  ""INSTITUTION  NOlifUIJLSNI  ~ NVIN0SH1IWS  S3  I HVHE 

z co  z co 


^ — p - 

> \!vass>-  ^ > ''  2 

linillSNI  NVINOSHJUIAJS^SS  I d VH  a llZLI  B RAR  I ES^SMITHSONIAN ^INSTITUTI 
~ 00  = co 


CO 


O Q 

lini!iSNI_lNVIN0SHllWS:ZS3  I avaa  n^LIBRAR  I ES^SMITHSONIANJNSTITUTIC 


00 
3D 
> 

33 

3RAR I ES  SMITHSONIAN  INSTITUTION  NOlinillSNI  NVIN0SH1MS  S31BVH9 

CO  21  r C/>  2 

T/7?V 


X O 

in 
O 

? * 

> 2 ' )^v  > ' 2 . 

linillSNlf  NVIN0SH1IWS^S3  I HVH  8 ll\l  BRAR  I ES^SMITHSONIAN  JNSTITUTK 

l | /#%.  l S <jfkf  l 

J S ' K (tfLoSl)  c (|§.  .31)  < c W 


kE©?'  5 " x^ss^  5 5 

2 ' «l  2 " j Z 

3RARIES  SMITHSONIAN  INSTITUTION  NOlinillSNI  NVIN0SH1IWS  S3IBVaa 

- - z r-  _ # 2 

c 


m 

mniliSNI  NVINOSHJLIWS  S3  I dVd  8 n-LI  B RAR  I ES  SMITHSONIAN  INSTITUTI 

C/5  Z 


2 /■»«* A --  • I 1 i A;- 

O X 

S I x\  > 

BRAR  I ES^SMITHSONIAN  INSTITUTION  NOlinillSNI  NVINOSHIIKIS^SS  I UM& E 
</>  =:  co 


eg 
< 
c/  a 

m § - B ^ ■ 

ilinillSNI^NVINOSHHIAIS^SS  I d VB  8 II^LI  B RAR  I ES^SMITHSONIAN^INSTITUTI 


BRARIES  SMITHSONIAN  INSTITUTION  NOlinillSNI  NVIN0SH1MS  S3IHVH8 


PSYCHE 

A Journal  of  Entomology 


Volume  88 
1981 


Editorial  Board 

Frank  M.  Carpenter,  Editor  P.  J.  Darlington,  Jr. 

W.  L.  Brown,  Jr.  H.  W.  Levi 

E.  O.  Wilson  Alfred  F.  Newton,  Jr. 

B.  K.  Holldobler  R.  E.  Silberglied 

Ronald  J.  McGinley 


Published  Quarterly  by  the  Cambridge  Entomological  Club 
Editorial  Office:  Biological  Laboratories 
16  Divinity  Avenue 
Cambridge,  Massachusetts,  U.S.A. 


The  numbers  of  Psyche  issued  during  the  past  year  were  mailed  on  the  following 
dates: 

Vol.  87,  no.  1-2,  for  1980,  April  27,  1981 

Vol.  87,  no.  3-4,  for  1980,  August  31,  1981 

Vol.  88,  no.  1-2,  for  1981,  December  28,  1981 


Wj  / 

PSYCHE 

A JOURNAL  OF  ENTOMOLOGY 

founded  in  1874  by  the  Cambridge  Entomological  Club 

Vol.  88  1981  No.  1-2 

CONTENTS 

Anti-predator  Strategies.  II.  Grasshoppers  (Orthoptera,  Acrididae)  Attacked 
by  Prionyx  parkeri  and  Some  Tachysphex  Wasps  (Hymenoptera,  Sphecinae 

and  Larrinae):  A Descriptive  Study.  A.  L.  Steiner  1 

A Comparison  of  the  Nest  Phenologies  of  Three  Species  of  Pogonomyrmex 

Harvester  Ants  (Hymenoptera:  Formicidae).  William  P.  MacKay  25 

Laboratory  Evaluation  of  Within-species,  Between-Species,  and  Partheno- 
genetic  Reproduction  in  Reticulitermes  flavipes  and  Reticulitermes  vir- 
ginicus.  Ralph  W.  Howard,  Eldon  J.  Mallette,  Michael  /.  Haverty,  and 

Richard  V.  Smythe  75 

Ecology  and  Life  History  of  the  Rhytidoponera  impressa  Group  (Hymenop- 
tera: Formicidae).  I.  Habitats,  Nest  Sites,  and  Foraging  Behavior. 

Philip  S.  Ward 89 

Ecology  and  Life  History  of  the  Rhytidoponera  impressa  Group  (Hymenoptera: 
Formicidae).  II.  Colony  Origin,  Seasonal  Cycles  and  Reproduction. 

Philip  S.  Ward 109 

The  Ontogeny  of  Lyssomanes  viridis  (Walckenaer)  (Araneae:  Salticidae)  on 
Magnolia  grandiflora  L. 

David  B.  Richman  and  Willard  H.  Whitcomb 127 

The  Emigration  Behavior  of  Two  Species  of  the  Genus  Pheidole  (Hymenoptera: 

Formicidae).  Robert  Droual  and  Howard  Topoff 135 

Statary  Behavior  in  Nomadic  Colonies  of  Army  Ants:  The  Effect  of  Overfeed- 
ing. Howard  Topoff,  Aaron  Rothstein,  Susan  Pujdak,  and 

Tina  Dahlstrom  151 

Life  History  of  Antaeotricha  sp.  (Lepidoptera:  Oecophoridae:  Stenomatinae)  in 

Panama.  Annette  Aiello  163 

Polistes  gallicus  in  Massachusetts  (Hymenoptera:Vespidae). 

Mary  Hathaway 169 

Notes  on  the  Population  Ecology  of  Cicadas  (Homoptera:  Cicadidae)  in  the 
Cuesta  Angel  Forest  Ravine  of  Northeastern  Costa  Rica. 

Allen  M.  Young 


175 


CAMBRIDGE  ENTOMOLOGICAL  CLUB 
Officers  for  1980-1981 


President  

Vice-President  

Secretary  

Treasurer  

Executive  Committee 


Jo  Brewer  Winter 
Barbara  L.  Thorne 
Heather  Hermann 
Frank  M.  Carpenter 
William  D.  Winter,  Jr. 
Ernest  Leblanc 


EDITORIAL  BOARD  OF  PSYCHE 

F.  M.  Carpenter  (Editor),  Fisher  Professor  of  Natural  History, 
Emeritus,  Harvard  University 

W.  L.  Brown,  Jr.,  Professor  of  Entomology,  Cornell  University  and 
Associate  in  Entomology,  Museum  of  Comparative  Zoology 
P.  J.  DARLINGTON,  Jr.,  Professor  of  Zoology,  Emeritus,  Harvard 
University 

B.  K.  HoLLDOBLER,  Professor  of  Biology,  Harvard  University 
H.  W.  Levi,  Alexander  Agassiz  Professor  of  Zoology,  Harvard  University 
R.  J.  McGlNLEY,  Assistant  Professor  of  Biology,  Harvard  University 
Alfred  F.  Newton,  Jr.,  Curatorial  Associate  in  Entomology,  Harvard 
University 

R.  E.  SlLBERGLlED,  Smithsonian  Tropical  Research  Institute,  Panama 
E.  O.  WILSON,  Baird  Professor  of  Science,  Harvard  University 


PSYCHE  is  published  quarterly  by  the  Cambridge  Entomological  Club,  the  issues 
appearing  in  March,  June,  September  and  December.  Subscription  price,  per  year, 
payable  in  advance:  $11.00,  domestic  and  foreign.  Single  copies,  $3.50. 

Checks  and  remittances  should  be  addressed  to  Treasurer,  Cambridge 
Entomological  Club,  16  Divinity  Avenue,  Cambridge,  Mass.  02138. 

Orders  for  missing  numbers,  notices  of  change  of  address,  etc.,  should  be  sent  to 
the  Editorial  Office  of  Psyche,  16  Divinity  Avenue,  Cambridge,  Mass.  02138.  For 
previous  volumes,  see  notice  on  inside  back  cover. 

IMPORTANT  NOTICE  TO  CONTRIBUTORS 
Manuscripts  intended  for  publication  should  be  addressed  to  Professor  F.  M. 
Carpenter,  Biological  Laboratories,  Harvard  University,  Cambridge,  Mass.  02138. 

Authors  are  expected  to  bear  part  of  the  printing  costs,  at  the  rate  of  $27.50  per 
printed  page.  The  actual  cost  of  preparing  cuts  for  all  illustrations  must  be  borne  by 
contributors:  the  cost  for  full  page  plates  from  line  drawings  is  ordinarily  $18.00 
each,  and  for  full  page  half-tones,  $30.00  each;  smaller  sizes  in  proportion. 


Psyche,  vol.  87,  no.  3-4,  for  1980,  was  mailed  August  31,  1981 


The  Lexington  Press,  Inc.,  Lexington,  Massachusetts 


PSYCHE 


Vol.  88 


1981 


No.  1-2 


ANTI-PREDATOR  STRATEGIES. 

II.*  GRASSHOPPERS  (ORTHOPTERA,  ACRIDIDAE) 
ATTACKED  BY  PRIONYX  PARKERI  AND  SOME 

TACHYSPHEX  WASPS  (HYMENOPTERA,  SPHECINAE 
AND  LARRINAE):  A DESCRIPTIVE  STUDY 

By  A.  L.  Steiner 

Department  of  Zoology,  University  of  Alberta 
Edmonton,  Alberta,  Canada,  T6G  2E9 

Introduction 

Predator  and  anti-predator  adaptations,  strategies,  have  been 
studied  extensively  in  recent  years  (see  for  instance  Curio  1976  and 
Edmunds  1974  for  some  recent  reviews).  Problems  of  predator-prey 
coevolution,  mimicry,  protective  coloration  (e.g.,  Cott’s  monu- 
mental work,  1940),  optimal  strategies,  etc.,  have  received  a great 
deal  of  attention.  Defense  mechanisms  are  extremely  diverse  and 
can  even  involve  use  of  a commensal  species  (e.g.  Ross  1971).  A 
variety  of  sensory  channels  can  be  used  such  as  visual  (e.g.  Cott 
1940;  Robinson  1969),  acoustical  (e.g.  Roeder  1965),  chemical  (e.g. 
Eisner  and  Meinwald  1966;  Eisner  1970),  mechanical,  vibratory  (e.g. 
Tautz  and  Markl  1975)  to  mention  only  a few  examples.  Predators 
such  as  mammals,  birds,  reptiles  (e.g.  Curio  1970),  fish,  mollusks 
have  been  extensively  studied. 

Among  insects,  solitary  and  social  wasps  have  also  been  inten- 
sively studied  but  on  the  whole  surprisingly  little  is  known  about  the 
defensive  mechanisms  of  their  “helpless”  prey.  Prey  capture  is  often 
very  difficult  to  observe  and  even  more  so  to  study  extensively  in 
natural  conditions.  The  few  exceptions  mostly  deal  with  prey  that 
represent  a potentially  formidable  opponent  (e.g.  spider,  praying 


*For  part  I see  Steiner  1968  in  the  Literature  Cited. 
Manuscript  received  by  the  editor  May  11,  1981 


2 


Psyche 


[Vol.  88 


mantis,  etc.).  Counter-attacks  by  such  prey  and  occasional  killing  of 
the  predator  have  even  been  reported  (e.g.,  Deleurance  1941,  pp. 
287-288,  for  a praying  mantis  attacked  by  the  sphecid  wasp  Stizus 
distinguendus;  also  1945,  p.  29  for  Tachysphex  costai  Dest.).  Dead 
spider  wasps  have  also  been  found  in  spider  webs  in  natural 
conditions  (pers.  obs.).  Non-predaceous  prey  can  also  exhibit 
defense  reactions,  however,  as  shown  before  for  crickets  attacked  by 
Liris  nigra  wasps  (Steiner  1968). 

The  anti-predator  system  of  acridid  grasshoppers  is  now  de- 
scribed, analyzed,  as  observed  both  in  nature  and  captivity  (sum- 
marized in  Steiner  1976).  The  prey  are:  (1)  mainly  adult  or  subadult 
Oedipodinae,  but  also  a few  Cyrtacanthacridinae,  all  attacked  by 
the  sphecid  wasp  Prionyx  parkeri  Bohart  and  Menke,  (2)  to  a much 
lesser  extent  smaller,  earlier,  instars  preyed  upon  by  Tachysphex 
wasps  (details  in  next  section).  For  the  latter  prey,  defense  reactions 
were  essentially  the  same,  except  for  the  ones  involving  the  wings, 
undeveloped  at  these  stages.  Prey  hunting  and  stinging  by  Prionyx 
parkeri  are  described  in  detail  in  Steiner  1981  (in  press). 

Materials  and  Methods 
Field  observations 

Prionyx  parkeri  wasps  were  observed  mainly  in  the  grassland 
desert  and  adjacent  riparian  habitat  of  S.E.  Arizona,  U.S.A.,  at  the 
foot  of  the  Chiricahua  Mountains,  East  of  Willcox,  during  the 
summer  of  1972. 


Observations  in  captivity 

Individually  marked  Prionyx  parkeri  and  Tachysphex  [mostly 
tarsatus  (Say)]  wasps  were  observed  in  controlled  laboratory  units 
about  60  X 50  X 50  cm  (general  method  described  in  Steiner  1965): 
(1)  at  the  Southwestern  Research  Station,  Portal,  Arizona,  during 
the  sprng  and  part  of  -the  summer  1973  (=  Arizona  study);  (2)  in 
central  Oregon,  U.S.A.,  near  Bend,  using  a field  trailer,  during  the 
summer  of  1977  (=  Oregon  study).  The  following  acridid  grass- 
hoppers taken  from  the  wasps’  habitats  were  used  in  the  Arizona 
study;  (1)  for  P.  parkeri,  adult  or  last  instar  nymphs  of:  Oedipodi- 
nae, mostly  Trimerotropis  pallidipennis  p.  (Burm.),  also  Conozoa 
carinata  Rehn,  a few  Cibolacris  parviceps  (Walker)  — Cyrtacan- 
thacridinae, a few  Psoloessa  delicatula  Scudder  and  an  occasional 


1981] 


Steiner — Anti-predator  Strategies 


3 


Eritettix  variabilis  Bruner;  (2)  for  Tachysphex  wasps,  small  acridid 
nymphs  of:  Oedipodinae,  mostly  Conozoa  carinata  Rehn  and  also 
a few  Trimerotropis  pallidipennis  p.  (Burm.);  Cyrtacanthacridinae, 
a few  Psoloessa  delicatula  and  an  occasional  Melanoplus  sp., 
Derotmema  sp.,  Rehnita  sp.  Rather  similar  but  un-determined 
grasshoppers  were  used  in  the  Oregon  study,  in  captivity.  The  grass- 
hoppers were  provided  either  ad  libitum,  or  in  staged  encounters. 

Observations  were  mostly  continuous,  with  “all  occurrences” 
sampling  of  wasp-prey  interactions.  Precise  quantifications  were 
difficult  or  impossible  because  initial  stages  of  encounters  were 
often  sudden  and  unpredictable.  Generally  speaking  proof  of  effects 
of  escape-defense  reactions  is  often  very  difficult  to  establish  (e.g. 
Edmunds  1974,  p.  240).  This  study  is  basically  descriptive. 

Total  observation  times  were;  (1)  for  captive  P.  parkeri  in  the 
Arizona  study  about  178  h over  a period  of  30  observation  days 
(X  = about  6 h-day)  and  in  the  Oregon  study  about  142  h for  14 
observation  days  (X=  about  6V2  h-day);  (2)  for  captive  Tachysphex 
wasps  in  the  Arizona  study  about  224  h for  37  observation  days  (X  = 
about  6h-day)  and  in  the  Oregon  study  about  224^  h for  35 
observation-day  (X  = 6!4  h-day). 

Results:  Description  of  Responses,  Conditions 

Common  responses:  escape  by  jumping  (flying)  away, 
staying  put  = first  line  of  defense. 

a)  Field  observations 

Visually  hunting  Prionyx  (parkeri?)  wasps  were  observed  in  the 
short  and  sparse  grassy  vegetation,  characteristic  of  the  upper 
Sonoran  desert  grassland.  Acridid  grasshoppers  were  abundant, 
particularly  Oedipodinae  such  as  Mestobregma  plattei  rubripenne 
(Bruner)  adults,  also  found  stored  in  the  nests  of  these  wasps.  The 
most  common  response  to  wasps  approaching  or  pouncing  was  a 
very  sudden,  even  startling,  escape  by  jumping  (Fig.  5a)  and  flying 
away  (Fig.  5b).  The  bright  flash  of  the  colorful  banded  wings  came 
in  sharp  contrast  with  the  sudden  disappearance  from  sight,  after 
landing  (crypticity:  Fig.  5c).  The  wasps  seldom  followed  the  escap- 
ing grasshoppers  in  flight,  but  occasionally  did  so  (Fig.  5b)  and  even 
managed  to  cling  to  them  in  mid  air  and  to  deliver  stings  before 
landing.  Most  stung  grasshoppers  were  apparently  caught  by  sur- 
prise or  at  the  preparatory  stages  of  escape.  Close  range  and 


4 Psyche  [Vol.  88 


Fig.  1:  Attack  of  an  adult  acridid  grasshopper  (Oedipodinae)  by  a Prionyx parkeri 
wasp.  The  wasp  uses  both  the  strong  mandibles  and  long,  powerful  legs,  to  firmly 
hold  the  prey  and  prevent  escape.  The  grasshopper  tries  (in  vain)  to  push  away  the 
wasp  with  both  powerful  hind  legs  by  applying  strong  pressure  on  the  points  where 
the  wasp  is  anchored  (head  and  one  fore  leg).  Several  drops  of  regurgitated  repelling 
fluid  are  indicated  by  arrows.  The  wasp  already  assumes  the  appropriate  posture  for 
the  first  paralyzing  sting,  delivered  in  the  throat  of  the  victim. 


quantitative  observations  were  almost  impossible.  At  times  the 
grasshoppers  stayed  put  instead  of  escaping,  for  no  apparent  reason. 
Attack  of  the  wasp  does  not  necessarily  follow  detection  of  a 
suitable  prey,  however,  since  hunting  wasps  go  through  periods  of 
temporary  refractoriness  (Steiner  1962,  1976,  1978,  1979).  This 
considerably  complicates  the  study  of  possible  effects  of  prey- 
defenses  on  the  wasps. 

b)  Observations  in  captivity 

The  same  responses  were  also  recorded  in  captivity.  Flying  away 


1981] 


Steiner — Anti-predator  Strategies 


5 


Fig.  2:  Regurgitation  of  a repelling  fluid.  An  acridid  grasshopper  (Oedipodinae), 
just  paralyzed  by  a Prionyx  parkeri  wasp,  lies  on  its  back  and  a huge  drop  of  fluid 
covers  a large  surface  of  the  ventral  thoracic  area  where  all  four  stinging  sites  are 
located  (indicated  by  white  dots  and  arrows).  Wasps  often  hesitate  to  dip  their 
abdomen  tip  into  this  viscous,  probably  offensive,  fluid.  Accidental  contact  triggers 
vigorous  body  rubbing  in  an  attempt  to  eliminate  the  unpleasant  fluid  from  the  body 
surface. 


and  long-range  escape  were  impossible,  however,  because  of  space 
limitations. 

There  was  no  evidence  of  active  avoidance  of  Prionyx  or  Tachy- 
sphex  wasps  by  grasshoppers  (“predator  recognition”),  even  after 
repeated  attacks.  Escape  was  always  in  direct  response  to  attack, 
imminent  attack,  or  at  least  sudden  movements  such  as  a wasp 
running  and/or  pouncing.  Thus  predator  and  prey  were  often  seen 
basking  together.  Immediately  following  an  attack,  the  escape 
threshold  was  clearly  lowered,  however. 

Mechanical  defenses  after  contact:  kicking,  pushing  and/or 

brushing  away  the  wasp:  biting;  wing  fluttering  and  flying 
= second  line  of  defense  (Fig.  1) 

After  contact,  Prionyx  wasps  attempt  to  anchor  themselves  to  the 
struggling  or  escaping  grasshopper.  They  try  to  gain  a firm  grip 


6 


Psyche 


[Vol.  88 


using  their  powerful  spinose  legs,  terminal  claws,  and  also  mandi- 
bles. These  wasps  tightly  “embrace”  the  grasshopper,  in  an  anti- 
parallel posture  and  strongly  cling  to  them  (Fig.  1).  In  contrast, 
many  larrine  wasps  (e.g.  Liris,  Tachysphex ) are  comparatively  frail, 
short-legged,  and  cannot  physically  overpower  their  prey  as  success- 
fully as  Prionyx  wasps  do.  Their  prey  often  struggles  free,  in 
contrast  to  Prionyx  prey  which  seldom  succeed,  after  the  “embrac- 
ing” stage,  in  spite  of  frantic  efforts  to  kick  and / or  brush,  push  away 
the  attacker  with  the  powerful  hind  legs.  Prionyx  prey  also  try  to 
deny  free  access  of  the  wasp  to  the  dorsal  side  by  raising  their  long, 
folded,  hind  legs,  often  beyond  the  vertical,  headwards  (hind  leg 
raising:  Fig.  5e).  Powerful  kicks  (Fig.  5e)  sometimes  send  the  wasp  a 
few  cm  from  the  grasshopper,  but  this  works  mostly  before  the  wasp 
can  secure  a firm  grip.  Pushing  action  with  the  tarsi  of  the  powerful 
hind  legs  can  also  be  recorded.  They  are  very  precisely  directed  at 
the  points  seized  by  the  wasp  as  shown  in  Fig.  1 . In  the  latter,  drawn 
from  a photograph,  the  grasshopper  tries,  with  its  right  hind  leg,  to 
push  away  the  left  front  leg  of  the  wasp  while  it  attempts,  with  the 
left  hind  leg,  to  exercise  strong  pressure  on  the  head,  jaws,  of  the 
attacker  and  presumably  get  the  wasp  to  release  its  mandibular  grip 
(in  Fig.  5f  these  “points  of  pressure”  have  been  circled).  Wing 
fluttering  and  even  flying  attempts  can  also  be  observed  in  reponse 
to  the  grasping  action  of  the  wasp.  The  orthopteran  also  performs 
snapping  motions  with  the  jaws  but  is  seldom  able  to  bite  the  wasp. 
The  very  globulous  abdomen  of  Prionyx  wasps  appears  to  be 
especially  well  adapted  to  prevent  such  biting.  The  abdomen  is 
particularly  exposed  since  the  wasp  delivers  the  first  sting  in  the 
throat  of  the  prey,  dangerously  close  to  the  powerful  jaws  (Fig.  5g). 

Chemical  defenses:  regurgitated  fluid  (Fig.  2) 

In  addition  and  often  as  a last  ditch  defense  the  grasshopper 
regurgitates  through  the  mouth  a large  drop  of  dark  fluid  (“tobacco 
juice”)  that  usually  spreads  rapidly  over  the  body  areas  closest  to  the 
mouth,  ventrally,  namely  the  thoracic  surface  (Fig.  2).  This  surface 
sometimes  becomes  completely  covered  with  the  substance.  From 
there  it  can  spread  to  other  body  areas,  if  struggling  is  intense 
enough.  On  Fig.  1 one  drop  can  be  seen  on  the  right  antenna  of  the 
grasshopper  and  one  on  the  tibia  of  the  right  hind  leg  (arrows). 


1981] 


Steiner — Anti-predator  Strategies 


7 


Fig.  3:  Postural  defense  replacing  escape  (startle  and/or  death  feigning  display?). 
The  attacked  grasshopper  froze  into  a hunched  posture,  with  appendages  tucked  in, 
thus  protecting  the  vulnerable  ventral  surface.  The  colorful  wings,  showing  striking 
semi-circular  dark  markings,  are  fully  extended  and/or  flutter  convulsively.  The 
wasp,  after  many  vain  efforts,  managed  to  slip  under  the  grasshopper  (one  leg  is  still 
visible  on  the  right  of  the  grasshopper  head)  and  will  attempt  to  reach  the  vulnerable 
ventral  surface  of  the  thorax  made  less  accessible  by  the  posture  and  interposition  of 
appendages  (obstruction  behavior). 

Uncommon  and  odd  postural  defenses  replacing  escape: 
stationary  wing  flashing  or  extension;  body  arching;  freezing 
(Fig.  3)  = first  line  of  defense. 

a)  Field  observations 

These  rare  occurrences  guarantee  that  such  responses  are  not 
reducible  to  captivity  artifacts. 

The  first  observation  was  made  on  Sept.  4,  1972,  near  the  end  of 
the  morning,  in  the  Arizona  grassland  desert.  One  hunting  Prionux 
(parkeri?)  suddenly  pounced  on  a motionless  grasshopper.  Instead 
of  trying  to  escape,  as  usual,  the  latter  was  seen  with  the  colorful 
wings  open,  fluttering  convulsively,  with  a startling  suddenness, 
thus  producing  a striking  color  flash.  The  hind  legs  were  rigidly 
extended  behind  like  in  the  flying  posture  (Fig.  5b).  However  the 


8 Psyche  [Voi.  88 


Fig.  4:  A Prionyx  parkeri  wasp  succeeded  in  overturning  a “frozen”  oedipodine 
grasshopper.  This  makes  the  ventral  surface  of  the  thorax  more  accessible  to  the 
stings  of  the  wasp.  One  small  drop  of  repellent  fluid  can  be  seen  on  the  abdomen  of 
the  wasp.  After  stinging  is  over,  the  wasp  will  vigorously  rub  its  abdomen  on  the 
substrate,  in  an  effort  to  eliminate  this  unpleasant,  perhaps  noxious,  fluid.  Note  (also 
in  Fig.  2)  the  dot  of  Testor  paint  on  the  dorsal  surface  of  the  wasp  thorax,  for 
individual  identification. 


whole  body  was  strongly  arched  downward  as  in  Fig.  3.  For  the 
observer,  it  looked  as  if  the  “frozen”  grasshopper  was  disabled  or 
dying.  The  wasp  left  the  grasshopper  alone  and  pursued  her  hunting 
trip.  Under  the  impression  that  the  prey  had  received  a sting  or  two, 
I picked  it  up  only  to  see  it  instantly  recover  without  the  slightest 
trace  of  paralysis.  Obviously  the  grasshopper,  later  identified  as  an 
adult  Mestobregma  plattei  rubripenne  (Bruner),  had  not  been  stung 
and  was  not  disabled  at  all.  This  species  is  an  acceptable  prey  since  it 
was  also  found  in  two  nests  dug  up  the  same  day,  nearby.  In 
another,  'Similar,  instance  the  upper  wings  (tegmina)  opened  only 
slightly,  just  enough  to  uncover  the  triangular  base  of  the  vivid  red 
wings  that  remained  folded.  Again  the  wasp  failed  to  paralyze  the 
frozen  grasshopper  which  later  escaped  just  as  suddenly  as  the  first 
one,  unharmed.  The  latter  case  might  be  a less  intense  version  of  the 
first  case.  Presumably  all  gradations  could  be  observed. 


1981] 


Steiner — Anti-predator  Strategies 


9 


The  eliciting  stimuli  of  such  reactions  could  not  be  determined, 
because  of  the  suddenness  and  unpredictability  of  such  encounters. 
Sight  of  the  rapidly  approaching  predator  and/or  mechanical 
contact  are  likely  candidates. 

b)  Observations  in  captivity  (Figs.  3 and  4) 

Similar  or  identical  responses  were  also  observed  in  captivity  at 
close  range  and  in  better  conditions.  Confinement  seemed  to  even 
somehow  favor  appearance  of  this  behavior  perhaps  because  of 
restricted  escape  and/or  greater  concentration  of  attacks.  Often  the 
extended  wings  and  whole  body  were  also  strongly  curved  down- 
wards, sometimes  even  tightly  pressed  against  the  substrate  (Fig.  3). 
The  appendages  and  head  were  tucked  in  and  more  or  less  invisible 
under  the  protective  “umbrella”  of  the  wings.  The  sudden  flash  of 
the  colorful  wings  and  dark  semi-circular  markings,  followed  by  the 
appearance  of  convulsive  movement  and  finally  the  illusion  of  a 
disabled  or  dying  grasshopper  were,  indeed,  an  arresting  sight,  at 
least  for  a human  observer. 

Curiously  such  frozen  grasshoppers  mostly  failed  to  suddenly 
“resuscitate”  and  escape  after  it  had  become  evident  that  their 
postural  defense  had  failed  to  stop  the  wasp  attack.  Such  misfiring 
might  be  a cost  of  this  strategy  because  of  the  strong  inhibitory 
influences  apparently  involved.  Sometimes  wing  fluttering  resumed 
as  the  wasp  attempted  to  deliver  the  paralyzing  stings.  If  left  alone 
by  the  wasp  the  grasshoppers  would  however  invariably  recover 
without  any  sign  of  discomfort,  like  in  the  wild. 

Such  displays  were  never  observed  with  Tachysphex  wasps, 
perhaps  because  the  much  smaller  grasshopper  nymphs  they  attack 
have  undeveloped  wings  . . . that  cannot  be  used. 

If  the  Prionyx  wasps  succeed  in  overcoming  all  these  various 
defense  mechanisms  or  hurdles,  as  they  often  do,  they  then  attempt 
to  deliver  an  average  four  successive  stings,  always  on  the  same 
stinging  sites  and  in  a predictable  order  (summarized  in  Steiner 
1976;  details  in  Steiner  1981).  The  paralyzed  grasshopper  can  then 
be  safely  and  freely  manipulated  and  stored  in  the  nest,  without  any 
resistance,  obstruction. 

Analysis,  Discussion,  Comparisons 

Discussion  is  concerned  mainly  with  possible  or  plausible  inter- 
pretations and  evolutionary  significance  of  these  various  defense 


10 


Psyche 


[Vol.  88 


reactions,  their  degree  of  predator-specificity.  Comparisons  are 
made  with  other  orthopterans,  with  similar  and  different  anti- 
predator strategies.  Effectiveness,  always  difficult  to  prove,  particu- 
larly when  attacks  or  lack  thereof  depend  on  the  internal  state  of  the 
predator  like  in  the  present  case,  will  be  assessed  rather  than 
analyzed  mathematically. 

All  defenses  described  before  (except  crypticity)  are  secondary 
rather  than  primary  defenses  since  they  are  exhibited  during 
encounters  (Edmunds  1974,  pp.  1,  136).  Defenses  are  often  anti- 
location, anti-capture  or  anti-consumption  devices  (i.e.  Alcock 
1975,  p.  333).  Furthermore,  many  species  have  several  lines  of 
defense  (integrated  defense  systems:  Edmunds  1974,  p.  243).  Thus 
the  mantid  Polyspilota  aeruginosa  may  run,  fly,  give  a startle 
display,  slash  at  the  attacker.  It  can  also  feign  death  if  persistently 
handled  in  a rough  way.  It  soon  recovers,  however.  The  brightly 
colored  abdomen  might  also  represent  flash  behavior  (in  Edmunds 
1974,  p.  245).  Each  aspect  of  the  defense  system  will  now  be 
discussed  separately. 

Escape  by  jumping,  flying  away 

This  is  a classical  and  common  case  of  sudden  startling  (flash  or 
deimatic  behavior  Fig.  5b)  followed  by  sudden  disappearance  into 
crypsis  (landing;  Fig.  5c)  (Edmunds  1974,  pp.  146-148)  by  using 
protective  colors  (e.g.  Isely  1938).  This  is  usually  a very  efficient 
mechanism  but  Prionyx  wasps  occasionally  dash  at  flying  grass- 
hoppers (Fig.  5b),  even  sting  them  in  mid  air,  or  take  them  by 
surprise  before  they  can  escape.  Pygmy  mole  crickets  that  escape  by 
flying  away  are  also  grasped  and/or  stung  during  flight  by  the 
sphecid  wasp  Tachytes  mergus  (Yoshimoto,  in  Krombein  and 
Kurczewski  1963,  p.  147)  and  also  by  Tachytes  minutus  (Kurczewski 
1966).  This  defense  is  not  especially  aimed  at  digger  wasp  predators. 

Detection  of  the  predator  is  probably  visual  but  could  also  be 
based  on  hairs  sensitive  to  airborne  vibrations,  as  in  some  caterpil- 
lars such  as  Barathra  brassicae  (Tautz  and  Markl  1978). 

Use  of  hind  legs  other  than  for  jumping:  kicking  or  obstructive 

behavior  such  as  hind  leg  raising  or  interpositions,  brushing 
away,  pushing  away 

Hind  leg  autotomy  used  by  crickets  (Steiner  1968)  was  never 
observed  in  grasshoppers  in  the  present  study  but  Prionyx  wasps 


1981] 


Steiner — Anti-predator  Strategies 


11 


Fig.  5:  Summary  of  oedipodine  grasshopper  anti-predator  actions  and  Prionyx- 
prey  interactions,  a:  the  wasp  detected  a grasshopper  which  is  at  the  preparatory 
stage  of  jumping  (J);  b:  the  prey  flies  away  (F),  suddenly  opening  very  colorful  wings 
with  conspicuous  semi-circular  dark  markings  (startle  display)  and  in  some  cases  the 
wasp  follows  the  grasshopper  in  flight  and  even  stings  it  in  midair;  c:  the  escaping 
prey  suddenly  lands  and  blends  with  the  substrate  (crypsis);  d:  instead  of  escaping  the 
grasshopper  sometimes  “freezes”  into  an  odd  posture  somewhat  remindful  of  an 
inhibited  flying  action;  the  posture  and  convulsive  wing  fluttering  give  the  impression 
that  the  orthopteran  is  disabled,  dying  (disablement  display?  thanatosis?);  at  the  same 
time  the  posture  and  hunching  appear  to  emphasize  the  semi-circular  dark  markings 
on  the  wings  (eyespot  intimidation  display,  “bluff’?);  furthermore  in  this  posture, 
access  of  the  vulnerable  ventral  side  of  the  thorax,  where  stings  are  delivered,  is 
reduced  or  impossible  for  the  wasp  (obstruction  behavior);  e:  hind  leg  raising  ( HLR) 
is  another  obstructive  behavior  that  makes  initial  posturing  of  the  wasp  difficult  or 
impossible;  kicking  can  also  send  the  wasp  a few  cm  away;  f:  hind  legs  are  also  used 
for  brushing  (B)  and/or  pushing  away  (P)  the  wasp;  pressure  is  applied  on  the  circled 
areas  so  as  to  try  to  force  the  wasp  to  release  her  mandibular  and  leg  grip;  g:  as  a last 
ditch  defense  the  grasshopper  can  release  a repellent  fluid  through  the  mouth,  which 
rapidly  spreads  over  the  ventral  thoracic  surface  where  all  stinging  sites  are  located; 
the  wasp  often  hesitates  to  dip  into  this  pool  her  abdomen  tip  (circled);  the  latter  is 
also  exposed  to  powerful  bites  from  the  grasshopper;  therefore  the  throat  of  the  prey 
must  be  quickly  stung  to  stop  these  mouth-based  defenses.  Solid  and  dashed  arrows 
indicate  prey  and  wasp  movements,  respectively;  open  arrows  show  possible 
sequences  of  events  but  these  sequences  can  also  be  broken  if  the  defenses  are 
effective  and  the  wasp  gives  up. 


12 


Psyche 


[Vol.  88 


usually  seize  the  wing  base(s)  or  abdomen  rather  than  one  hind  leg 
(Fig.  1).  Hind  legs  of  crickets,  grasshoppers,  sometimes  phasmids, 
often  covered  with  strong  spines,  are  one  of  their  major  systems  of 
escape  and/or  defense.  Some  wingless  phasmids  can  jab  the  spines 
into  an  aggressor  (Robinson  1968b).  Overt  defense  by  kicking  has 
also  been  described  in  some  large  aphids  (in  Edmunds  1974,  p.  245) 
and  in  a number  of  orthopterans  such  as  crickets  (Steiner  1968)  and 
Locusta  migratoria  for  instance  (Parker  et  al.  1974).  In  the  latter 
case  it  can  be  so  violent  that  the  attacker  is  knocked  20-30  cm  away. 
According  to  Parker  et  al.  (1974)  hind  leg  raising  often  precedes 
kicking  (threat?).  It  is  also  part  of  the  defense  postures  of  male  L. 
migratoria,  the  giant  weta  ( Deinacrida ) of  New  Zealand  (in  Sebeok 
1977,  Fig.  5a,  p.  342)  and  mormon  crickets  when  attacked  by  the 
digger  wasp  Pa/modes  laeviventris  (Parker  and  Mabee  1928,  p.  9). 
In  the  latter  case,  as  in  Prionyx  and  Tachysphex,  the  wasps 
succeeded  in  stinging  only  with  considerable  difficulty.  In  the 
present  study  hind  legs  were  often  raised  past  the  vertical  line  (Fig. 
5e)  and  even  as  far  forward  as  the  level  of  the  head,  as  in  Fig.  1 for 
instance,  in  addition  to  tail  or  body  raising.  This  was  also  observed 
once  in  response  to  an  approaching  Tachysphex  tarsatus.  Freezing 
into  such  postures  made  access  to  the  dorsal  area  and  wasp 
posturing  very  difficult,  sometimes  impossible  (Fig.  5e)  (obstructive 
behavior)  and  the  efficiency  of  this  behavior  appeared  even  to 
increase  as  a result  of  repeated  attacks.  Interposition  of  legs 
(obstruction  behavior)  was  also  observed  in  mole  crickets  attacked 
by  Larra  wasps  (Williams  1928). 

Brushing  and  pushing  away  (Figs.  1 and  5f)  are  more  difficult  to 
evaluate  since  they  are  more  graded  and  variable  responses  which 
are  not  easy  to  detect,  let  alone  quantify,  in  the  confusion  of  the 
attack.  Plausibly  these  responses  work  best  (if  at  all)  at  early  stages 
of  contact  with  the  wasp,  also  if  the  prey  is  very  large  and  vigorous 
or  if  the  wasp  is  more  likely  to  easily  give  up,  for  instance  at  early 
stages  of  hunting  (Steiner  1976).  It  is  doubtful  that  a firmly 
anchored  wasp  can  easily  be  dislodged  in  this  way. 

[Remark:  some  orthopterans  extend  or  raise  their  fore  legs, 
vertically,  as  part  of  a threat-intimidation  posture  (e.g.,  Neobarettia: 
Cohn,  in  Sebeok  1977,  p.  342,  Fig.  5b)]. 

Orthopteran  hind  legs  are  often  given  special  attention  and  are 
paralyzed  first  by  some  predatory  waps  such  as  Liris  and  Tachy- 
sphex (Steiner  1962,  1976).  Prionyx  wasps  can  give  priority  to  the 


1981] 


Steiner — Anti-predator  Strategies 


13 


mouth-based  defenses  (biting,  regurgitating),  now  discussed,  since 
they  effectively  neutralize  hind  leg  defenses  with  their  powerful 
“embracing”  legs.  Correspondingly,  these  wasps  deliver  the  first 
sting  in  the  throat,  not  around  the  hind  legs  (Steiner  1976). 


Biting  and  retaliation  ( aggressive  defense: 

Edmunds  1974,  p.  182) 

Orthopterans  commonly  use  their  powerful  jaws  for  threat, 
intimidation  or  even  active  defense,  retaliation,  if  not  for  predation. 
The  predaceous  North  American  katydid  Neobarettia  severely  bites 
and  displays  the  open  mandibles  as  part  of  the  threat-intimidation 
display  (Cohn,  in  Sebeok  1977,  p.  342,  Fig.  5b). 

In  one  observation  in  captivity  (Arizona,  June  24  1973,  1335  h)  a 
wrongly  positioned  Tacky sphex  tarsatus  (No  + 1042)  was  clearly 
and  severely  bitten  by  a nymph  Trimerotropis  pal/idipennis  p. 
(Burm.)  (No  + 1098)  during  a stinging  attempt.  This  suggests  that 
the  wasp  is  particularly  vulnerable  before  proper  positioning  is 
achieved  and  that  strong  selection  pressures  in  the  direction  of 
minimum  risk  must  have  shaped  the  usual  stinging  postures.  The 
penalty  for  wrong  posturing  can  be  very  heavy.  Thus  the  above 
wasp  was  found  dying  in  the  cage  the  next  day,  June  25,  most  likely 
as  a result  of  this  violent  retaliation  of  the  prey. 

Importance  of  mouth-based  defenses  is  confirmed  by  the  fact  that 
many  orthoptera-hunters  deliver  a special  throat  sting  (Steiner  1962, 
1976)  sometimes  even  before  any  other  sting  (e.g.  Prionyx  parkeri). 
This  also  eliminates  opposition  to  prey-transport  and  storage  in  the 
nest  (and  furthermore  “de-activates”  the  prey  that  recovers  in  part 
from  paralysis,  later:  Steiner  1963a).  In  sharp  contrast,  Oxybelus 
uniglumis  wasps  omit  the  throat  sting  when  they  paralyze  their  non- 
recovering fly-prey  devoid  of  subesophageal  ganglion  and  of  poten- 
tially dangerous  mouth  parts  (Steiner  1978,  1979).  Orthoptera- 
hunting  wasps  with  missing  legparts  or  damaged  antennae  are  often 
found,  particularly  late  in  the  season.  This  might  be  a testimony  to 
the  efficiency  of  bites  of  their  prey  but  also  result  from  intra-specific 
fighting  (see  for  instance  Brockmann  and  Dawkins  1979,  for  Sphex 
ichneumoneus ) and/or  accidents  during  nesting.  A female  Pal- 
modes  carbo  with  two  deep  dents  on  the  back  of  her  abdomen  was 
found  in  southern  British  Columbia.  It  is  probable  that  this 
represented  severe  bites  received  from  one  of  their  large,  often 


14 


Psyche 


[Vol.  88 


predaceous,  decticine  grasshopper-prey  rather  than  beak  marks  of 
some  bird. 


Chemical  defenses:  regurgitated  fluid  (R) 

Chemical  defenses  are  particularly  widespread  among  insects  (see 
for  instance  Eisner  and  Meinwald  1966;  Wallace  and  Blum  1971, 
etc.)  including  Opthopterans.  Some  of  them  have  specialized  glands 
and  the  substance  can  be  ejected  with  considerable  force  (e.g. 
Poekilocerus  buforus,  from  an  opening  located  on  the  first  abdom- 
inal tergite:  Fishelson  1960).  A froth  can  also  be  discharged  through 
a thoracic  spiracle  (e.g.  Romalea  microptera:  in  Eisner  and  Mein- 
wald 1966).  Such  repellents  make  their  owner  distasteful  or  un- 
palatable. The  same  apparently  holds  for  fluids  regurgitated  from 
the  gut  through  the  mouth  (Edmunds  1974,  p.  199)  by  grasshoppers 
for  instance  = enteric  discharges  (Matthews  and  Matthews  1978,  p. 
335).  Digger  wasps,  however,  do  not  consume  their  prey  usually  but 
avoid  contact  with  this  fluid  which  is  apparently  a contact  repellent. 
Functioning  of  the  receptors  located  around  the  stinger  could  be 
impaired  (jamming  effect?)  chemically  and/or  mechanically  (Steiner 
1976).  Stinging  remains  possible,  however,  even  with  stinging  sites 
covered  with  the  fluid  (Figs.  2 and  5b)  but  the  wasp  clearly  hesitates 
or  even  gives  up  half  way  through  stinging.  Contact  triggers 
vigorous,  sometimes  frantic,  rubbing  against  the  ground  and/or 
hyper-grooming  as  in  ants  (Matthews  and  Matthews  1978,  p.  335)  as 
in  hunters  of  regurgitating  caterpillars  like  cutworms  (e.g.,  Am- 
mophila,  Podalonia  wasps).  Body  contact  is  clearly  unpleasant  if 
not  deleterious,  particularly  for  some  small  Tachysphex  wasps 
(Steiner  1976). 

One  of  the  latter  ( tarsatus  No  + 874)  had  her  abdomen  tip  covered 
with  a thick  coat  of  sand  particles  as  a result  of  her  attempts  to  rub 
off  the  sticky  substance.  The  wasp  was  found  dying  the  next  day, 
June  19  (Arizona  study)  (the  same  probably  happened  to  another 
tarsatus  (No  + 887)  which  died  on  June  6). 

The  same  wasp  (No  + 874)  was  also  observed  the  day  before  (June 
18,  1405  h)  in  the  process  of  carefully  removing  with  the  mandibles, 
bit  by  bit,  a large  crust  of  dried  up  fluid,  from  the  ventral  surface  of 
the  thorax  and  throat  of  a grasshopper.  This  was  done  right  after 
“malaxation”  of  the  fore  leg  bases  which  in  some  larrine  wasps  is  a 
preparatory  stage  of  egg-laying  (details  in  Steiner  1971).  Since  the 


1981] 


Steiner — Anti-predator  Strategies 


15 


egg  is  invariably  glued  right  behind  the  fore  legs,  where  the  crust  was 
also  located,  this  would  indicate  that  the  regurgitated  fluid  could 
also  be  a serious  obstacle  to  egg-laying  or  egg  development.  Prionyx 
wasps  lay  their  egg  at  the  base  of  one  hind  leg  . . . where  the  risk  of 
such  “flooding”  is  clearly  much  reduced  or  even  nil!  Furthermore, 
paralyzed  grasshoppers  cannot  remove  the  spilled  fluid  by  groom- 
ing, as  they  normally  do.  Consequently  “cleaning”  of  the  soiled  prey 
can  be  done  only  by  the  wasps,  if  at  all. 

This  chemical  defense  is  apparently  even  more  effective  in  mole 
crickets  against  another  larrine  wasp:  Larra  (Williams  1928).  Thus 
Larra  sanguinea  wasps  were  found  with  their  mouthparts  com- 
pletely glued  together  by  the  very  viscous  fluid.  Remarkably,  some 
of  these  wasps  managed  to  catch  their  mole  cricket  in  spite  of  such 
crippling  handicap!  Ants  are  repelled  by  fecal  material  or  chryso- 
melid  beetle  larvae  (in  Matthews  and  Matthews  1978,  p.  343),  and 
refuse  to  carry  away  pieces  of  grasshopper  treated  with  their  own 
repelling  fluid  (Eisner  1970). 

In  conclusion,  the  importance  of  mouth-based  regurgitative 
defenses  can  be  assessed  by  (1)  the  care  with  which  these  wasps  try 
to  eliminate  the  fluid  from  the  prey  and  from  their  own  body,  (2) 
evolution  of  a specialized  sting  in  the  throat  that  abolishes  mouth- 
based  defenses,  (3)  the  priority  given  by  Prionyx  wasps  to  mouth- 
based  defenses  (first  sting  in  the  throat),  (4)  dramatic  effects, 
including  death,  observed  on  some  wasps  like  small  Tachysphex,  (5) 
toxic  effects  reported  in  the  literature,  for  mammals,  such  as  topical 
irritation  of  eyes,  vomiting  when  swallowed  and  severe  symptoms 
caused  by  injection  (Matthews  and  Matthews  1978,  p.  335). 

Such  defenses  are  therefore  particularly  efficient  against  smaller 
predators  like  arthropods,  wasps  included.  More  experimentation  is 
clearly  needed,  however. 

Postural  defenses,  displays,  replacing  escape 
(Figs.  3,  4 and  5d) 

Such  complex  postures  and  displays  will  be  analyzed  in  terms  of 
their  various  components  or  aspects. 

a)  Color  flash,  startle  response 

Sudden  display  of  colored  wings,  of  hidden  and  bright  structures 
(deimatic  behavior)  is  common  in  insects,  particularly  in  otherwise 
cryptically  colored  moths  such  as  Catocala  scripta,  Triphaena 


16 


Psyche 


[Vol.  88 


pronuba  (in  Edmunds  1974)  and  also  many  orthopterans.  For  the 
latter,  wing  opening  (lifting)  is  for  instance  part  of  the  dramatic 
threat-intimidation  display  of  Neobarettia  already  mentioned  (in 
Sebeok  1977,  p.  342)  or  the  one  of  Phymateus.  Since  these  latter 
species  are  potentially  dangerous  and/or  distasteful  such  displays 
are  usually  interpreted  as  warning  (in  Edmunds  1974,  pp.  148,  154; 
see  also  for  instance  Frazer  and  Rothschild  1962).  The  first  species 
bites  severely  while  the  latter  has  strong  hind  leg  spines  and  secretes 
a repelling  fluid  if  further  molested.  When  exhibited  by  harmless 
species  such  as  the  stick  insect  Metriotes  diocles  (e.g.  Bedford  and 
Chinnick  1966;  Robinson  1968a)  or  common  grasshoppers  it  is 
considered  as  mere  “bluff’  based  on  a startle  effect  and/or  an 
apparent  increase  in  size,  height,  volume,  etc.  (intimidation  beha- 
vior). Similar  actions  are  reported  from  some  cicadas  and  mantids 
and  are  particularly  dramatic  in  the  African  mantid  Idolium 
diabolicum  (in  Wickler  1968). 

b)  Display  of  dark  markings  or  “eyespots” 

Eyespots  are  commonly  displayed  by  moths  (see  for  instance  Blest 
1957,  1964).  If  even  very  imperfect  imitations  are  considered 
effective  then  perhaps  this  also  applies  to  the  semi-circular  dark 
markings  displayed  by  grasshoppers  (Figs.  3 and  5d).  Rarity  of  the 
display  is  essential  (in  Edmunds  1974,  p.  168). 

c)  Appearance  of  disabled,  dying  or  dead  insect  (thanatosis)  with 
freezing,  hunching  and  appendages  tucked  in  (Fig.  5d). 

Inhibition  of  movement  in  itself  or  freezing  is  likely  to  lower  the 
probability  of  detection  and / or  attack  by  predators  that  hunt 
moving  live  prey  visually  (e.g.,  Steiner  1962,  1976  for  cricket- 
hunting Liris  wasps).  This  probably  includes  many  digger  wasps. 
Thanatosis  is  known  from  a number  of  insects,  also  orthopterans 
(Edmunds  1974,  p.  172;  Robinson  1968a).  The  prey  might  also  be 
considered  unsuitable  because  of  the  unusual  appearance  as  such 
(oddity  effects).  The  latter  is  illustrated  by  “protean  defenses”  an 
unpredictable,  erratic  and  highly  diverse  behavior  (in  Edmunds 
1974,  pp.  144-145;  see  also  Chance  and  Russel  1959;  Humphries 
and  Driver  1971,  etc.). 

Furthermore,  grasshoppers  with  wings  spread,  appendages 
tucked  in  and  body  strongly  arched  (Fig.  3)  also  seem  less  exposed 
because  of  reduced  access  to  the  vulnerable  stinging  sites,  all  located 
on  the  well  protected  ventral  surface  of  the  thorax  (Steiner  1981). 


1981] 


Steiner — Anti-predator  Strategies 


17 


Some  Prionyx  wasps  experienced  great  difficulties  in  squeezing 
themselves  under  such  grasshoppers  (Fig.  3,  one  leg  of  the  wasp 
visible).  Sometimes  also  the  wasps  succeeded  in  turning  over  such 
grasshoppers,  venter  up  (Fig.  4).  Even  so,  stinging  was  difficult. 

Reduced  accessibility  might  be  an  accidental  by-product  of  the 
“disablement”  display  or  a more  direct  result  of  wasp-grasshopper 
coevolution.  The  apparent  immunity  of  Acrotylus  grasshopper 
nymphs  to  Tachysphex  peetinipes  was  also  attributed  to  restricted 
accessibility  linked  with  dense  and  long  pilosity  (Ferton  1910,  p. 
158).  Body  arching  has  also  been  observed  on  some  other  orthop- 
terans  and  is  sometimes  associated  with  the  release  or  violent 
expulsion  of  repellent  fluid,  as  in  Poekilocerus  buforus  (Fishelson 
1960). 

d)  “Intimidating”  and  aggressive  defensive  elements  (Fig.  5d). 

If  the  posture  shown  in  Figs.  3 and  5d  is  also  an  eyespot  display 
then  it  has  an  intimidating  as  well  as  “bluff’  value. 

Sideways  rocking,  known  from  some  mantids  (Crane  1952)  and 
also  forward-backward  rocking  were  often  observed  in  crickets,  just 
before  or  after  contact  with  Liris  wasps  (Steiner  1968),  suggesting  an 
intimidating  function.  This  was  also  observed  in  Empusa  egena  in 
response  to  attacks  by  the  sphecid  wasp  Stizus  distinguendus  Handl. 
(Deleurance  1941,  pp.  287-288),  along  with  other  aggressive  re- 
sponses such  as  wings  open,  striking  with  the  raptorial  fore  legs. 
Rocking  was  also  observed  in  some  phasmids  (Crane  1952)  and 
roaches  such  as  Periplaneta  fuliginosa  (Simon  and  Barth  1977,  p. 
307).  Crickets  also  sometimes  froze  into  odd  or  intimidating  erect 
postures  difficult  to  interpret  as  “death  feigning”  (Steiner  1962, 
1968).  Absence  of  stinging  in  such  cases,  if  related  at  all  to  the 
display,  might  depend  on:  (1)  the  oddity  of  the  posture,  as  Chauvin 
and  Chauvin  (1977)  suggest  (the  vertical  posture  is  in  sharp  contrast 
with  the  usual  horizontal  one),  or  (2)  the  possible  intimidating 
effects  associated  with  increased  height  (bluff  behavior),  (3)  preda- 
tor mimicry,  namely  a mantis-like  appearance  (see  Steiner  1968, 
Fig.  i,  p.  267).  [Remark:  this  latter  possibility  was  considered  far- 
fetched by  one  reviewer  of  the  paper  cited  and  consequently 
eliminated  from  the  text. . .and  yet  Simon  and  Barth  (1977,  p.  307, 
Fig.  2)  describe  a somewhat  comparable  rare  posture  from  the  roach 
Periplaneta  fuliginosa  which  they  interpreted  (probably  rightly)  as  a 
“Mantis-threat”!] 


18 


Psyche 


[Vol.  88 


Fig.  6:  Proportions  of  cases  where  negative  effects  on  the  wasp  were  present 
(hatched  bars  = wasp  inhibited  or  stopped,  stinging  incomplete  or  no  stinging  at  all) 
or  not  present  (bars  in  solid  black  = complete  stinging  without  apparent  negative 
effects).  For  cases  where  prey  defenses  were  recorded  (left  pair  of  bars)  the 
proportion  of  negative  effects  is  greater  than  no  effects,  whereas  it  is  the  reverse  for 
cases  where  no  defenses  were  recorded  or  this  information  was  unavailable  (right  pair 
of  bars).  This  indicates  that  prey  defenses  (all  cases  pooled)  do  have  some  negative 
effects  on  the  wasps.  It  is  only  a trend,  however,  since  the  differences  do  not  reach 
significance. 

It  has  been  suggested  that  some  protective  and  intimidating 
displays  (e.g.  in  saturniid  and  sphingid  moths)  could  have  evolved 
from  flight  movements  (Blest  1957)  and  can  be  classified  as  (1) 
rhythmic,  (2)  static,  (3)  mixed  and  (4)  cryptic.  Category  (1),  that 
appears  to  best  fit  the  data  (Figs.  3 and  5d)  would  be  closest  to  the 
original  flight  movements.  Extension  of  hind  legs,  wing  beats,  even 
if  convulsive,  are  clearly  part  of  flying  which  is  strongly  inhibited. 


1981] 


Steiner — Anti-predator  Strategies 


19 


Similar  explanations  would  seem  to  apply  to  the  odd  cricket 
postures  (Steiner  1968)  but  in  the  form  of  “frozen  jumping  and/ or 
kicking”  rather  than  “frozen  flight”  and  reduced  access  to  the 
vulnerable  ventral  stinging  sites  is  also  indicated.  Startle  displays 
have  also  been  interpreted  in  terms  of  conflict  between  flying  and 
freezing  for  some  mantids  (Crane  1952). 

Efficiency  of  such  defenses  has  been  clearly  demonstrated  in  only 
a few  cases.  Parker  et  al.  (1974),  for  instance,  showed  that  defense 
postures  exhibited  by  Locusta  migratoria  had  a significant  negative 
effect  on  bout  continuance  between  conspecifics.  With  wasp  studies 
the  problem  is  further  complicated  by  wide  moment-to-moment 
fluctuations  in  responsiveness  of  the  hunting  wasps  (Steiner  1962, 
1976,  1979).  Such  variables  must  be  controlled,  manipulated  or 
eliminated  to  get  clear  answers  and  this  was  not  done  in  the  present 
study. 


Quantitative  Data 

Quantifications  were  too  limited  and  inappropriate  to  make  a 
statistical  analysis  of  the  effectiveness  of  such  defenses  very  mean- 
ingful. Only  128  cases  were  known  in  sufficient  detail  to  be  included 
in  the  analysis.  In  41 .27%  (n  = 26)  of  the  cases  the  defenses  (lumped 
together)  had  no  apparent  effect  and  complete  stinging  followed  and 
in  58.73%  (n  = 37)  at  least  some  possible  effects  were  recorded,  such 
as  temporary,  permanent,  interruption  or  even  deletion  of  stinging. 
When  no  defenses  were  observed  (or  unknown  status)  the  percent- 
ages of  complete  vs  incomplete  stinging  were  approximately  re- 
versed as  predicted:  55.38%  (n  = 36)  and  44.62%  (n  = 29).  These 
differences  in  proportions  (Fig.  6)  were  not  significant,  however, 
since  the  calculated  x2  was  only  3.689  for  a critical  value  of  5.991 
(p  ^ 0.05;  df  = 2;  G-test  of  independence  of  rows  and  columns: 
Sokal  and  Rohlf  1969,  p.  599).  A slight  advantage  can  have  a 
decisive  selective  value  in  the  long  run,  however. 

Conclusion 

Prey  as  harmless  as  herbivorous  crickets  and  grasshoppers  pos- 
sess a rather  complex,  well  integrated,  system  of  anti-predator 
devices  they  can  use  against  their  wasp  enemies.  Even  if  some  of 
these  responses  are  merely  obstructive,  they  do  in  fact  increase  the 


20 


Psyche 


[Vol.  88 


cost  of  predation  to  the  wasps  by  making  capture  more  difficult, 
more  costly,  and/or  less  probable.  Natural  selection  should  there- 
fore promote  evolution  of  such  anti-predator  strategies  which  in  the 
long  run  increase  the  fitness  of  the  prey. 

Some  components  of  the  system  such  as  flying  away  and  cryptic- 
ity,  perhaps  regurgitation,  are  of  a very  generalized  nature  whereas 
other  devices  are  more  predator-specific.  Thus  startle  displays  with 
exposure  of  dark  semi-circular  markings  are  probably  most  efficient 
against  small  avian  predators,  whereas  biting,  mouth  regurgitation, 
hind  leg  raising  and  obstruction  behaviors  are  presumably  more 
useful  against  smaller,  more  vulnerable  predators  such  as  other 
insects,  including  digger  wasps.  Matthews  and  Matthews  (1978,  p. 
352)  state  that  “protective  adaptations  in  insects  are  intimately 
related  to  the  behavior  and  physiology  of  their  predators.”  This  also 
applies  well  to  wasp  predators. 

Acknowledgements 

The  Arizona  study  was  part  of  a sabbatical  project,  while  on  an 
exchange  program  with  the  American  Museum  of  Natural  History, 
New  York,  in  1972-73.  Research  was  conducted  at  the  Southwest- 
ern Research  Station,  Portal,  Arizona,  and  in  the  surrounding 
areas,  including  the  Chiricahua  National  Monument  and  the  Erick- 
son Ranch.  Help,  advice  and  hospitality  of  many  persons  and 
friends,  I cannot  mention  individually,  are  gratefully  acknowledged. 
Wasp  specimens  were  kindly  identified  by  A.  S.  Menke,  U.S. 
National  Museum  (Entomology),  Washington;  R.  M.  Bohart,  Uni- 
versity of  California,  Davis,  and  W.  J.  Pulawski,  Wroclaw  Univer- 
sity, Poland,  and  grasshopper  specimens  by  D.  C.  Rentz,  the 
Academy  of  Natural  Sciences,  Philadelphia,  Pennsylvania.  The 
study  was  supported  in  part  by  an  operating  grant  (A3499)  from  the 
National  Research  Council  of  Canada  and  funds  from  the  Univer- 
sity of  Alberta,  Edmonton,  Canada.  I would  like  to  thank  J. 
Scheinas  for  typing  the  manuscript. 

Summary 

Harmless  herbivores  such  as  acridid  grasshoppers  exhibit  a 
complex  anti-predator  behavior  when  attacked  by  Prionyx  and 
Tachysphex  sphecid  wasps.  Besides  jumping  and  flying  away  with 


1981] 


Steiner — Anti-predator  Strategies 


21 


exposure  of  colorful  wings  (flash  behavior)  and  sudden  return  to 
crypticity  upon  landing,  these  insects  show  freezing,  often  in  odd 
postures,  with  the  colorful  wings  and  dark  markings  (“eyespots”?) 
prominently  exposed.  Such  postures  also  reduce  access  to  the 
vulnerable  ventral  surface  usually  stung  by  these  wasps  (obstruction 
behavior).  After  contact  with  the  wasp  a second  line  of  defense 
comes  into  effect  such  as  kicking,  brushing  and  pushing  actions.  In 
addition  to  these  hind-leg  based  defenses,  the  attacked  prey  can  also 
use  mouth-based  defenses:  biting  and / or  regurgitating  a repelling, 
perhaps  even  noxious,  fluid  (“tobacco  juice”).  Such  defenses  pre- 
sumably lower  the  probability  of  capture  or  at  least  increase  the  cost 
to  the  predator  and  have  therefore  a selective  value. 


Literature  Cited 

Alcock,  J. 

1975.  Animal  Behavior,  an  evolutionary  approach.  Sunderland,  Mass.:  Si- 
nauer,  547  p. 

Bedford,  G.  O.  and  L.  J.  Chinnick 

1966.  Conspicuous  displays  in  two  species  of  Australian  stick  insects.  Anim. 
Behav.  14:  518-21. 

Blest,  A.  D. 

1957.  The  evolution  of  protective  displays  in  the  Saturnioidea  and  Sphingidae 
(Lepidoptera).  Behaviour  11:  257  309. 

1964.  Protective  display  and  sound  production  in  some  New  World  arctiid  and 
ctenuchid  moths.  Zoologica,  New  York  49:  161-81. 

Brockmann,  H.  J.  and  R.  Dawkins 

1979.  Joint  nesting  in  a digger  wasp,  an  evolutionary  stable  preadaptation  to 
social  life.  Behaviour  71:  203-45. 

Chance,  M.  R.  A.  and  W.  M.  S.  Russell 

1959.  Protean  displays:  a form  of  allaesthetic  behaviour.  Proc.  Zool.  Soc. 
Lond.  132:  65-70. 

Chauvin,  R.  and  B.  Chauvin 

1977.  Le  monde  animal  et  ses  comportements  complexes.  Plon,  Paris. 

Cott,  H.  B. 

1940.  Adaptive  coloration  in  animals.  London,  Methuen. 

Crane,  J. 

1952.  A comparative  study  of  the  innate  defensive  behaviour  in  Trinidad 
mantids  (Orthoptera,  Mantoidea).  Zoologica,  New  York  37:  259-93. 

Curio,  E. 

1970.  Die  Selektion  dreier  Raupenformen  eines  Schwarmers  (Lepidopt., 
Sphingidae)  durch  einen  Anolis  (Rept.  Iguanidae).  Z.  Tierpsychol.  27: 
899  914. 

1976.  The  ethology  of  predation.  Berlin,  Springer  Verlag. 


22 


Psyche 


[Vol.  88 


Deleurance,  E.  P. 

1941.  Contributions  a l’etude  biologique  de  la  Camargue  - Observations 
entomologiques.  Bull.  Mus.  Hist.  Natur.  Marseille  1(4):  275-89. 

1945.  Sur  Fethologie  d’un  Tachytes  chasseur  de  Mantes  Tachysphex  costai 
Dest.  (Hym.  Sphegidae).  Bull.  Mus.  Hist.  Natur.  Marseille  1(1-2): 
25-29. 

Edmunds,  M. 

1974.  Defence  in  Animals.  A survey  of  anti-predator  defences.  New  York, 
Longman. 

Eisner,  T. 

1970.  Chemical  defense  against  predation  in  arthropods.  In  Chemical  Ecology, 
E.  Sondheimer  and  J.  B.  Simeone  (eds.).  New  York,  Academic  Press, 
(pp.  157-215). 

Eisner,  T.  and  J.  Meinwald 

1966.  Defensive  secretions  of  arthropods.  Science  153:  1341-50. 

Ferton,  C. 

1910.  Notes  detachees  sur  l’lnstinct  des  Hymenopteres  melliferes  et  ravisseurs 
(6e  Serie).  Ann.  Soc.  Entom.  Fr.  79:  145-78. 

Fishelson,  L. 

1960.  The  biology  and  behaviour  of  Poekilocerus  bufonius  Klug,  with  special 
reference  to  the  repellent  gland  (Orth.  Acrididae).  Eos,  Madrid  36: 
41-62. 

Frazer,  J.  F.  D.  and  M.  Rothschild 

1962.  Defence  mechanisms  in  warningly-coloured  moths  and  other  insects. 
Internat.  Congr.  Entom.  11,3:  249-56. 

Humphries,  D.  A.  and  P.  M.  Driver 

1971.  Protean  defence  by  prey  animals.  Oecologia  5:  285-302. 

Isely,  F.  B. 

1938.  Survival  value  of  acridian  protective  coloration.  Ecology  19:  370-89. 

Krombein,  K.  V.  and  F.  E.  Kurczewski 

1963.  Biological  notes  on  three  Floridian  wasps.  Proc.  Biol.  Soc.  Wash.  76: 
139-52. 

Kurczewski,  F.  E. 

1966.  Behavioral  notes  on  two  species  of  Tachytes  that  hunt  pygmy  mole- 
crickets  (Hym.:  Sphecidae,  Larrinae).  J.  Kans.  Ent.  Soc.  39:  147-55. 

Matthews,  R.  W.  and  J.  R.  Matthews 

1978.  Insect  behavior.  New  York,  Wiley,  507  p. 

Parker,  G.  A.,  G.  R.  G.  Hayhurst  and  J.  S.  Bradley 

1974.  Attack  and  defence  strategies  in  reproductive  interactions  of  Locusta 
migratoria,  and  their  adaptive  significance.  Z.  Tierpsychol.  34:  1-24. 

Parker,  J.  R.  and  W.  B.  Mabee 

1928.  Montana  insect  pests  for  1927  and  1928.  Univ.  Montana  Agric.  Exper. 
Station,  Bozeman,  Mont.,  Bulletin  216,  p.  9. 

Robinson,  M.  H. 

1968a.  The  defensive  behaviour  of  the  Javanese  stick  insect,  Orxines  macklotti 
De  Haan,  with  a note  on  the  startle  display  of  Metriotes  diocles  Westw. 
(Phasmatodea,  Phasmidae).  Ent.  month.  Mag.  104:  46-54. 

1968b.  The  defensive  behavior  of  the  stick  insect  Oncotophasma  martini 


1981] 


Steiner — Anti-predator  Strategies 


23 


(Griffini)  (Orthoptera:  Phasmatidae).  Proc.  Royal.  Entom.  Soc.  Lond. 
43:  183-7. 

1969.  Defenses  against  visually  hunting  predators.  In:  Evolutionary  Biology  3, 
T.  Dobzhansky,  M.  K.  Hecht  and  W.  C.  Steere  (eds.).  New  York, 
Meredith,  (pp.  225-59). 

Roeder,  K.  D. 

1965.  Moths  and  ultrasound.  Sci.  Amer.  212:  94  102. 

Ross,  D.  M. 

1971.  Protection  of  hermit  crabs  (Dardanus  spp.)  from  Octopus  by  commensal 
sea  anemones  ( Calliactis  spp.).  Nature,  London  230:  401-2. 

Sebeok,  T.  A. 

1977.  How  animals  communicate.  Bloomington,  Indiana  University  Press. 

Simon,  D.  and  R.  H.  Barth 

1977.  Sexual  behavior  in  the  cockroach  genera  Periplaneta  and  Blatta.  III. 
Aggression  and  sexual  behavior.  Z.  Tierpsychol.  44:  305-22. 

SOKAL,  R.  R.  AND  F.  J.  ROHLF 

1969.  Biometry.  San  Francisco,  Freeman,  776  p. 

Steiner,  A.  L. 

1962.  Etude  du  comportement  predateur  d‘un  Hymenoptere  Sphegien:  Liris 
nigra  V.d.L.  (=  Notogonia  pompiliformis  Panz.)  Ann.  Sci.  Natur.  (Zool. 
Biol.  Anim.)  Ser.  12,4:  1-126. 

1963.  Interpretation  neuro-  et  psycho-physiologique  de  l’etat  des  victimes  de 
certaines  Guepes  paralysantes  ( Liris  nigra  V.d.L.  = Notogonia  pompili- 
formis Panz.).  (Contribution  a letude  des  facteurs  motivationnels  et 
activateurs  chez  l’lnsecte).  C.R.  Acad.  Sci.  Ser.  D (Sci.  Nat.)  Paris  257: 
3480-82. 

1965.  Mise  au  point  d’une  technique  d’elevage  d — Hymenopteres  fouisseurs  en 
laboratoire  (Note  preliminaire).  Bull.  Soc.  Entom.  Fr.  70:  12-18. 

1968.  Behavioral  interactions  between  Liris  nigra  V.d.L.  (Hym.  Sphecidae) 
and  Gryllus  domesticus  L.  (Orthoptera:  Gryllidae)  Psyche  75:  256-73. 

1971.  Behavior  of  the  hunting  wasp  Liris  nigra  V.d.L.  (Hym.,  Larrinae)  in 
particular  or  in  unusual  situations.  Can.  J.  Zool.  49:  1401-15. 

1976.  Digger  wasp  predatory  behavior  (Hym.,  Sphecidae).  II.  Comparative 
study  of  closely  related  wasps  (Larrinae:  Liris  nigra,  Palearctic;  L. 
argentata  and  L.  aequalis,  Nearctic)  that  all  paralyze  crickets  (Orthop- 
tera, Gryllidae).  Z.  Tierpsychol.  42:  343-80. 

1978.  Evolution  of  prey-carrying  mechanisms  in  digger  wasps:  possible  role  of 
a functional  link  between  prey-paralyzing  and  carrying  studied  in 
Oxybelus  uniglumis  (Hym.,  Sphecidae,  Crabroninae).  Quaest.  ent.  14: 
393-409. 

1979.  Digger  wasp  predatory  behavior  (Hym.,  Sphecidae):  fly  hunting  and 
capture  by  Oxybelus  uniglumis  (Crabroninae:  Oxybelini);  a case  of 
extremely  concentrated  stinging  pattern  and  prey  nervous  system.  Can. 
J.  Zool.  57:  953-62. 

1981.  Digger  wasp  predatory  behavior  (Hym.,  Sphecidae).  IV.  Comparative 
study  of  some  distantly  related  Orthoptera-  hunting  wasps  (Sphecinae 
vs.  Larrinae),  with  emphasis  on  Prionyx  parkeri  (Sphecini).  Z.  Tier- 
psychol. (in  press). 


24 


Psyche 


[Vol.  88 


Tautz,  J.  and  J.  Markl 

1978.  Caterpillars  detect  flying  wasps  by  hairs  sensitive  to  airborne  vibration. 
Behav.  Ecol.  Sociobiol.  4:  101-10. 

Wallace,  J.  B.  and  M.  S.  Blum 

1971.  Reflex  bleeding:  a highly  refined  defence  mechanism  in  Diabrotica 
larvae  (Coleoptera:  Chrysomelidae).  Ann.  Ent.  Soc.  Amer.  64:  1021-4. 
Wickler,  W. 

1968.  Mimicry  in  plants  and  animals.  London,  Weidenfeld  and  Nicholson. 
Williams,  F.  X. 

1928.  Studies  in  tropical  wasps  — their  hosts  and  associates  (with  descriptions 
of  new  species).  Bull.  Exper.  Station  Hawaii.  Sugar  Planter’s  Assoc. 
(Entom.)  (19):  1-179. 


A COMPARISON  OF  THE  NEST  PHENOLOGIES  OF 
THREE  SPECIES  OF  POGONOMYRMEX  HARVESTER 
ANTS  (HYMENOPTERA:  FORMICIDAE)* 


By  William  P.  MacKay 
Departamento  de  Entomologia 
Colegio  de  Graduados 
Escuela  Superior  de  Agricultura 
Ciudad  Juarez,  Chih.  Mexico 

Introduction 

Ants  are  among  the  most  abundant  animals  in  most  habitats 
(Petal  1967)  and  may  even  be  the  dominant  insects  in  many 
ecosystems  (Nielsen  1972;  Nielsen  and  Jensen  1975).  Harvester  ants 
of  the  genus  Pogonomyrme x are  a major  component  of  the  energy 
flux  through  ecosystems  (Golley  and  Gentry  1964).  Ants  of  this 
genus  have  become  increasingly  important  in  ecological  studies, 
including  mutualism  (O’Dowd  and  Hay  1980),  competition  (Mares 
and  Rosenzweig  1978;  Reichman  1979;  Davidson  1980),  predation 
(Whitford  and  Bryant  1979),  foraging  (Whitford  and  Ettershank 
1975;  Holldobler  1976a;  Whitford  1976,  1978a;  Davidson  1977a,  b; 
Taylor  1977),  community  structure  (Davidson  1977a,  b;  Whitford 
1978b),  and  impact  on  ecosystems  (Clark  and  Comanor  1975; 
Reichman  1979).  It  is  difficult  to  investigate  harvester  ants  as 
seasonal  processes  occurring  inside  the  nest  are  generally  unknown 
and  the  nest  populations  are  usually  underestimated. 

This  investigation  compares  the  nest  phenologies  of  three  species 
of  Pogonomyrmex  harvester  ants:  P.  montanus  MacKay,  P. 
subnitidus  Emery,  and  P.  rugosits* Emery,  which  occur  at  high,  mid, 
and  low  altitudes  respectively.  These  data  form  the  basis  for  a 
comparison  of  the  ecological  energetics  of  the  three  species 
(MacKay  1981). 

Materials  and  Methods 
The  species  investigated. 

The  altitudinal  comparison  is  based  on  three  species  of  harvester 

*This  research  constitutes  Chapter  3 of  a dissertaion  submitted  to  the  faculty  of  the 
University  of  California,  Riverside,  in  partial  fulfillment  of  the  requirements  for  the 
Degree  of  Ph.D.  in  Population  Biology. 

Manuscript  received  by  the  editor  May  28,  1981. 


25 


26 


Psyche 


[Vol.  88 


ants:  Pogonomyrmex  montanus  MacKay,  P.  subnitidus  Emery,  and 
P.  rugosus  Emery.  All  three  belong  to  the  subgenus  Pogonomyrmex. 
Pogonomyrmex  subnitidus  and  P.  montanus  are  very  closely 
related,  both  belong  to  the  occidentalis  complex  (MacKay  1980). 
Pogonomyrmex  rugosus  belongs  to  the  barbatus  complex  (Cole 
1968).  Pogonomyrmex  montanus  is  unusual  for  the  genus  in  being  a 
high  mountain  species  occurring  in  pine  forests  in  the  mountains  of 
southern  California.  Pogonomyrmex  subnitidus  is  a mid-altitude 
species  in  the  San  Jacinto  Mountains.  Pogonomyrmex  subnitidus  is 
distributed  throughout  southern  and  south  central  California  and 
Baja  California,  occurring  at  lower  elevations  throughout  much  of 
its  range.  Pogonomyrmex  subnitidus  is  sympatric  with  P.  rugosus  in 
parts  of  Riverside  County,  but  is  uncommon  in  such  areas. 
Pogonomyrmex  rugosus  is  a low  altitude  species  near  Riverside  and 
occurs  at  lower  elevations  throughout  much  of  southwestern  United 
States.  It  rarely  occurs  at  higher  elevations.  For  example,  in  the 
Joshua  Tree  National  Monument  it  is  present  up  to  1350  meters,  in 
New  Mexico  it  occurs  at  over  2100  meters. 

Study  areas. 

Populations  of  all  three  species  were  studied  in  southern  Cali- 
fornia: P.  montanus — in  a yellow  pine  forest  community  between 
Fawnskin  and  Big  Pine  Flat  at  2100  meters  elevation  in  the  San 
Bernardino  Mountains  of  San  Bernardino  Co.,  P.  subnitidus — in 
the  chaparral  near  the  Vista  Grande  Ranger  Station  at  1500  meters 
in  the  San  Jacinto  Mountains  of  Riverside  Co.,  P.  rugosus — in  the 
coastal  sage  scrub  community  at  Box  Springs  at  300  meters  near 
Riverside,  Riverside  Co.  The  three  species  occur  in  clearings  within 
these  different  plant  communities. 

Estimation  of  nest  populations. 

Two  primary  methods  are  used  in  the  estimation  of  ant  nest 
populations:  mark-recapture  methods  and  nest  excavation.  Mark- 
recapture  methods  are  used  to  compare  a population  before  and 
after  seasonal  production.  This  method  has  been  criticized  as  one  of 
the  assumptions  is  that  workers  mix  randomly  in  the  nest.  The 
workers  of  all  three  species  are  stratified  within  the  nests  and  there  is 
strong  evidence  that  other  species  are  stratified  as  well  (MacKay 
1981).  Also  I could  find  no  reliable  way  to  mark  the  individuals  such 
that  the  marks  were  permanent,  could  not  be  passed  on  to  other 
individuals,  and  would  not  disrupt  normal  activities.  In  any  case, 


1981]  Mac  Kay— Nest  Phenologies  of  Pogonomyrmex 


27 


such  a method  would  only  estimate  the  numbers  of  foragers  in  a 
Pogonomyrmex  nest,  not  the  actual  nest  population.  In  addition, 
mark-recapture  methods  do  not  provide  an  estimate  of  the  repro- 
ductives  produced  in  a nest. 

Excavation  of  nests  destroys  them  for  further  study  and  requires  a 
large  expenditure  of  time  and  effort.  I chose  periodic  nest  excava- 
tion as  the  method  of  estimating  production  as  counts  of  the 
sexuals,  brood,  and  workers  can  be  made. 

Our  experience  indicates  that  most  of  the  nest  population  is 
collected.  Pogonomyrmex  spp.  colonies  may  live  15  to  20  years 
(Barnes  and  Nearney  1953),  and  will  live  at  least  two  years  after  the 
removal  of  the  queen  (pers.  obs.).  Nest  longevity  is  unknown  in  the 
three  species  investigated,  but  based  on  data  from  other  species,  I 
expect  at  least  5%rl0%  of  the  nests  should  not  have  queens.  The 
high  proportion  of  nest  queens  collected  (84%  in  P.  montanus,  77% 
in  P.  subnitidus,  and  80%  in  P.  rugosus ) supports  the  hypothesis 
that  most  of  the  nest  population  is  collected.  The  queens  do  not 
reside  in  any  special  “queen  chamber”  and  are  of  a similar  size  as  a 
worker.  Therefore,  it  is  not  any  easier  to  find  the  queen  than  it  is  to 
find  any  individual  worker  in  the  nest.  In  all  cases  excavation  was 
continued  at  least  50  cm  deeper  than  the  position  of  the  last  ant 
found  or  the  end  of  a burrow. 

Nest  excavation  procedure. 

The  procedure  was  as  follows:  The  surface  dimensions  of  the  nest 
were  determined  by  removal  of  the  top  10  cm  of  the  nest.  The  hole 
was  then  extended  at  least  50  cm  on  all  sides.  A square  ditch  was 
dug  around  the  perimeter  of  the  nest  to  a depth  of  one  meter  in  the 
case  of  P.  montanus  nests  and  over  1.5  meters  around  the  nests  of  P. 
rugosus  and  P.  subnitidus.  We  were  able  to  sit  in  the  ditches  while 
carefully  excavating  the  nests  in  10  cm  levels.  As  the  hole  became 
deeper,  the  ditches  were  proportionally  deepened.  All  of  the 
contents  of  the  burrows,  including  ants,  brood,  guests,  stored  seeds, 
and  dirt  were  placed  in  labeled  half  or  one  liter  plastic  containers. 
Later  the  animals  were  separated  from  the  dirt,  and  counted.  Nest 
excavation  usually  began  between  06:00  and  07:00,  before  the  ants 
became  active.  If  foragers  were  needed  for  other  investigations, 
excavation  began  later  in  the  morning  or  early  in  the  afternoon. 
Excavation  and  counting  of  a P.  montanus  nest  requires  6-10  hours, 
of  a P.  subnitidus  nest  20-30  hours  and  of  a P.  rugosus  nest  60-90 


28 


Psyche 


[Vol.  88 


hours.  Whenever  excavation  was  stopped  to  be  continued  on  the 
following  day,  the  nest  was  covered  with  a heavy  vinyl  cloth  and  10 
cm  deep  layer  of  dirt.  This  was  necessary  to  keep  the  inhabitants, 
especially  the  males,  in  the  nest.  A total  of  80  P.  montanus,  26  P. 
subnitidus,  and  20  P.  rugosus  nests  were  completely  excavated 
between  1977  and  1980. 

It  appeared  that  the  excavation  procedure  disrupted  stratification 
of  individuals  within  the  nest  only  slightly.  When  nest  chambers 
were  exposed,  many  individuals  emerged,  but  most  of  the  popula- 
tion remained  in  the  chambers,  and  assumed  a defensive  position 
involving  opening  of  the  mandibles  and  forward  extension  of  the 
antennae. 

The  numbers  of  workers  at  each  level  and  the  position  of  the 
queen  were  recorded.  When  the  nests  were  in  production,  the 
presence  or  absence  of  eggs  was  noted,  but  the  eggs  were  not 
counted,  as  they  were  extremely  small  and  are  easily  lost  in  the  dirt. 
The  larvae,  pupae,  females,  males,  and  callows  (immature,  under- 
pigmented  workers)  were  counted  when  they  were  present  in  the 
nests.  The  contents  of  each  level  were  summed  to  obtain  an  estimate 
of  the  entire  nest  population. 

Seed  storage  in  nests. 

The  seeds  were  separated  from  the  soil  by  filling  a 1000  ml  beaker 
about  full  of  soil  and  seeds.  The  contents  were  washed  into  a sieve 
with  0.5  mm  mesh.  The  washing  and  swirling  were  continued  until 
all  of  the  seeds  were  removed  from  the  soil.  The  material  caught  in 
the  sieve  was  washed  again  until  only  seeds  remained  in  the  sieve. 
The  seeds  were  then  dried  (60°  C)  to  constant  weight. 

Nest  structure. 

In  the  process  of  nest  excavation  it  was  noted  that  the  general 
form  and  shape  of  the  nests  were  comparable  in  all  three  species. 
The  P.  montanus  nest  structure  was  studied  by  pouring  a thin 
solution  of  plaster  of  Paris  (3  tablespoons /liter  of  water)  into  one 
nest.  The  solution  was  dilute  enough  that  the  walls  of  most  of  the 
tunnel  system  were  coated  with  plaster.  The  nest  was  excavated  in 
1-2  cm  layers  and  the  tunnel  structure  at  each  layer  was  measured 
and  sketched.  The  resulting  series  of  “cross  sections”  of  the  nest 
resulted  in  a composite  drawing  of  the  nest. 

Nest  temperature  and  humidity. 

Temperature  data  were  recorded  from  approximately  weekly 


1981] 


Mac  Kay — Nest  Phenologies  of  Pogonomyrmex 


29 


readings  of  thermisters  permanently  implanted  in  nests  of  the  three 
species.  The  data  were  supplemented  with  readings  taken  during 
nest  excavation,  following  the  procedure  of  Rogers  et  al.  (1972). 
Soil  temperatures  taken  within  the  excavation  hole  (at  least  20  cm 
distant  from  ant  burrows)  and  within  the  adjacent  undisturbed  soil 
at  the  same  level  were  not  significantly  different  in  two  cases 
involving  P.  montanus  nests  (F  = 0.00001ns,  F = 0.13ns).  Similar 
comparisons  were  not  made  in  the  cases  of  P.  rugosus  and  P. 
subnitidus  as  the  soils  were  too  compacted  to  allow  the  insertion  of 
a thermometer  in  undisturbed  soil  to  a depth  of  30  or  40  cm. 

Soil  samples  (160  grams)  were  collected  at  various  depths  and 
oven  dried  (60°  C)  to  constant  weight  to  determine  water  content.  At 
least  three  replicates  of  soil  temperature  and  soil  moisture  content 
were  collected  at  each  level.  It  was  anticipated  that  these  parameters 
would  determine  the  position  of  the  brood  within  the  nest.  I 
assumed  a correlation  existed  between  the  humidity  within  the 
burrows  and  water  content  of  the  soil  as  well  as  a uniformity  of  the 
soil  structure  in  the  first  100  cm  of  the  nest  where  most  of  the 
seasonal  changes  in  the  positions  of  the  inhabitants  occurred.  Sandy 
soils  would  release  more  water  vapor  to  burrows  than  would  clay 
soils,  if  both  had  the  same  level  of  soil  moisture  (Marshall  and 
Holmes  1979).  The  amount  of  water  present  within  the  soil  changes 
continuously  under  field  conditions  (Marshall  1959),  which  would 
also  modify  the  relative  humidity. 

Food  input  into  nest. 

Food  input  was  estimated  by  channeling  the  flow  of  foragers  and 
sampling  a fraction  of  foragers  at  regular  intervals  to  determine  the 
numbers  of  trips  made  and  the  amount  of  food  brought  back  to  the 
nest. 

Twenty-eight  nests  of  the  three  harvester  ant  species  (13  P. 
montanus,  10  P.  subnitidus,  and  5 P.  rugosus ),  were  surrounded  by 
strips  of  25  gauge  sheet  metal.  The  diameters  of  the  enclosures  were 
approximately  one  meter  for  P.  montanus,  1.5  meters  for  P. 
subnitidus,  and  2 meters  for  P.  rugosus.  The  sheet  metal  strips  were 
buried  to  a depth  such  that  10  cm  of  the  metal  were  exposed.  Sheet 
metal  with  a total  width  of  20  cm  was  sufficient.  The  ants  could  not 
normally  climb  over  the  enclosure  as  the  sheet  metal  was  very 
smooth.  The  ants  would  occasionally  begin  to  climb  the  enclosure  at 
the  junction  of  the  two  ends.  In  such  cases  the  area  was  covered  with 
Tanglefoot(R). 


30 


Psyche 


[Vol.  88 


In  some  cases,  especially  with  P.  montanus,  the  ants  would 
attempt  to  tunnel  under  the  enclosure.  When  this  occurred,  the  ants 
were  removed  from  the  site  of  the  tunneling  and  placed  near  the  nest 
entrance  inside  the  enclosure.  In  such  cases  the  tunneling  was 
completely  controlled  by  destroying  the  tunnel  system  and  replacing 
it  with  soil. 

The  ants  were  allowed  to  enter  and  exit  the  colony  through  two  2 
cm  diameter  vinyl  tubes,  6 cm  in  length.  Entrance  of  the  ants  to  the 
colony  through  the  “exit”  tube  was  prevented  by  having  a 0.5-1  cm 
distance  between  the  end  of  the  tube  and  the  soil.  In  a similar 
manner  exit  via  the  “entrance”  tube  was  prevented.  The  ants  were- 
apparently  not  affected  by  this  short  distance,  they  either  simply 
dropped  with  no  hesitation  or  rapidly  climbed  down  from  the  tube 
to  the  soil.  The  tubes  were  within  15  cm  of  each  other  and  were 
placed  on  the  side  of  the  nest  where  most  of  the  foraging  occurred. 
A 0.448  liter  glass  jar  could  be  placed  under  the  tube  by  which  the 
ants  entered  the  nest,  thus  collecting  the  foragers  with  the  food  items 
they  carried.  The  foragers  were  counted  and  the  food  items 
collected.  The  foragers  were  released  into  the  nest  enclosure  with  a 
quantity  of  food  (native  seeds)  which  approximated  the  amount  of 
food  removed.  The  nests  were  sampled  at  approximately  weekly 
intervals  throughout  the  foraging  seasons,  during  1978  to  1980.  All 
of  the  foragers  entering  P.  montanus  nests  were  collected,  1/5  to  1/6 
of  those  entering  P.  subnitidus  nests,  and  1 / 60  of  those  entering  the 
P.  rugosus  nests.  With  these  proportions,  one  person  could  handle 
the  activity  of  5 nests  during  a single  day.  The  forager  populations 
were  estimated  by  capturing  all  of  the  foragers  throughout  the  day, 
as  they  returned  to  the  nests. 

Statistical  analysis. 

Unless  otherwise  indicated,  the  5%  level  of  significance  was  used 
in  all  comparisons.  A single  asterisk  indicates  statistical  significance 
at  the  5%  level,  double  asterisks  at  the  1%  level,  triple  asterisks 
indicate  significance  at  the  0.1%  level.  Means  are  listed  plus  or 
minus  one  standard  error.  The  percentages  of  the  nest  populations 
were  used  to  make  comparisons  between  the  species  possible.  The 
data  obtained  were  fit  to  least  squares  polynomial  regressions 
(Snedecor  and  Cochran  1967).  The  curves  were  constructed  from 
the  equations. 


1981]  Mac  Kay — Nest  Phenologies  of  Pogonomyrmex 


31 


Figure  1.  The  structure  of  a typical  Pogonomyrmex  montanus  nest. 


CENTIMETERS 


32 


Psyche 


[Vol.  88 


Results 

Nest  structure. 

The  nest  of  P.  montanus  has  numerous  burrows  in  the  upper 
levels  (Figure  1).  Below  this,  there  is  often  only  a single  main  tunnel 
to  the  bottom  of  the  nest.  Most  of  the  ants  are  found  in  the  burrows 
which  branch  from  the  main  tunnel.  The  main  tunnel  contains  few 
ants  and  is  apparently  used  only  for  movement  between  the  side 
burrows.  In  many  cases  there  are  two  separate  “major  tunnels”,  as  is 
shown  in  Figure  1.  In  P.  subnitidus  the  two  major  tunnels  may  be 
separated  by  more  than  100  cm  and  may  appear  as  two  separate 
nests.  One  major  tunnel  may  contain  no  brood  and  the  other  may 
contain  all  of  the  brood  in  the  nest.  The  queen  and  brood  are  usually 
found  in  the  major  tunnel  which  goes  to  the  deeper  level. 

The  structure  of  the  nests  of  P.  subnitidus  and  P.  rugosus  are  not 
shown,  but  are  similar  except  that  they  are  larger  and  deeper,  often 
extending  to  300  or  400  cm  deep.  There  was  no  relationship  between 
the  worker  populations  and  the  nest  depth  (for  P.  montanus  r = 
0.16ns  (65),  for  P.  subnitidus  r = 0.03ns  (26),  and  for  P.  rugosus  r = 
0.32ns  (20)). 

Nest  microclimatology:  temperature. 

The  seasonal  changes  in  nest  temperatures  are  similar  for  all  three 
species  (Figure  2).  The  nest  warms  rapidly  in  the  spring  and 
temperatures  reach  a maximum  at  the  end  of  June  or  July.  The  soil 
temperature  begins  to  drop  in  August  and  levels  out  during  the 
winter  months.  As  the  species  occur  at  different  altitudes,  the 
temperature  ranges  are  different.  The  range  of  P.  montanus  extends 
from  slightly  below  zero  to  20°  C,  that  of  P.  subnitidus  from  slightly 
above  zero  to  25°  C,  and  that  of  P.  rugosus  from  slightly  below  10 
to  30°  C. 

Only  the  changes  at  the  20  and  50  cm  depths  are  shown  in  Figure 
2 as  the  other  levels  are  similar.  The  differences  between  the  levels 
deeper  than  40  cm  were  generally  not  significant.  The  only 
important  difference  between  the  curves  of  the  20  cm  level  and  50 
cm  level  is  that  the  shallow  level  warmed  sooner  in  the  spring  and 
cooled  sooner  in  the  fall. 

Nest  microclimatology:  humidities. 

The  seasonal  changes  in  soil  moisture  are  similar  in  the  nests  of  all 
three  species  (Figure  3).  Soil  moistures  are  high  in  the  winter  and 


1981]  Mac  Kay — Nest  Phenologies  of  Pogonomyrmex 


33 


MONTHS 

Figure  2.  Seasonal  changes  in  the  mean  daily  nest  temperatures  of  three  species 
of  Pogonomyrmex  harvester  ants. 


Free  water  in  soil  (%)  Free  water  in  soij_(%) 


34 


Psyche 


[Vol.  88 


Figure  3.  Seasonal  changes  in  the  nest  humidities  of  three  species  of 
Pogonomyrmex  harvester  ants. 


spring  and  low  in  the  summer  and  fall.  Throughout  the  winter,  the 
soils  receive  relatively  large  amounts  of  rain  or  snow  which  raise  the 
soil  moistures  to  high  levels.  After  this  time,  the  surface  and  upper 
levels  lose  water  rapidly  by  evaporation.  The  lower  levels  of  the  nest 
retain  water  throughout  the  entire  season,  although  the  percentage 
decreases.  Soil  moistures  at  levels  below  30  cm  are  essentially  the 
same  for  all  three  species.  Summer  showers  rapidly  increase  soil 
moistures  of  the  upper  levels  (note  the  peaks  in  the  Figure  3),  but 
have  little  effect  on  the  levels  below  30  cm.  This  water  input  into  the 
soil  is  rapidly  lost  by  evaporation. 

The  soil  moisture  of  the  lower  levels  is  generally  higher  than  that 
of  the  upper  levels,  possibly  forming  a relative  humidity  gradient. 
There  are  more  fluctuations  in  the  higher  levels,  both  in  soil 


1981]  Mac  Kay — Nest  Phenologies  of  Pogonomyrmex 


35 


moisture  and  temperature.  This  probably  accounts  for  much  of  the 
brood  being  kept  in  the  lower  nests  levels. 

The  harvester  ants  apparently  obtain  water  from  several  sources. 
Some  metabolic  water  may  be  available  to  the  ants,  as  it  has  been 
shown  that  harvester  ants  increase  their  metabolism  when  they  are 
water  stressed,  without  increasing  their  activity  (Ettershank  and 
Whitford  1973;  Kay  and  Whitford  1975).  Morning  dew  would  not 
normally  be  available  as  foraging  begins  after  dew  has  evaporated.  I 
have  seen  harvester  ants  actively  drink  rain  drops  on  the  soil 
surface,  demonstrating  a curious  pumping  action  of  the  gaster,  but 
precipitation  is  not  common  in  the  three  habitats  during  the  summer 
(U.S.  Weather  Bureau  Climatological  Data).  Capillary  condensa- 
tion occurs  in  the  soil  at  relative  humidities  above  eighty  percent 
(Rode  1955)  and  may  allow  the  ants  free  water.  Arthropods, 
especially  insects,  are  able  to  actively  absorb  water  vapor  from 
unsaturated  air,  although  the  mechanism  is  not  understood  (Edney 
1974;  Cloudsley-Thompson  1975).  It  is  not  known  if  harvester  ants 
have  the  ability  to  actively  absorb  water  vapor. 

Seasonal  changes  in  nest  populations. 

The  data  on  nest  populations  obtained  from  the  nest  excavations 
are  summarized  in  Appendix  1.  Absolute  counts  could  not  be  easily 
compared  because  the  numbers  of  individuals  present  in  the  nests  of 
the  three  species  are  very  different.  To  reduce  this  variation  between 
nest  populations  of  the  three  species,  the  data  are  compared  in  the 
form  of  percentages.  The  seasonal  changes  in  the  brood  and  sexual 
populations  are  similar  for  all  three  species,  when  the  percentage 
composition  of  each  of  the  classes  are  compared  (Figs.  4 & 5).  In  the 
three  species,  egg  laying  begins  in  late  April  to  late  May,  similar  to 
P.  owyheei  (Willard  and  Crowell  1965)  and  P.  occidentalis  (Lavigne 
1969).  Development  from  egg  to  callow  in  the  species  requires  five 
to  six  weeks  compared  to  25  days  for  P.  badius  (Gentry  1974)  and  30 
days  in  P.  occidentalis  (Cole  1934).  It  is  very  difficult  to  determine 
the  number  of  larval  instars  in  the  development  of  ants  (Wheeler 
and  Wheeler  1976),  although  Marcus  (1953)  suggests  that  there  are 
four  instars  in  P.  marcusi.  As  a consequence,  all  of  the  instars  were 
combined  into  a single  group.  The  first  larvae  appear  about  a week 
after  the  eggs  are  laid,  first  pupae  about  two  weeks  later.  Callows 
are  found  in  the  nest  about  5 or  6 weeks  after  the  eggs  were  laid  and 


36 


Psyche 


[Vol.  88 


MONTHS 


MONTHS 


MONTHS 


Figure  4.  Seasonal  changes  in  the  brood  populations  of  three  species  of 
Pogonomyrmex  harvester  ants.  The  arrows  indicate  the  dates  when  eggs  were  first 
found  in  the  nests.  Nests  excavated  which  contained  only  adult  workers  are  not 
represented  in  the  figure. 


1981]  Mac  Kay — Nest  Phenologies  of  Pogonomyrmex 


37 


remain  pale  for  about  three  weeks.  Thus,  development  from  the  egg 
through  the  larval  instars  requires  about  three  weeks,  the  pupal 
stage  2 3 weeks,  and  the  callow  stage  three  weeks. 

Most  of  the  eggs  are  laid  in  the  spring  as  large  amounts  are  found 
early  in  the  season.  The  amounts  found  in  later  excavations  decrease 
and  eggs  are  rarely  found  after  the  pupae  begin  to  appear  in  the  nest. 

The  larval  population  reaches  a maximum  in  late  July  in  P. 
montanus,  and  mid  August  in  P.  subnitidus  and  P.  rugosus.  The 
pupal  population  reaches  a maximum  in  mid  August  in  P. 
montanus  and  late  August  in  P.  subnitidus  and  P.  rugosus.  The 
callow  population  reaches  a maximum  in  early  to  mid  September  in 
all  three  species.  The  callows  are  easy  to  distinguish  from  adult 
workers  in  P.  montanus  as  they  remain  pale  for  at  least  three  weeks 
(based  on  laboratory  observations).  The  callows  of  P.  rugosus  and 
P.  subnitidus  are  much  more  difficult  to  distinguish  from  the  adult 
workers.  Pogonomyrmex  rugosus  callows  darken  to  a color  indis- 
tinguishable from  mature  workers  within  five  days.  Pogonomyrmex 
subnitidus  mature  workers  are  pale  making  it  difficult  to  distinguish 
them  from  the  callows,  even  if  the  callows  remain  pale  for  many 
days. 

As  the  majority  of  the  first  individuals  produced  are  sexuals,  most 
of  the  larvae  and  pupae  formed  in  the  first  part  of  the  season 
become  reproductives.  Workers  are  also  produced  early  in  the 
season,  especially  in  P.  rugosus.  All  of  the  later  brood  become 
workers  as  was  also  found  in  P.  owyheei  (Willard  and  Crowell 
1965).  The  reproductives  remain  in  the  nest  only  until  late  August  or 
early  September.  In  P.  owyheei  they  remain  in  the  nests  until  mid 
December  (Willard  and  Crowell  1965). 

The  first  winged  reproductives  appear  in  the  nests  in  late  June  (P. 
rugosus ) or  late  July  ( P . montanus  and  P.  subnitidus ).  The  mating 
flights  are  completed  by  the  first  part  of  September.  The  highest 
sexual  populations  occur  in  mid  August.  Therefore  the  colony 
begins  production  of  reproductives  early  in  the  year  and  allows 
them  to  remain  in  the  nest  for  extensive  periods  of  time,  even 
though  they  are  consuming  food.  This  is  true  to  a lesser  extent  in  P. 
subnitidus,  where  the  reproductives  appear  in  the  nest  in  late  July 
and  most  have  left  the  nest  by  mid  August  (Figure  5). 

There  are  several  interesting  points  in  Figs.  4 & 5.  Although  P. 
rugosus  begins  production  earlier  in  the  year  than  do  the  other  two 
species,  the  populations  of  brood  in  the  nest  reach  peaks  later  in  the 


38 


Psyche 


[Vol.  88 


year.  Pogonomyrmex  rugosus  spreads  reproduction  out  over  the 
year  to  a greater  extent  than  does  P.  montanus.  Pogonomyrmex 
montanus  produces  relatively  more  sexuals  than  does  P.  rugosus  or 
P.  subnitidus  and  in  general  the  production  is  much  higher. 

Mating  flights. 

The  mating  flights  occur  either  in  the  morning  (P.  subnitidus ) or 
the  afternoon  (P.  montanus  and  P.  rugosus ).  Reproductives  of  P. 
montanus  first  appeared  on  the  nest  surface  on  10  August  1978.  The 
reproductives  emerged  from  the  nest  entrance,  scurried  over  the 
mound  for  a few  seconds  and  then  returned  to  the  nest.  They  may 
have  been  evaluating  environmental  conditions  to  determine  when  it 
was  optimal  for  the  mating  flight.  This  behavior  was  found  in  all 
three  species.  A small  flight  occurred  on  29  August  1978  between 
15:30  and  16:20,  a second  larger  flight  occurred  on  9 September 

1978  between  13:20  and  14:10.  The  nests  of  P.  montanus  normally 
have  a single  entrance-exit  hole.  During  the  large  flight  on  9 
September  1978  the  nests  had  2.7  ± 0.3SE  (12)  exit  holes  per  nest 
(range  = 2 to  4).  These  supplemental  exit  holes  allowed  the 
reproductives  to  exit  the  nest  more  rapidly.  I did  not  observe  this 
behavior  in  the  other  two  species.  Reproductives  of  P.  subnitidus 
were  seen  on  the  nest  surface  as  early  as  23  July  1980.  The  flights 
occurred  on  6,  7,  and  8 August  1980  between  8:00  and  9:30.  In  P. 
rugosus , reproductives  first  appeared  on  the  nest  surfaces  on  1 
August  1979.  A large  mating  swarm  was  observed  on  24  October 

1979  between  14:00  and  15:00. 

During  the  time  the  reproductives  left  the  nest,  the  surfaces  of  the 
nests  swarmed  with  workers.  Apparently  most  or  all  of  these 
workers  were  foragers  as  they  were  lighter  in  weight  than  the  other 
ants  in  the  nest  (MacKay,  unpubl.).  The  reproductives  often  had 
considerable  difficulty  becoming  airborne,  especially  the  females, 
which  usually  climbed  up  plant  stems  before  flying. 

Large  mating  swarms  were  observed  in  P.  rugosus  and  were 
similar  to  those  described  by  Holldobler  (1976b).  The  males  waited 
on  the  tops  of  hills  (over  100  m altitude  above  surrounding  terrain) 
for  the  females.  The  males  displayed  considerable  competition  for 
females  as  was  shown  by  Markl  et  al.  (1977).  As  a result  mating  was 
a frenzied  activity  in  which  numerous  males  competed  for  single 
females  by  biting,  pushing,  and  in  general  attempting  to  exclude 


1981] 


Mac  Kay — Nest  Phenologies  of  Pogonomyrmex 


39 


Figure  5.  Seasonal  changes  in  the  populations  of  reproductives  of  three  species  of 
Pogonomyrmex  harvester  ants.  Note  that  the  percentage  scale  for  reproductives  in  P. 
montanus  has  twice  the  range  of  the  scales  for  reproductives  in  the  other  two  species. 


40 


Psyche 


[Vol.  88 


other  males  (See  Figure  2 of  Holldobler  1976b  and  Figure  4 of 
Markl  et  al.  1977).  Prior  to  the  mating  flight,  male  respiratory  rates 
doubled  or  tripled  (MacKay  1981).  The  individuals  with  higher 
activity  levels  may  be  able  to  increase  their  fitness  by  excluding 
other  males  from  a female  or  by  capturing  a female  quickly  and 
moving  into  the  copulatory  position  before  other  males  arrive. 

After  the  female  has  copulated  for  a short  time,  she  bites  the 
gaster  of  the  male  which  is  copulating  with  her.  He  usually 
relinquishes  his  position  to  another  male.  There  is  considerable 
fighting  and  tumbling  so  it  is  difficult  to  determine  the  numbers  of 
times  a female  mates.  Observations  suggest  that  a single  female 
mates  at  least  3 or  4 times.  She  may  have  mated  previously  with  one 
or  more  of  her  brothers  in  the  nest.  I observed  one  mating  within  the 
nest  of  a laboratory  colony  of  P.  montanus.  In  all  three  species,  the 
males  attempt  to  mate  with  their  sisters  during  emergence  from  the 
nest,  although  a complete  copulation  was  never  observed. 

After  several  copulations  the  females  leave  the  mating  swarm 
either  by  flying  or  walking  away.  The  males  no  longer  show  interest 
in  such  females,  as  the  females  apparently  stop  releasing  a phero- 
mone (Holldobler  1976b).  Most  females  then  fly  away  from  the 
area.  A few  remain  and  within  a few  minutes  begin  excavating  nests 
near  the  mating  site.  As  the  density  of  such  nests  is  very  high  (more 
than  4 per  square  meter)  the  success  rate  is  undoubtedly  low. 
Several  times  I saw  females  near  the  mating  area  attempt  to  “steal” 
the  excavation  hole  of  another  female,  but  were  chased  away  by  the 
resident  female.  Such  attempts  are  common  and  are  occasionally 
successful  (Markl  et  al.  1977). 

Seasonal  changes  in  the  positions  of  inhabitants  within  the  nests. 

The  seasonal  movements  in  the  positions  of  the  inhabitants  of  the 
nests  depicted  in  Figures  6,  7 and  8 are  similar  to  those  described  in 
P.  owyheei  (Willard  and  Crowell  1965)  and  P.  occidentalis  (Lavigne 
1969).  The  depths  are  not  comparable  between  the  three  species  as 
the  nests  of  P.  rugosus  are  deeper  than  those  of  P.  subnitidus  which 
are  in  turn  deeper  than  those  of  P.  montanus  (Appendix  1).  In  most 
cases  the  time  axis  is  expressed  in  months  of  the  year  with  the 
exception  of  the  sexuals  in  which  only  four  months  are  shown.  In  all 
cases,  the  proportions  represent  means  of  all  nests  excavated. 

Most  of  the  nest  population  of  P.  montanus,  including  the 


1981]  MacKay—Nest  Phenologies  of  Pogonomyrmex  41 


Figure  6.  Seasonal  movements  of  the  populations  of  the  various  member  groups  in  the  nests  of  P.  man, anus.  The  grid  has  a 
value  of  zero.  The  value  of  the  proportion  of  each  element  in  the  array  is  represented  both  by  the  height  of  the  box  above  the 
grid  and  the  linear  dimensions  of  the  box. 


42 


Psyche 


[Vol.  88 


workers  and  the  nest  queen,  overwinter  near  the  40  cm  level  of  the 
nest  (Figures  6 and  9).  In  the  early  spring  the  soil  temperatures  are 
low  (Figure  2)  and  the  ants  are  very  sluggish.  When  the  snow  begins 
to  melt,  the  lowest  chambers  of  the  nest  fill  with  water.  If  the  ants 
were  at  the  lowest  levels,  they  would  probably  be  killed.  In  April 
and  May  the  P.  montanus  worker  population  begins  to  spread 
throughout  the  nest.  In  June,  July,  and  August,  nearly  80%  of  the 
worker  population  moves  into  the  upper  10  cm  of  the  nest  (Figure 
6).  During  this  time  the  nest  temperatures  are  high  and  much  of  the 
worker  population  is  involved  in  foraging,  brood  care,  and  nest 
construction.  In  September  as  the  soil  temperature  begins  to  cool, 
foraging  decreases  and  the  workers  begin  to  spread  throughout  the 
levels  of  the  nest.  In  December  the  workers  are  again  at  the  40  or  50 
cm  level  of  the  nest.  The  worker  population  in  the  20  and  30  cm 
levels  remains  low  and  relatively  constant  throughout  the  year. 
There  is  apparently  no  temporal  movement  in  the  larvae  or  pupae, 
but  they  are  present  within  the  nest  for  only  part  of  the  year.  In 
general,  they  are  located  at  the  30  or  40  cm  level  where  temperature 
and  humidity  are  relatively  constant  throughout  the  season.  The 
callows  tend  to  occur  in  the  deeper  levels  of  the  nest  together  with 
the  brood.  As  most  of  the  worker  population  is  in  the  upper  levels  of 
the  nest,  the  responsibilities  of  brood  care  are  left  to  the  callows. 

It  is  difficult  to  make  inferences  concerning  the  sexuals  as 
individuals  begin  to  leave  the  nest  in  the  middle  of  August.  Thus, 
what  appears  to  be  a downward  movement  may  simply  be  the  result 
of  the  individuals  in  the  upper  levels  leaving  the  nest.  The  females  do 
tend  to  occur  deeper  in  the  nest  than  do  the  males.  They  may  be  in 
lower  levels  in  the  nest  in  order  to  assist  in  caring  for  the  brood,  as 
has  been  observed  in  the  laboratory.  It  has  been  shown  in  Formica 
polyctena  that  workers  must  learn  brood  care  during  an  early  period 
of  their  lives  or  they  will  never  care  for  brood  (Jaisson  1975).  This 
could  occur  in  Pogonomyrmex  where  the  female  reproductives  may 
“learn”  brood  care  so  they  can  later  rear  their  own  brood. 

The  seasonal  movement  in  P.  subnitidus  nests  is  similar  to  that 
found  in  P.  montanus  nests  (Figure  7).  A high  proportion  of  the 
workers  remains  in  the  upper  30  cm  of  the  nest.  In  October  there  is  a 
dispersion  throughout  the  nest.  By  December,  much  of  the  popula- 
tion is  at  the  120  to  180  cm  level,  with  little  of  the  population  in  the 
lowest  parts  of  the  nest.  The  study  area  receives  less  snow  than  the 


1981]  Mac  Kay — Nest  Phenologies  of  Pogonomyrmex 


43 


Figure  7.  Seasonal  movements  of  the  populations  of  the  various  member  groups  in  the  nests  of  P.  subnitidus. 


44 


Psyche 


[Vol.  88 


area  containing  P.  montanus,  but  the  lower  levels  of  the  nest  may 
also  become  flooded  when  the  snow  melts.  Many  of  the  larvae  and 
pupae  are  found  in  the  upper  levels  of  the  nest,  but  there  is 
apparently  a downward  movement  of  the  brood  and  callows  in 
October  and  November.  By  December  there  is  no  brood  in  the  nest. 
Most  of  the  reproductives  are  found  in  the  upper  30  cm  of  the  nest 
(Fig.  7). 

The  seasonal  movements  in  P.  rugosus  nests  are  similar  to  the 
other  two  species  (Figure  8).  Most  of  the  worker  population  is  in  the 
upper  levels  of  the  nest  throughout  the  spring  and  summer.  In 
September  and  October  until  December,  the  ants  become  distri- 
buted throughout  the  nest.  The  larvae  are  dispersed  throughout  the 
nest  during  most  of  the  year,  but  appear  to  be  moved  into  the  deeper 
levels  of  the  nest  at  the  beginning  of  the  winter.  The  pupae  are 
located  in  the  upper  levels  of  the  nest  but  also  appear  to  be  moved 
into  the  deeper  regions  of  the  nest  in  the  fall.  The  callows  also 
demonstrate  a movement  into  the  deeper  nest  levels  in  the  fall. 
Again,  it  is  difficult  to  make  inferences  concerning  the  sexuals  as 
they  are  in  the  nest  for  a short  period  of  time,  but  both  sexes  appear 
to  be  in  the  upper  levels. 

In  the  winter  the  ants  seem  to  be  dispersed  throughout  the  nest 
and  do  not  avoid  the  lowest  levels  of  the  nest.  There  is  no  winter 
snow  at  Riverside  and  the  temperatures  are  higher  than  those  in  the 
mountains  (Figure  1),  therefore  the  ants  remain  somewhat  active 
throughout  the  year. 

The  seasonal  patterns  of  distribution  within  the  nests  are  similar 
in  all  three  species.  The  reproductives  (when  present)  and  workers 
are  most  abundant  in  the  upper  levels  of  the  nest,  except  in  the 
winter.  The  brood  are  in  the  deeper  levels  where  the  microclimate 
undergoes  little  change.  The  callows  are  in  the  lower  levels  of  the 
nests  in  all  three  species  and  apparently  care  for  the  brood.  This  is 
common  in  ants  in  general  (Wilson  1971)  and  in  P.  badius  (Gentry 
1974).  No  callows  were  ever  seen  foraging.  They  do  not  quickly 
darken  on  exposure  to  sunlight. 

It  is  commonly  stated  that  ants  keep  the  larvae  and  pupae 
separate  within  the  nest  to  take  advantage  of  the  optimal  conditions 
for  the  development  of  each  (Wheeler  1910;  Protomastro  1973).  In 
Pogonomyrmex,  at  least  P.  marcusi  is  reported  to  practice  such 
behavior  (Marcus  and  Marcus  1951).  I have  no  evidence  that  the 


200 
250  / 


1981] 


MacKay — Nest  Phenologies  of  Pogonomyrmex 


45 


Figure  8.  Seasonal  movements  of  the  populations  of  the  various  member  groups  in  the  nests  of  P.  rugosus 


46 


Psyche 


[Vol.  88 


Table  1.  Three-way  analysis  of  variance  comparisons  of  the  positions  of  larvae  and 
pupae  in  26  nests  of  P.  montanus  collected  in  1978  and  1979,  3 nests  of  P.  subnitidus 
collected  in  1979,  and  9 nests  of  P.  rugosus  collected  in  1979.  (As  the  data  were 
expressed  as  percentages  of  the  total  nest  population,  they  were  subjected  to  an  arcsin 
transformation  before  analysis.) 


Source 

df 

MS 

F 

P.  montanus 

Different  nests 

25 

0.005 

0.556  ns 

Positions  of  larvae  and  pupae 

1 

0.007 

0.778  ns 

Levels  in  nests 

7 

0.315 

35.000*** 

Nests  X brood 

25 

0.004 

0.444  ns 

Nests  X levels 

175 

0.110 

12.222*** 

Brood  X levels 

7 

0.028 

3.111** 

error 

174 

0.009 

P.  subnitidus 

Different  nests 

2 

0.000 

0.000  ns 

Positions  of  larvae  and  pupae 

1 

0.000 

0.000  ns 

Levels  in  nests 

22 

0.017 

5.667*** 

Nests  X brood 

2 

0.000 

0.000  ns 

Nests  X levels 

44 

0.017 

5.667*** 

Brood  X levels 

22 

0.003 

1.000  ns 

error 

43 

0.003 

P.  rugosus 

Different  nests 

8 

0.000 

0.000  ns 

Positions  of  larvae  and  pupae 

1 

0.000 

0.000  ns 

Levels  in  nests 

39 

0.019 

9.500*** 

Nests  X brood 

8 

0.000 

0.000  ns 

Nests  X levels 

312 

0.010 

5.000*** 

Brood  X levels 

39 

0.003 

1.500  ns 

error 

311 

0.002 

larvae  and  pupae  are  placed  in  separate  levels  of  the  nests  in  any  of 
the  three  species  (Table  1).  There  is  a significant  difference  between 
the  levels  of  the  nests,  which  is  evident  in  Figures  6,  7,  and  8.  The 
brood  tend  to  be  in  the  lower  levels  of  the  nest.  Although  it  is 
commonly  assumed  there  is  segregation  of  the  larvae  and  pupae, 
statistical  analysis  has  not  been  performed  in  the  past  to  support  the 
assumption. 

In  one  instance,  a P.  montanus  nest  placed  a large  number  of 
brood  on  the  soil  surface  near  the  nest  entrance  after  a late-summer 


1981]  Mac  Kay — Nest  Phenologies  of  Pogonomyrmex 


47 


Table  2.  Analysis  of  variance  comparisons  of  the  positions  of  males  and  females  in 
17  nests  of  P.  montanus  collected  in  1978  and  1979,  4 nests  of  P.  subnitidus  collected 
in  1980,  and  one  P.  rugosus  nest  collected  in  1979.  (The  data  were  subjected  to  an 
arcsin  transformation  before  analysis.) 


Source 

df 

MS 

F 

P.  montanus 

Different  nests 

16 

0.007 

0.538ns 

Males  and  females 

1 

0.000 

0.000ns 

Levels  in  nests 

7 

0.232 

17.846*** 

Nests  X sexuals 

16 

0.005 

0.385ns 

Nests  X levels 

112 

0.072 

5.538*** 

Sexuals  X levels 

7 

0.062 

4.769*** 

error 

111 

0.013 

P.  subnitidus 

Different  nests 

3 

0.000 

0.044ns 

Males  and  females 

1 

0.002 

0.231ns 

Levels  in  nests 

7 

0.455 

45.083*** 

Nests  X sexuals 

3 

0.001 

0.065ns 

Nests  X levels 

21 

0.006 

0.630ns 

Sexuals  X levels 

7 

0.040 

3.924** 

error 

20 

0.010 

P.  rugosus 

Males  and  females 

1 

0.002 

1.000ns 

Levels  in  nest 

17 

0.006 

3.000* 

error 

16 

0.002 

rain,  possibly  because  the  upper  levels  of  the  nest  had  become 
waterlogged.  A considerable  number  of  workers  guarded  the  brood 
during  this  time  and  when  disturbed,  the  workers  immediately 
moved  the  brood  back  into  the  nest.  This  behavior  has  not  been 
observed  in  the  other  two  species. 

The  positions  of  the  males  and  females  were  compared  with  an 
analysis  of  variance  (Table  2).  Although  it  appears  from  Figures  6, 
7,  and  8 and  our  impressions  in  the  field,  that  females  are  in  deeper 
levels  of  the  nest  than  the  males,  there  is  no  statistical  support 
(Table  2).  There  were  significant  differences  between  the  levels. 
Figures  6,  7,  and  8 illustrate  that  the  reproductives  tend  to  be  in  the 
upper  levels  of  the  nests. 

In  Pogonomyrmex  spp.  there  is  evidence  that  little  mixing  of 
adult  workers  occurs  within  the  nests  (Chew  1960;  Golley  and 


48 


Psyche 


[Vol.  88 


Gentry  1964;  Gentry  1974).  MacKay  (1981)  presents  data  on  the 
respiratory  rates  and  fat  contents  of  workers  taken  from  the 
different  levels  of  the  nests  of  the  three  species.  In  winter,  spring, 
and  fall,  there  are  significant  differences  between  the  levels  with 
regard  to  both  of  these  parameters.  If  mixing  of  the  workers  did 
occur  between  the  different  levels  of  the  nest,  we  would  not  have 
found  these  consistent  differences  between  workers  taken  from 
different  levels. 

There  is  little  evidence  of  seasonal  movements  of  the  nest  queens 
(Figure  9).  In  the  spring  P.  occidental is  queens  ascend  into  the 
upper  levels  from  the  lower  levels  (Lavigne  1969).  The  queens  may 
be  moved  into  the  deeper  regions  during  the  winter  for  greater 
protection.  In  the  spring,  the  soil  begins  to  warm  sooner  in  the 
superficial  levels.  The  queen  may  be  moved  to  the  higher  warmer 
levels  in  order  to  increase  her  metabolism  for  initiation  of  egg 
production. 

Guests. 

Many  species  of  insects  and  spiders  were  collected  within  the  ant 
nests.  The  occurrence  of  most  of  these  species  is  probably  accidental 
and  individuals  of  most  species  were  found  only  in  small  numbers 
(one  or  two  individuals  per  nest).  Those  species  most  commonly 
found  include:  Orthoptera — Myrmecophila  manni  Schimmer,  in  the 
nests  of  all  three  species;  Coleoptera — Echinocoleus  setiger  Horn,  in 
P.  montanus  and  P.  subnitidus  nests,  Hetarius  hirsutus  Martin  and 
H.  sp.#l  with  P.  montanus,  H.  morsus  Leconte  and  H.  sp.#2  with  P. 
subnitidus,  Cremastocheilus  westwoodi  Horn  in  the  nests  of  P. 
subnitidus.  There  are  at  least  two  species  of  unidentified  staphylin- 
ids  that  are  common  in  P.  subnitidus  nests  (more  than  10  per  nest). 
Hymenoptera — Solenopsis  molesta  (Say)  is  common  in  P.  mon- 
tanus and  P.  subnitidus  nests,  Pheidole  spp.  in  P.  rugosus  nests.  Of 
the  three  harvester  ant  species,  P.  subnitidus  has  the  greatest 
number  of  guests  and  diversity  of  species. 

Food  input  into  nests. 

All  three  species  demonstrate  similar  seasonal  changes  in  their 
foraging  patterns,  with  much  activity  in  mid-summer  and  no  activity 
in  the  winter  and  early  spring  (Figures  10  and  11).  There  are 
important  differences  between  the  three  species.  Foraging  in  P. 
rugosus  begins  earlier  in  the  spring  and  extends  later  into  the  fall 
than  in  the  other  two  species.  Pogonomyrmex  subnitidus  has  an 


1981]  Mac  Kay — Nest  Phenologies  of  Pogonomyrmex 


49 


MONTHS 


Figure  9.  Seasonal  changes  in  the  positions  of  the  nest  queen  in  three  species  of 
Pogonomyrmex  harvester  ants. 


especially  short  foraging  period.  Pogonomyrmex  montanus  begins 
the  spring  with  an  abrupt  increase  in  foraging  (Figure  10).  The  lower 
altitude  species,  P.  rugosus,  is  exposed  to  many  sunny  days  during 
the  winter.  During  most  of  this  time  the  nests  of  the  high  altitude 
species,  P.  montanus,  are  covered  with  snow.  The  nests  of  the  mid 
altitude  species,  P.  subnitidus,  are  covered  by  snow  part  of  the  time. 
In  May  or  June  foraging  begins,  increases  throughout  the  summer 
and  decreases  again  in  the  fall.  This  foraging  pattern  corresponds 
well  with  the  production  of  workers  and  reproductives  within  the 
nest. 

Only  a small  portion  of  the  population  is  involved  in  foraging. 
The  mean  number  of  foragers  per  day  (recorded  during  July  and 
August,  the  months  of  peak  foraging)  were  378  ± 73.2  (6)  for  P. 
montanus,  648  + 177.3  (4)  forP.  subnitidus,  and  1427  ± 187.3  (5)  for 
P.  rugosus.  Later  excavation  of  the  nests  indicated  that  the 
population  of  foragers  comprised  22.9%,  19.4%,  and  18.4%  of  the 
total  nest  populations  of  P.  montanus,  P.  subnitidus,  and  P. 
rugosus,  respectively.  Others  have  estimated  that  10%  of  the 
population  is  involved  in  foraging  in  such  species  as  P.  badius 
(Golley  and  Gentry  1964),  P.  calif ornicus  (Erickson  1972)  and  P. 


50 


Psyche 


[Vol.  88 


MONTHS 

Figure  10.  A comparison  of  the  number  of  daily  foraging  trips  in  three  species  of 
Pogonomyrmex  harvester  ants.  The  horizontal  lines  indicate  the  means,  the  black 
rectangles  the  standard  errors  on  each  side  of  the  mean,  and  the  vertical  lines  indicate 
the  ranges. 


Table  3.  Nest  densities,  populations  and  biomasses  of  several  ant  species  of  the  genus  Pogonomyrmex.  The  values  are 
±1  standard  error,  n is  presented  in  parenthesis. 


1981]  Mac  Kay — Nest  Phenologies  of  Pogonomyrmex  51 


\o  ~o 

On  C 

— cd 

s- 

C C 

<u  aj 

a o 


c/3 


gf 

C cd 

o Q 

T3 

X "O 

£ g 


--  (N 
X r- 

£ ^ 


1 * 


O' 


O *-!  O 


£■1 
a>  u. 
X ,0 

3 X 
O 3 
C/3  U 


-3  <u 

1 ^ 

cd  0 

£ 


60 

C On 

2 £ 


C4  3 
O a -3 

Z.  Cl  "3 
^ S'  -Cl 


desertorum  115  400-600  New  Mexico  Whitford  and 

Bryant  1979 


Table  3 continued 


52 


Psyche 


[Vol.  88 


6 o 
£ 


■g  m ^ 

3 t,  c 

_ ° c 
£ 


3 3 

T3  *n 

3 o S 

tt  £ e 


o 

<u 

Oh 

C/3 


>>  .2 

5 | 

1 " 

2 3 


3 33 
O 3 
C/3  U 


ON 
NO 
O ON 
NO  — 
ON 

<U 

c 

£ 60 
Si  ’> 

33  efl 

U J 


60 

3 .s 

§ I 

■si 

< £ 


OO  II  <N 

“ y>3  £5 

m 60  I O ’'3’ 
NO  o c ON  O (N 

vo  fC  sj  vo  r~  o 
— ' w u-  m oo  cn 


.03  .«n 

2 2 

s*  c 

8 8 

o o 


NT) 

r- 

■o  2 
c ~ 


OJ 

rsi 

03  1-^ 

« ON 
60  — 
O 

04  3 


do  o 
o o 
o o 
— m 


o 

o 

m 

+1 

m 

ON  _ 
60 


O ” — 


^ W ^ u 


2 D 


o *3 
C/3  U 


II  CN 

S 15 

^ © oo 

^ M ^ 

r»  w <n 


>3  -2 

5 l 

I & 

S 3 


O as 

£ U 


•5  <2 

3 33 
O 3 
C/3  U 


- ^ 

£ £ 

> § 
<D  2 


II  NO 

g 

m 3 22 

+l  2 7 

^ o 
^ m 
ON  fO  oo 

mo  — 


1981]  Mac  Kay — Nest  Phenologies  of  Pogonomyrmex  53 


in  in 
Os  Os 


> > 

o o 

N N 

CL>  <D 

3 C 

C/3  C/3 

3 3 


3 

C 


C 

!U 

oo 


< 


< < 


3 

.3 

3 

<u 

00 

< 


3 

3 


< CQ 


3 

'JS 

B 

_o 

o 

U 


3 

C 

3 

D 

00 

< 


* Estimations  based  on  data  from  literature. 
**  From  Brian  (1965) 


Specie 


(II) 


0.8-8.3* 


1. 2-3.9*  8.6-27.9* 


I 


i 

SS 


1 


I 


b3 


54 


Psyche 


[Vol.  88 


occidentalis  (Rogers  et  al.  1972).  Chew  (1960)  estimated  that  no 
more  than  !/2  of  P.  occidentalis  workers  were  out  of  the  nest  at  any 
one  time.  In  a mark  recapture  analysis,  Whitford  et  al.  (1976) 
estimated  the  forager  population  at  2786  in  P.  rugosus.  This 
estimate  is  higher  than  the  one  I determined  which  may  indicate  that 
the  nest  populations  of  P.  rugosus  in  New  Mexico  are  larger  than 
those  in  southern  California.  My  estimates  are  minimal:  there  may 
have  been  foragers  which  remained  within  the  nest.  Also  the 
experimental  channeling  of  the  forager  population  may  have 
affected  the  natural  foraging  activity.  The  whole  work  force  may  not 
have  been  activated  because  of  a reduction  of  recruitment  (Hbll-' 
dobler,  Pers.  Comm.). 

A comparison  of  the  number  of  foragers  given  above  and  the 
number  of  foraging  trips  per  day  (Figure  10)  indicates  that 
individual  P.  montanus  foragers  make  two  or  three  trips  per  day,  P. 
subnitidus  foragers  about  nine,  and  P.  rugosus  foragers  make  more 
than  ten  trips  per  day.  There  are  considerable  differences  between 
the  three  species  in  the  numbers  of  foraging  trips  made  (Figure  10), 
which  compares  with  the  differences  in  the  sizes  of  the  nest 
populations  (Table  3). 

The  seasonal  changes  in  the  daily  amount  of  food  brought  to  the 
nest  are  similar  to  those  found  in  the  numbers  of  foraging  trips 
(Figure  11).  As  with  the  forager  number,  P.  rugosus  brings  in  food 
earlier  in  the  spring  and  extends  foraging  later  into  the  fall, 
compared  to  the  other  two  species.  Pogonomyrmex  montanus 
abruptly  increases  the  food  input  once  foraging  begins  and  de- 
creases it  slowly  until  fall.  Pogonomyrmex  montanus  is  the  only 
species  of  the  three  which  does  not  store  seeds  in  the  nests.  It  may 
have  to  bring  in  large  amounts  of  food  once  the  larvae  begin  to 
appear  in  the  nest.  The  other  two  species  have  seed  reserves  and  may 
thus  avoid  such  an  abrupt  increase  in  foraging  in  the  spring. 

Comparisons  of  the  food  sources  of  the  three  species  (Figure  12) 
indicate  that  the  harvester  ants  utilize  a wide  variety  of  food  items, 
although  most  materials  are  either  seeds  or  plant  parts.  Pogono- 
myrmex rugosus  relies  almost  exclusively  on  seeds.  Pogonomyrmex 
subnitidus  and  especially  P.  montanus  bring  a much  greater 
diversity  of  food  items  to  the  nest.  Pogonomyrmex  montanus  relies 
more  heavily  on  plant  parts  and  insects  than  does  P.  subnitidus. 
Pogonomyrmex  subnitidus  brings  in  a greater  proportion  of  feces 
than  does  P.  montanus,  although  the  ratio  of  bird  to  mammal  feces 


DAILY  FOOD  INPUT  (g) 


1981] 


Mac  Kay — Nest  Phenologies  of  Pogonomyrmex 


55 


1.4 


0.6  - 


0.4  - 


0.2  - 


5 

4 

3 

2 

I 

40 

30 

20 

101- 


P montonus 


I 


P.  subnitidus 


if 


^ .. 


P.  rugosus 


I 


t- 


JFMAMJ  JASOND 
MONTHS 


Figure  11.  A comparison  of  the  daily  food  input  (grams)  in  the  three  species  of 
Pogonomyrmex  harvester  ants. 


56 


Psyche 


[Vol.  88 


SOURCES  OF  FOOD 


Plant  parts 
Seeds 
Soft  insects 
Hard  insects 
Bird  feces 
Mammal  feces 


Figure  12.  A comparison  of  the  food  sources  in  the  three  species  of  Pogono- 
myrmex  harvester  ants. 


1981]  MacKay — Nest  Phenologies  of  Pogonomyrmex 


57 


is  similar  in  both  species.  Pogonomyrmex  rugosus  brings  in  more 
bird  feces  than  mammal  feces,  P.  montanus  and  P.  subnitidus  bring 
in  more  mammal  feces  than  bird  feces.  A distinction  was  made 
between  “hard”  insects  and  “soft”  insects.  Hard  insects  included 
those  heavily  chitinized  forms,  especially  the  Coleoptera  and  certain 
Formicidae.  Soft  insects  included  Homoptera,  most  Hemiptera, 
most  Diptera,  larvae  and  pupae  of  most  orders  and  a few  non- 
insects such  as  spiders.  It  appears  that  the  degree  of  chitinization 
may  not  be  important  as  the  proportions  of  hard  and  soft  insects 
were  similar.  All  three  species  have  chitinase  activity  in  their  gasters 
(MacKay,  unpub.  data). 

Plant  parts  consist  of  pieces  of  leaves  and  flowers  and  in  the  case 
of  P.  montanus,  pine  resin.  Flowers  of  Penstemon  spp.  and 
Arctostaphylos  spp.  are  transported  to  the  nest  and  placed  around 
the  brood,  possibly  to  increase  the  humidity.  Later  the  intact  flowers 
are  discarded  at  the  nest  surface.  This  indicates  the  flowers  are  not 
placed  around  the  brood  to  protect  them  from  predators.  In  the  case 
of  pieces  of  leaves,  apparently  they  are  eaten  by  the  ants  as  they  do 
not  later  appear  on  the  nest  surface.  There  is  considerable  seasonal 
change  in  the  food  composition  of  P.  montanus  and  P.  subnitidus 
(Figure  13).  The  percentages  of  insects  brought  into  P.  montanus 
nests  changes  little  seasonally.  There  is  a seasonal  reduction  in  the 
percentage  of  utilization  of  insects  in  P.  subnitidus.  There  is  little 
seasonal  change  in  the  proportion  of  the  food  sources  composed  of 
feces  in  the  two  species,  although  a slight  reduction  may  occur.  In 
both  species,  especially  P.  montanus,  there  is  a seasonal  decrease  in 
the  proportion  of  plant  parts  brought  to  the  nest.  In  both  species, 
there  is  a dramatic  increase  in  the  utilization  of  seeds  after  July.  This 
increase  is  probably  related  to  a greater  availability  of  seeds  after  the 
flowering  period  of  annual  plants.  A similar  comparison  was  not 
made  in  the  case  of  P.  rugosus  as  non-seed  materials  are  a very  small 
portion  of  their  diet  (Fig.  12).  In  P.  rugosus,  there  was  a seasonal 
drop  in  the  proportion  of  the  diet  composed  of  Erodium  eicutarium 
(L.)  L’Her.  seeds  (May  90.3%,  June  91.0%,  July  88.9%,  August 
89.7%,  September  84.1%,  and  October  80.9%).  Other  seeds,  espe- 
cially those  of  Pectocarya  linearis  DC  and  Festuca  octoflora  Walt., 
made  up  most  of  the  difference. 

Caloric  analysis  of  the  food  entering  the  nests  of  the  three  species 
indicates  that  a P.  montanus  colony  receives  an  average  of  166.6 


PERCENT  OF  TOTAL  FOOD  INPUT 


58 


Psyche 


[Vol.  88 


Figure  13.  The  seasonal  changes  in  the  food  sources  of  P.  montanus  and  P. 
subnitidus. 


1981]  Mac  Kay — Nest  Phenologies  of  Pogonomyrmex 


59 


kcals,  a P.  subnitidus  nest  1267.0  kcals,  and  a P.  rugosus  nest  7613.6 
kcals  of  food  during  a year  (MacKay  1981).  Of  these  amounts,  a P. 
rugosus  colony  discards  seed  husks  and  other  such  materials,  a 
quantity  consisting  of  5004.5  kcals  or  65.7%  of  the  intake.  This  is 
indicated  in  the  field  by  large  discard  piles  of  seed  husks  being 
deposited  around  the  nests.  A few  seeds  are  discarded  and  germinate 
from  the  piles  in  the  spring.  Another  harvester  ant,  Veromessor 
pergandei  (Mayr)  forages  in  the  piles  and  removes  many  of  the 
discarded  seeds.  Pogonomyrmex  montanus  and  P.  subnitidus 
discard  few  materials,  the  amounts  are  too  small  to  be  estimated. 

Seed  storage. 

Seasonal  changes  in  seed  storage  in  P.  subnitidus  and  P.  rugosus 
are  shown  on  Figure  14.  Pogonomyrmex  rugosus  began  both  1979 
and  1980  (data  for  January)  with  0.04^0.06  grams  of  seed  storage 
per  ant.  The  correlation  of  ant  number  vs.  seed  weight  was  very  high 
(r  = 0.997,  p < 0.01).  This  amount  dropped  until  May,  possibly  the 


Figure  14.  A comparison  of  the  seasonal  changes  of  seed  storage  in  P.  rugosus 
and  P.  subnitidus. 


60 


Psyche 


[Vol.  88 


result  of  seed  consumption  by  the  developing  larvae.  I have  no 
explanation  for  the  other  two  peaks  which  appear.  There  is  some 
evidence  of  a drop  in  seed  storage  in  the  spring  in  P.  subnitidus,  but 
it  is  not  as  great  as  that  found  in  P.  rugosus.  Pogonomyrmex 
subnitidus  also  appears  to  begin  the  season  with  a constant  amount 
of  seeds,  about  0.002-0.004  g/ant,  much  smaller  quantities  than  P. 
rugosus.  There  are  also  many  unexplained  peaks  in  P.  subnitidus 
seed  storage,  especially  the  high  peak  in  September.  Pogono- 
myrmex montanus  does  not  store  seeds  in  the  nest.  In  the 
population  at  Big  Pine  Flats  in  the  San  Bernardino  Mountains,  we 
occasionally  encountered  very  small  caches  of  seeds  (less  than 
0.0001  g/ant)  which  were  apparently  only  small  daily  accumulations 
of  seeds  that  had  not  been  eaten  at  that  time. 

Production. 

Production  in  the  three  species  is  summarized  in  Table  4.  The 
proportion  of  energy  invested  in  production  varies  considerably 
between  the  three  species,  but  in  all  cases  it  is  relatively  low.  Total 
production  constitutes  12.2,  8.3,  and  7.9  per  cent  of  the  total  energy 
flow  in  P.  montanus,  P.  subnitidus,  and  P.  rugosus  respectively 
(MacKay  1981).  In  all  three  species,  a higher  percentage  of  the  total 
production  is  invested  in  workers  than  reproductives  (Table  4). 
Pogonomyrmex  subnitidus  and  P.  rugosus  both  invest  heavily  in 
workers,  P.  montanus  invests  heavily  in  reproductives.  The  data  on 
Table  4 suggest  that  the  three  species  invest  more  in  the  production 
of  females  than  in  males.  The  costs  of  respiration  of  males  are  higher 
than  of  females  (MacKay  1981).  When  respiration  costs  are  taken 
into  account,  the  colonies  of  each  species  invest  about  equally  in  the 
production  of  males  and  females  (MacKay  1981).  More  numbers  of 
males  than  females  are  produced  in  all  three  species  (Table  4). 
Individual  females  are  more  expensive  to  produce  than  are  indi- 
vidual males  (MacKay  1981). 

Most  of  the  workers  are  replaced  each  year.  Pogonomyrmex 
montanus  colonies  produce  1516  workers  per  year  (Table  4),  which 
is  similar  to  the  mean  worker  population  of  1665  (Table  3). 
Pogonomyrmex  subnitidus  colonies  produce  3988  workers  as 
compared  to  a worker  nest  population  of  5934;  P.  rugosus  colonies 
produce  5298  workers  per  year  compared  to  a worker  nest 
population  of  7740. 


1981] 


Mac  Kay — Nest  Phenologies  of  Pogonomyrmex 


61 


Table  4.  A comparison  of  the  investments  in  production  in  three  species  of 
Pogonomyrmex  harvester  ants. 


Species 

Group 

Number  of 
Individuals 

Dry  wt 
(g) 

kcals 

Percent  Total 
Production 

montanus 

Workers 

1516+  95 

2.4 

12.7 

51.8 

Females 

187+  30 

1.2 

7.8 

31.8 

Males 

239+  41 

0.8 

4.0 

16.3 

24.5 

subnitidus 

Workers 

3988+438 

11.6 

87.4 

91.5 

Females 

111+  65 

0.9 

5.4 

5.7 

Males 

251+  87 

0.6 

2.7 

2.8 

95.5 

rugosus 

Workers 

5298+763 

30.3 

208.2 

86.6 

Females 

118+100 

2.7 

18.9 

7.9 

Males 

312+  73 

2.5 

13.4 

5.6 

240.5 

Discussion 

Comparison  with  other  species  in  the  genus  Pogonomyrmex. 

The  genus  Pogonomyrmex  belongs  to  the  tribe  Myrmicini,  one  of 
the  most  primitive  tribes  in  the  subfamily  Myrmicinae.  The  genus 
has  existed  at  least  since  the  Oligocene  (Burnham  1978),  and  is 
distributed  throughout  North  and  South  America  from  Canada  to 
Patagonia,  from  sea  level  to  at  least  4500  meters  in  altitude.  At  the 
present  time  there  are  24  valid  species  in  North  America  and  about 
33  in  Central  and  South  America.  The  genus  may  have  originated  in 
South  America  and  migrated  northward  (Kusnezov  1951)  or 
originated  in  North  America  and  migrated  southward  (Wheeler 
1914;  Creighton  1952). 

Considerable  work  has  been  done  on  nest  densities,  populations, 
and  biomasses  of  ants  of  various  species  of  the  genus  Pogono- 
myrmex (Table  3).  Examples  of  biomasses  from  other  genera  would 
include  the  following  (expressed  as  mg  dry  weight/  m2),  Tetramori- 
um  caespitum  at  200  (Brian  et  al.  1967)  and  1480  (Nielsen  1974), 
Lasius  niger  at  60  (Odum  and  Pontin  1961)  and  1060  (Nielsen  1974), 
L.  alienus  at  2090  (Nielsen  1974),  L.  flavus  at  1400  (Odum  and 
Pontin  1961)  and  15,000  (Waloff  and  Blackith  1962),  Leptothorax 
acervorum  at  3000  (Brian  1956),  and  Formica  rufa  at  12,000 


62 


Psyche 


[Vol.  88 


(Marikovsky  1962).  In  general,  the  biomasses  of  Pogonomyrmex 
are  much  lower  than  those  found  in  other  genera. 

The  species  investigated,  especially  P.  subnitidus  and  P.  rugosus, 
are  comparable  to  most  of  the  North  American  representatives  of 
the  genus  (Table  3).  The  South  American  species  apparently  have 
much  smaller  populations,  but  few  nests  have  been  excavated  and 
most  were  partial  excavations  in  which  the  queen  was  not  found  or 
after  the  excavation  was  finished,  additional  ants  were  found  later. 
Species  from  arid  regions  tend  to  have  larger  colonies  than  those 
from  mesic  environments,  with  the  exception  of  P.  laticeps.  The 
colonies  of  North  American  species  live  longer  than  South  Ameri-* 
can  species  (Kusnezov  1951).  Pogonomyrmex  montanus  is  some- 
what atypical  for  the  genus  in  occurring  at  higher  altitudes,  but  is 
similar  to  other  species  in  several  aspects.  The  number  of  nests  per 
hectare  is  comparable  to  several  other  species  including  P.  badius, 
P.  barbatus,  P.  occidentals,  P.  owyheei,  P.  rugosus,  and  P. 
subnitidus.  The  nest  populations  of  P.  montanus  are  smaller  than 
those  of  most  of  the  other  species,  but  the  number  of  workers/ m.sq. 
and/or  the  dry  wt/m.sq.  are  comparable  to  P.  badius,  P.  calif orni- 
cus,  P.  occidentals,  P.  owyheei,  P.  rugosus,  and  P.  subnitidus. 

With  regards  to  the  populations,  the  three  species  investigated 
appear  to  be  “typical”  North  American  Pogonomyrmex  harvester 
ants.  It  would  be  very  interesting  to  do  a comparable  study  of 
“typical”  South  American  Pogonomyrmex  harvester  ants. 

Effect  of  altitude. 

It  was  anticipated  that  altitude  would  have  three  primary 
effects:  1)  The  higher  altitude  species,  P.  montanus,  would  be 

subjected  to  lower  average  temperatures.  2)  The  higher  altitude 
species  would  be  subjected  to  shorter  foraging  seasons,  thus 
reducing  the  yearly  food  input  into  the  nest,  resulting  in  lower 
production.  3.  The  higher  altitude’s  shorter  growing  season  would 
result  in  fewer  available  seeds  from  annual  plants. 

Although  P.  montanus  is  subjected  to  the  lowest  seasonal 
temperatures  of  the  specific  populations  of  the  three  species 
investigated  (Figure  2),  it  metabolically  compensates  for  this  by 
having  higher  respiratory  rates  than  the  other  species  (MacKay 
1981).  Apparently  altitude  has  an  effect  on  foraging,  although  it  was 
not  as  large  as  expected.  The  foraging  season  was  somewhat 
reduced  in  P.  montanus  and  P.  subnitidus,  when  they  are  compared 


1981] 


Mac  Kay — Nest  Phenologies  of  Pogonomyrmex 


63 


with  P.  rugosus  (Figure  10).  Pogonomyrmex  montanus,  and  to 
some  extent  P.  subnitidus,  are  in  habitats  with  winter  snow  cover.  In 
such  habitats  foraging  during  the  winter  is  not  possible.  Pogono- 
myrmex rugosus  occupies  a low  altitude  habitat  where  there  are 
many  warm  sunny  days  during  the  winter.  During  these  days,  it  does 
not  forage,  although  a few  workers  are  on  the  nest  surface  either 
sunning  themselves  or  working  on  nest  reconstruction. 

The  higher  altitudes  had  shorter  growing  seasons,  resulting  in 
fewer  annual  seed  producing  plants.  As  a result  P.  montanus  and  P. 
subnitidus  foraged  on  various  materials  but  began  to  rely  heavily  on 
seeds  later  in  the  year  (Figure  13).  This  was  especially  the  case  in  P. 
montanus,  which  relied  heavily  on  plant  parts  early  in  the  year. 
Later  when  seeds  became  more  available,  they  almost  completely 
replaced  plant  parts  in  the  diet  (Figure  13). 

Allocation  of  resources  between  worker  and  reproductive  produc- 
tion. 

As  was  expected,  the  highest  altitude  species  was  exposed  to  a 
shorter  foraging  season,  but  this  did  not  result  in  lower  production. 
The  highest  altitude  species,  P.  montanus,  invests  a larger  propor- 
tion of  energy  into  production  than  do  the  other  two  species.  The 
amount  invested  in  reproductives  is  especially  high  (Table  4). 
Pogonomyrmex  subnitidus  and  P.  rugosus  invested  about  equally  in 
production,  with  investment  in  reproductives  very  low  compared  to 
P.  montanus  (Table  4). 

Most  Pogonomyrmex  spp.  are  low  altitude  desert  species  (Cole 
1968).  Pogonomyrmex  montanus  appears  to  be  in  a marginal 
habitat  for  Pogonomyrmex  spp.  in  that  it  occurs  in  a high  altitude 
pine  forest.  The  nest  populations  are  among  the  smallest  for  the 
genus  (Table  3)  and  the  nests  are  also  very  shallow  (Appendix  1). 
Both  P.  montanus  and  P.  subnitidus  have  shorter  foraging  seasons 
and  apparently  are  not  able  to  exploit  their  optimal  food  source 
(seeds)  until  late  in  the  season  (Figure  13).  Simulations  of  the  effects 
of  bad  years  on  the  nests  indicate  that  P.  rugosus  and  P.  subnitidus 
are  able  to  withstand  moderately  large  reductions  in  food  input 
whereas  P.  montanus  is  not  (MacKay  in  prep.).  As  a result,  nests 
may  be  short-lived  as  compared  to  the  other  two  species  and  nest- 
extinction  may  be  a common  phenomenon.  Apparently,  as  a 
response  to  such  conditions,  P.  montanus  invests  a larger  propor- 
tion of  energy  in  the  production  of  reproductives  than  do  the  other 


64 


Psyche 


[Vol.  88 


two  species.  It  might  be  expected  that  the  South  American  species 
would  be  ecologically  similar  to  P.  montanus  as  they  share  many 
characteristics  (Table  3). 

Production  as  well  as  foraging  and  food  input  were  spread  over 
more  of  the  season  in  P.  rugosus  than  in  the  other  two  species  (Figs. 
4,  5,  10  & 1 1).  This  is  easily  explained  as  P.  rugosus  lives  in  a more 
moderate  climate  than  the  other  two  species.  Actually  it  was 
expected  that  these  processes  would  occur  over  the  entire  year  as 
there  are  many  warm  sunny  days  at  lower  elevations  during  the 
winter.  Yet,  activities  almost  stop.  Perhaps  these  processes  do  not 
continue  as  the  nest  temperatures  are  lower  during  the  winter  than 
they  are  in  the  summer  (Figure  2). 

The  sex  ratio  was  not  constant  between  years  (see  data  in 
Appendix  1).  In  P.  montanus  the  female:male  ratio  was  0.88:1  in 
1978,  1.41:1  in  1979,  and  0.42:1  in  1980.  In  1980  the  number  of 
males  produced  was  three  times  those  of  the  other  years.  An  excess 
of  females  in  1 979  was  not  found  in  P.  rugosus  (0.38: 1 ) as  was  found 
in  P.  montanus.  An  excess  of  males  was  found  in  P.  subnitidus 
(0.42:1)  in  1980  as  was  found  in  P.  montanus. 

Nests  are  extremely  heterogeneous  in  regards  to  sex  ratio 
(Appendix  1).  Correlations  were  investigated  between  the  female: 
male  ratio  and  the  apparent  age  of  the  nest.  Twelve  P.  montanus 
nests  at  the  peak  levels  of  production  were  used  in  the  analysis.  The 
age  of  a nest  should  be  related  to  the  numbers  of  adult  workers 
present  in  the  nest  and  the  depth  of  the  nest:  older  nests  should  be 
deeper  and  have  a larger  worker  population.  The  product-moment 
correlation  coefficients  (Sokal  and  Rohlf  1969)  of  the  sex  ratio  with 
worker  population  size  and  nest  depths  were  both  0.17.  Although 
the  coefficients  were  not  statistically  significant,  both  were  positive, 
suggesting  that  older  nests  produced  greater  proportions  of  females. 
The  product-moment  correlation  coefficient  comparing  the  sex 
ratio  with  the  numbers  of  workers  produced  by  the  nest  during  the 
year  was  negative  (r  = —0.38).  Although  the  relationship  was  not 
statistically  significant,  it  suggested  that  nests  involved  in  an 
increase  in  the  worker  population  (i.e.,  younger  nests)  produced  a 
smaller  proportion  of  females.  Data  were  presented  (MacKay  1981) 
which  indicated  that  food  stressed  nests  produced  a greater 
proportion  of  females;  nests  given  extra  food  produced  a greater 
proportion  of  males. 


1981] 


Mac  Kay — Nest  Phenologies  of  Pogonomvrmex 


65 


The  factors  influencing  the  determination  of  sex  ratios  in  the 
Hymenoptera  are  currently  of  much  interest  (Herbers  1979). 
Experimental  manipulation  of  food  input  and  excavation  of 
colonies  of  known  age  may  provide  information  on  the  factors 
which  determine  the  sex  ratio  in  a harvester  ant  nest. 

Summary 

This  investigation  compares  the  phenologies  of  foraging  and 
reproduction  in  three  species  of  Pogonomvrmex  harvester  ants 
along  an  altitudinal  transect  in  southern  California,  USA.  Periodic 
excavations  of  126  nests  of  the  three  species,  P.  montanus,  P. 
subnitidus,  and  P.  rugosus,  reveal  that  seasonal  changes  occur 
within  the  nests.  The  three  species  have  similarities  in  the  physical 
environment  of  the  nest  although  P.  montanus,  the  highest  altitude 
species,  has  lower  nest  temperatures.  Both  P.  montanus  and  P. 
subnitidus  are  snowbound  during  part  of  the  season.  Egg  laying 
begins  in  late  April  or  May;  development  to  adult  requires  five  to  six 
weeks.  The  brood  reach  maximum  numbers  in  late  July  to  late 
August.  Most  of  the  larvae  and  pupae  formed  in  the  first  part  of  the 
season  become  reproductives.  Mating  flights  begin  in  late  July  and 
are  completed  by  the  first  part  of  September.  The  highest  reproduc- 
tive populations  occur  in  mid  August. 

Much  of  the  nest  population  is  in  the  upper  levels  of  the  nest 
during  the  summer  and  in  the  lower  levels  during  the  winter.  During 
the  summer,  temperature  and  humidity  gradients  exist  in  the  nests 
with  deeper  levels  being  cooler  and  moister.  These  gradients  may 
account  for  the  placement  of  the  brood  in  the  lower  levels.  There  is 
no  evidence  of  segregation  of  the  larvae  and  pupae  within  the  nest, 
which  has  been  reported  by  other  investigators. 

All  three  species  demonstrate  similar  seasonal  changes  in  foraging 
patterns,  with  much  activity  in  the  mid  summer  and  no  activity 
during  the  winter.  Only  about  20%  of  the  nest  population  is 
involved  in  foraging.  Individual  foragers  make  up  to  9 or  more 
foraging  trips  per  day.  The  ants  utilize  a wide  variety  of  food  items, 
although  most  materials  are  either  seeds  or  plant  parts.  There  is  a 
considerable  seasonal  change  in  the  food  composition  of  P. 
montanus  and  P.  subnitidus. 

The  highest  altitude  species,  P.  montanus,  allocates  more  energy 
to  reproduction  than  do  the  mid  or  low  altitude  species.  The  nests 


66 


Psyche 


[Vol.  88 


invest  about  equally  in  the  production  of  males  and  females. 
Evidence  presented  suggests  that  the  sex  ratio  may  be  ecologically 
determined  and  that  there  may  be  a yearly  change  in  the  sex  ratio. 

Acknowledgements 

I would  like  to  thank  Clay  Sassaman,  Rodolfo  Ruibal,  Robert 
Luck  and  Bert  Holldobler  for  the  critical  review  of  the  manuscript. 
Walter  Whitford  provided  unpublished  information  on  several 
Pogonomvrmex  spp.,  Charles  Kugler  provided  unpublished  data  on 
Pogonomyrmex  mayri.  Jessie  Halverson,  Cecil  Hoff  and  the  U.S. 
Forest  Service  generously  granted  permission  to  conduct  the 
investigation  on  property  under  their  jurisdiction.  I am  especially 
grateful  to  Emma  MacKay,  who  assisted  in  all  aspects  of  the  field 
and  laboratory  work,  prepared  the  figures  and  made  important 
contributions  to  the  manuscript.  Kenneth  Cooper,  Fred  Andrews, 
Gary  Alpert,  David  Kistner,  and  Stewart  Peck  kindly  identified  the 
beetles  found  in  the  ant  nests,  and  provided  much  stimulating 
information  on  the  ecologies  of  the  beetles.  Oscar  Clarke  identified 
the  plant  seeds. 

The  research  was  supported  by  the  Theodore  Roosevelt  Memori- 
al Fund  of  the  American  Museum  of  Natural  History,  three  Grants- 
in-Aid  of  Research  from  Sigma  Xi,  The  Scientific  Research  Society 
of  North  America,  the  Chancellor’s  Patent  Fund  of  the  University 
of  California,  and  the  Irwin  Newell  Award  of  the  Department  of 
Biology  of  the  University  of  California  at  Riverside.  The  Depart- 
ment of  Entomology  of  the  Colegio  de  Graduados  of  Ciudad 
Juarez,  Mexico,  paid  the  costs  of  publication. 

Literature  Cited 
Barnes,  O.  L.  and  N.  J.  Nearney. 

1953.  The  red  harvester  ant  and  how  to  subdue  it.  USDA  Farmers  Bull, 
number  1668,  11  pages. 

Brian,  M.  V. 

1956.  The  natural  density  of  Mvrmica  rubra  and  associated  ants  in  west 
Scotland.  Ins.  Soc.  3:  474-487. 

1965.  Social  insect  pupulations.  Academic  Press,  vii  + 135  pp. 

Brian,  M.  V.,  G.  Elmes,  and  A.  F.  Kelly. 

1967.  Populations  of  the  ant  Tetramorium  caespitum  L.  J.  Anim.  Ecol.  36: 
337-342. 

Bruch,  C. 

1916.  Contribucion  al  estudio  de  las  hormigas  de  la  provincia  de  San  Luis. 
Revista  del  Museo  de  La  Plata,  Argentina.  23:  291-357. 

Burnham,  L. 

1978.  Survey  of  social  insects  in  the  fossil  record.  Psyche  85:  85-134. 


1981] 


Mac  Kay — Nest  Phenologies  of  Pogonomyrmex 


67 


Chew,  R.  M. 

I960.  Note  on  colony  size  and  activity  in  Pogonomyrmex  oecidentalis 
(Cresson).  J.  New  York  Entomol.  Soc.  68:  81  82. 

Clark,  W.  H.  and  P.  L.  Comanor. 

1975.  Removal  of  annual  plants  from  the  desert  ecosystem  by  western 
harvester  ants,  Pogonomyrmex  oecidentalis.  Environ.  Entomol.  4: 
52-56. 

Cloudsley-Thompson,  J.  L. 

1975.  Adaptations  of  arthropods  to  arid  environments.  Ann.  Rev.  Entomol. 
20:  261-283. 

Cole,  A.  C. 

1934.  A brief  account  of  aestivation  and  overwintering  of  the  Occident  ant, 
Pogonomyrmex  oecidentalis  Cresson,  in  Idaho.  Can.  Entomol.  66: 
193-198. 

1954.  Studies  of  New  Mexico  ants.  VII.  The  genus  Pogonomyrmex  with 
synonymy  and  a description  of  a new  species  ( Hymenoptera:For- 
micidae).  J.  Tenn.  Acad.  Sci.  29:  115-121. 

1968.  Pogonomyrmex  harvester  ants,  a study  of  the  genus  in  North  America. 
Knoxville  Univ.  Press,  x + 222  pp. 

Creighton,  W.  S. 

1952.  Studies  on  Arizona  ants  (3).  The  habits  of  Pogonomyrmex  huachucanus 
Wheeler  and  a description  of  the  sexual  castes.  Psyche  59:  71-81. 

Davidson,  D.  W. 

1977a.  Foraging  ecology  and  community  organization  in  desert  seed-eating 
ants.  Ecology  58:  725-737. 

1977b.  Species  diversity  and  community  organization  in  desert  seed-eating  ants. 
Ecology  58:  7 1 1-724. 

1980.  Some  consequences  of  diffuse  competition  in  a desert  ant  community. 
Amer.  Nat.  116:  92-105. 

Edney,  E.  B. 

1974.  Desert  arthropods.  In:  Desert  Biology  Vol.  II,  ed.  by  G.  W.  Brown  Jr. 
pp.  311  384.  Academic  Press,  New  York. 

Erickson,  J.  M. 

1972.  Mark-recapture  techniques  for  population  estimates  of  Pogonomyrmex 
ant  colonies:  An  evaluation  of  the  32  P technique.  Ann.  Entomol.  Soc. 
Am.  65:  57-61. 

Ettershank,  G.  and  W.  G.  Whitford. 

1973.  Oxygen  consumption  of  two  species  of  Pogonomyrmex  harvester  ants 
(Hymenoptera:  Formicidae).  Comp.  Biochem.  Physiol.  46A:  605  611. 

Gentry,  J.  B. 

1974.  Response  to  predation  by  colonies  of  the  Florida  harvester  ant, 
Pogonomyrmex  hadius.  Ecology  55:  1328-1338. 

Gentry,  J.  B.  and  K.  T.  Stiritz. 

1972.  The  role  of  the  Florida  harvester  ant,  Pogonomyrmex  hadius,  in  old  field 
mineral  nutrient  relationships.  Environ.  Entomol.  1:  39-41. 

Golley,  F.  B.  and  J.  B.  Gentry. 

1964.  Bioenergetics  of  the  southern  harvester  ant,  Pogonomyrmex  hadius. 
Ecology  45:  217-225. 


68 


Psyche 


[Vol.  88 


Herbers,  J.  M. 

1979.  The  evolution  of  sex-ratio  strategies  in  hymenopteran  societies.  Amer. 
Nat.  114:  818-834. 

Holldobler,  B. 

1976a.  Recruitment  behavior,  home  range  orientation  and  territoriality  in 
harvester  ants,  Pogonomyrmex.  Behav.  Ecol.  Sociobiol.  1:  3 44. 

1976b.  The  behavioral  ecology  of  mating  in  harvester  ants.  ( Hymenoptera: 
Formicidae:  Pogonomyrmex).  Behav.  Ecol.  Sociobiol.  1:  405  423. 

Jaisson,  P. 

1975.  L’impregnation  dans  l’ontogenese  des  comportements  de  soins  aux 
cocons  chez  la  jeune  fourmi  rousse  ( Formica  polvctena  Eorst.).  Beha- 
viour 52:  1-37. 

Kay,  C.  A.  and  W.  G.  Whitford. 

1975.  Influences  of  temperature  and  humidity  on  oxygen  consumption  of  five 
Chihuahuan  desert  ants.  Comp.  Biochem.  Physiol.  52A:  281  286. 

Kusnezov,  N. 

1951.  El  genero  Pogonomyrmex  Mayr.  Acta  Zoologica  Lilloana  11:  227-333. 

Lavigne,  R. 

1969.  Bionomics  and  nest  structure  of  Pogonomyrmex  occidentalis  (Hymen- 
optera: Formicidae).  Ann.  Entomol.  Soc.  Amer.  62:  1166-1175. 

Mack  ay,  W.  P. 

1980.  A new  harvester  ant  from  the  mountains  of  southern  California 
(Hymenoptera:  Formicidae).  Southwest.  Nat.  25:  151-156. 

1981.  A comparison  of  the  ecological  energetics  of  three  species  of  Pogono- 
myrmex harvester  ants  (Hymenoptera:  Formicidae).  Ph.D.  dissertation. 
University  of  California  at  Riverside.  240  pp. 

Marcus,  H. 

1953.  Estudios  mirmecologicos  11.  Observaciones  en  Pogonomyrmex  marcusi 
(Kusn.).  Folia  Universitaria  de  la  Universidad  de  Cochabamba  6:  45  55. 

Marcus,  H.  and  E.  E.  Marcus. 

1951.  Los  nidos  y los  organos  de  estridulacion  y de  equilibrio  de  Pogono- 
myrmex marcusi  y de  Dorymyrmex  emmaericaellus.  (Kusn.).  Folia 
Universitaria,  Universidad  Mayor  de  San  Simon,  Cochabamba,  Bolivia 
5:  117-143. 

Mares,  M.  A.  and  M.  L.  Rosenzweig. 

1978.  Granivory  in  North  and  South  American  deserts:  rodents,  birds,  and 
ants.  Ecology  59:  235-241. 

Marikovsky,  P.  1. 

1962.  On  intraspecific  relations  of  Formica  rufa  L.  Ent.  Rev.  1:  47-51. 

Markl,  H.,  B.  Holldobler  and  T.  Holldobler. 

1977.  Mating  behavior  and  sound  production  in  harvester  ants  {Pogono- 
myrmex, Formicidae).  Ins.  Soc.  24:  191  212. 

Marshall,  T.  J. 

1959.  Relations  between  water  and  soil.  Technical  Communication  Number 
50.  Commonwealth  Bureau  of  Soils,  Harpenden. 

Marshall,  T.  J.  and  J.  W.  Holmes. 

1979.  Soil  physics.  Cambridge  University  Press.  * + 345  pp. 


1981] 


Mac  Kay — Nest  Phenologies  of  Pogonomyrmex 


69 


Nielsen,  M.  G. 

1972.  Production  of  workers  in  an  ant  nest.  Ekologia  Polska  20:  65-71. 
Nielsen,  M.  G. 

Number  and  biomass  of  worker  ants  in  a sandy  area  in  Denmark. 
Natura  Jutlandica  17:  93-95. 

Nielsen,  M.  G.  and  T.  F.  Jensen. 

1975.  Ecological  studies  on  Lasius  alienus  ( Forst)  ( Hymenoptera:  Formicidae). 
Entomologiske  Meddelelser  43:  5-16.  (In  Danish). 

O’Dowd,  D.  J.  and  M.  E.  Hay. 

1980.  Mutualism  between  harvester  ants  and  a desert  ephemeral:  seed  escape 
from  rodents.  Ecology  61:  531-540. 

Odum,  E.  P.  and  A.  J.  Pontin. 

1961.  Population  density  of  the  underground  ant,  Lasius  flavus  as  determined 
by  tagging  with  P22.  Ecology  42:  186-188. 

Peck,  S.  B. 

1976.  The  myrmecophilous  beetle  genus  Echinoeoleus  in  the  southwestern 
United  States  (Leiodidae,  Catopinae).  Psyche  83:  51-62. 

Petal,  J. 

1967.  Productivity  and  the  consumption  of  food  in  the  Myrmica  laevinodis 
Nyl.  population.  Proceedings  of  the  International  Biological  Program 
on  secondary  productivity  of  terrestrial  ecosystems.  K..  Petrusewicz  (ed). 
pp.  841  857. 

Protomastro,  J.  J. 

1973.  Relacion  entre  el  comportamiento  de  crianza  y un  ritmo  diario  de 
preferencia  termica  en  la  hormiga  Camponotus  mus  Roger.  Physis, 
Seccion  C.  32:  123-128. 

Reichmann,  O.  J. 

1979.  Desert  granivore  foraging  and  its  impact  on  seed  densities  and  distribu- 
tions. Ecology  60:  1085  1092. 

Rode,  A.  A. 

1955.  The  moisture  properties  of  soils  and  underground  strata.  Academy  of 
Sciences  of  the  USSR,  Moscow.  117  pp. 

Rogers,  L.,  R.  Lavigne  and  J.  L.  Miller. 

1972.  Bioenergetics  of  the  western  harvester  ant  in  the  shortgrass  plains 
ecosystem.  Environ.  EntomoU  1:  763-768. 

SOKAL,  R.  R.  AND  F.  J.  ROHLF. 

1969.  Biometry.  W.  H.  Freeman  and  Co.  xxi  + 776  pp. 

Snedecor,  G.  W.  and  W.  G.  Cochran. 

1967.  Statistical  Methods.  6th  edition.  Iowa  State  University  Press,  xiv  + 593 

pp. 

Taylor,  F.  W. 

1977.  Foraging  behavior  of  ants:  Experiments  with  two  species  of  Myrmicine 
ants.  Behav.  Ecol.  Sociobiol.  2:  147  167. 

Waloff,  N.  and  R.  W.  Blackith. 

1962.  The  growth  and  distribution  of  the  mounds  of  Lasius  flavus  (Fabricius) 
(Hym:  Formicidae)  in  Silwood  Park,  Berkshire.  J.  Anim.  Ecol.  31: 
421  437. 


70 


Psyche 


[Vol.  88 


Wheeler,  G.  L.  and  J.  Wheeler. 

1976.  Ant  larvae:  review  and  synthesis.  Mem.  Entomol.  Soc.  Wash.  #7,  vi  + 

108  pp. 

Wheeler,  W.  M. 

1910.  Ants,  their  structure,  development  and  behavior.  Columbia  University 
Press,  xxv  + 663  pp. 

1914.  New  and  little  known  harvesting  ants  of  the  genus  Pogonomyrmex. 
Psyche  21:  149-157. 

Whitford,  W.  G. 

1 972.  Demography  and  bioenergetics  of  herbivorous  ants  in  a desert  ecosystem 
as  functions  of  vegetation,  soil  type  and  weather  variables.  Unpublished 
manuscript. 

1976.  Foraging  behavior  of  Chihuahuan  desert  harvester  ants.  Amer.  Midi. 
Nat.  95:  455-458. 

1978a.  Foraging  in  seed-harvester  ants  Pogonomyrmex  spp.  Ecology  59: 
185-189. 

1978b.  Structure  and  seasonal  activity  of  Chihuahua  desert  ant  communities. 
Ins.  Soc.  25:  79-88. 

Whitford,  W.  G.  and  M.  Bryant. 

1979.  Behavior  of  a predator  and  its  prey:  the  horned  lizard  ( Phrynosoma 
cornutum)  and  harvester  ants  ( Pogonomyrmex  spp.)  Ecology  60: 
686-694. 

Whitford,  W.  G.  and  G.  Ettershank. 

1975.  Factors  affecting  foraging  activity  in  Chihuahuan  desert  harvester  ants. 
Environ.  Entomol.  : 689-696 

Whitford,  W.  G.,  P.  Johnson  and  J.  Ramirez. 

1976.  Comparative  ecology  of  the  harvester  ants  Pogonomyrmex  barbatus  (F. 
Smith)  and  Pogonomyrmex  rugosus  (Emery).  Ins.  Soc.  23:  117-132. 

Wildermuth,  V.  L.  and  E.  G.  Davis. 

1931.  The  red  harvester  ant  and  how  to  subdue  it.  U.S.D.A.  Farmers’  Bull. 

#1668,  21  pp. 

Willard,  J.  R.  and  H.  H.  Crowell. 

1965.  Biological  activities  of  the  harvester  ant,  Pogonomyrmex  owyheei  in 
central  Oregon.  J.  Econ.  Entomol.  58:  484-489. 

Wilson,  E.  O. 

1971.  The  insect  societies.  Belknap  Press,  x + 548  pp. 

Appendix  1.  List  of  the  populations  of  the  excavated  Pogonomyrmex  spp.  nests, 
including  workers  (W),  larvae  (L),  pupae  (P),  callows  (C),  males  (M),  and  females 
(F).  The  position  of  the  queen  and  maximum  depth  of  the  nests  are  expressed  in 
centimeters.  The  dates  indicated  are  when  the  excavation  was  begun. 


Position 

Date  W L P C F M of  queen  Depth 

Pogonomyrmex  montanus 

21  Sept  77  793  7 514  0 0 0 50  50 

23  Sept  77  1918  123  578  0 0 0 40  40 


1981]  MacKay — Nest  Phenologies  of  Pogonomyrmex 


71 


Appendix  1.  cont. 


Position 

Date  W L P C F M of  queen  Depth 


1  1 Apr  78 
13  Apr  78 


16  Apr  78 

3 May  78 

8 May  78 

9 May  78 
18  May  78 

25  May  78 
31  May  78 

6 June  78 

8 June  78 
13  June  78 

13  June  78 

23  June  78 
27  June  78 

7 July  78 
12  July  78 
20  July  78 

24  July  78 

4 Aug  78 
1 1 Aug  78 

18  Aug  78 

22  Aug  78 
30  Aug  78 

9 Sept  78 

19  Sept  78 

23  Sept  78 

23  Sept  78 

24  Sept  78 

26  Sept  78 

26  Sept  78 
30  Sept  78 
30  Sept  78 
30  Sept  78 

7 Oct  78 

14  Oct  78 
16  Oct  78 
16  Oct  78 

27  Oct  78 
7 May  79 
7 May  79 

25  May  79 
25  May  79 


3142 

0 

1654 

0 

1874 

0 

2372 

0 

2240 

0 

3641 

0 

2277 

0 

1752 

0 

1951 

0 

491 

0 

1598 

83 

499 

0 

1841 

51 

695 

7 

1087 

141 

1488 

359 

1652 

367 

1604 

426 

1033 

228 

996 

41 1 

1013 

290 

1158 

277 

1057 

244 

1552 

306 

1113 

14 

1443 

54 

631 

1 

2271 

10 

2573 

4 

634 

0 

1734 

5 

1785 

24 

1024 

0 

652 

15 

1245 

16 

2105 

20 

1343 

24 

538 

49 

2812 

4 

1304 

0 

3585 

0 

2194 

0 

3141 

0 

0 0 

0 0 

0 0 

0 0 

0 0 

0 0 

0 0 

0 0 

0 0 

0 0 

0 0 

0 0 

0 0 

0 0 

38  0 

160  0 

177  0 

477  0 

131  0 

678  438 

323  496 

358  529 

359  784 

330  766 

163  254 

73  281 

0 80 

2 128 

6 96 

0 40 

3 117 

12  52 

0 13 

1 1 1 

0 1 1 

3 63 

0 0 

0 0 

0 4 

0 0 

0 0 

0 0 

0 0 


0 0 

0 0 

0 0 

0 0 

0 0 

0 0 

0 0 

0 0 

0 0 

0 0 

0 0 

0 0 

0 0 

0 0 

0 0 

0 0 

0 0 

188  54 

52  138 

64  87 

262  0 

147  148 

48  327 

111  405 

162  83 

81  25 

0 0 

1 0 

0 1 

0 I 

0 0 

0 0 

0 0 

0 0 

0 0 

0 0 

0 0 

0 0 

1 0 

0 0 

0 0 

0 0 

0 0 


50 

50 

40 

50 

* 

40 

38 

70 

36 

60 

55 

70 

65 

80 

43 

60 

75 

75 

32 

50 

47 

59 

* 

30 

10 

65 

55 

60 

* 

70 

39 

80 

75 

75 

68 

68 

40 

70 

39 

53 

80 

82 

60 

70 

40 

82 

40 

105 

40 

75 

50 

68 

40 

63 

70 

83 

60 

73 

* 

1 10 

* 

84 

60 

64 

59 

59 

** 

86 

* 

74 

80 

92 

** 

43 

** 

40 

50 

76 

30 

53 

* 

61 

50 

70 

40 

80 

72 


Psyche 


[Vol.  88 


Appendix  1.  cont. 


Position 


Date 

W 

L 

P 

C 

F 

M 

of  queen 

Depth 

3 June  79 

1605 

0 

0 

0 

0 

0 

** 

40 

15  June  79 

1979 

0 

0 

0 

0 

0 

30 

70 

20  June  79 

537 

63 

0 

0 

0 

0 

* 

50 

20  June  79 

1768 

0 

0 

0 

0 

0 

20 

60 

8 July  79 

599 

193 

67 

0 

0 

0 

50 

60 

16  July  79 

563 

84 

44 

0 

28 

8 

50 

50 

22  July  79 

2614 

1318 

750 

0 

86 

290 

50 

60 

29  July  79 

947 

813 

710 

305 

198 

2 

40 

50 

29  July  79 

1777 

383 

708 

7 

286 

146 

70 

80 

30  July  79 

369 

27 

21 

137 

15 

2 

* 

80 

30  July  79 

1 1 1 1 

835 

693 

13 

149 

51 

60 

60 

3 Aug  79 

1853 

325 

490 

609 

601 

37 

* 

120 

4 Aug  79 

1261 

483 

1049 

595 

305 

543 

* 

100 

4 Aug  79 

1625 

555 

1013 

225 

221 

277 

60 

80 

12  Aug  79  (a) 

976 

345 

534 

752 

167 

12 

50 

60 

12  Aug  79  (b) 

640 

124 

231 

701 

144 

0 

30 

80 

13  Aug  79  (a) 

1063 

619 

529 

495 

390 

11 

* 

50 

13  Aug  79  (b) 

697 

244 

466 

660 

387 

85 

40 

70 

18  Aug  79 

755 

234 

719 

729 

41 

113 

60 

70 

19  Aug  79  (c) 

370 

4 

0 

107 

221 

4 

50 

70 

19  Aug  79  (c) 

520 

108 

641 

322 

72 

262 

* 

50 

19  Aug  79  (d) 

836 

168 

632 

1068 

142 

55 

80 

80 

20  Aug  79  (d) 

792 

81 

401 

655 

141 

0 

40 

80 

20  Aug  79  (e) 

596 

169 

0 

36 

33 

318 

50 

50 

20  Aug  79  (e) 

1079 

574 

839 

289 

54 

436 

50 

60 

7 Sept  79 

1796 

468 

5 

931 

62 

1 

80 

110 

8 Sept  79 

711 

67 

19 

65 

0 

0 

30 

30 

8 Sept  79 

808 

126 

112 

559 

0 

0 

70 

70 

14  Oct  79 

1308 

35 

0 

0 

0 

0 

60 

60 

5 Dec  79 

1504 

0 

0 

0 

0 

0 

50 

70 

19  Aug  80 

1039 

nr 

nr 

780 

216 

309 

** 

nr 

20  Aug  80 

1388 

nr 

nr 

501 

255 

343 

** 

nr 

21  Aug  80 

1346 

nr 

nr 

144 

53 

274 

** 

nr 

21  Aug  80 

1687 

nr 

nr 

351 

39 

634 

** 

nr 

22  Aug  80 

1800 

nr 

nr 

185 

310 

516 

** 

nr 

Pogonomyrmex  subnitidus 

5 Nov  78 

5033 

47 

0 

0 

0 

0 

150 

210 

29  Aug  79 

2507 

894 

1212 

2819 

0 

0 

140 

170 

15  Sept  79 

3612 

867 

892 

4147 

0 

0 

120 

150 

7 Oct  79 

5442 

109 

321 

999 

0 

0 

220 

230 

3 Nov  79 

3452 

56 

0 

315 

0 

0 

160 

230 

30  Nov  79 

5182 

0 

0 

0 

0 

0 

180 

240 

1981] 


Mac  Kay — Nest  Phenologies  of  Pogonomyrmex 


73 


Appendix  1.  cont. 


Position 


Date 

W 

L 

P 

C 

F 

M 

of  queen 

Depth 

25  Jan  80 

6864 

0 

0 

0 

0 

0 

130 

210 

12  Mar  80 

13056 

0 

0 

0 

0 

0 

100 

200 

16  May  80 

5009 

0 

0 

0 

0 

0 

* 

210 

22  May  80 

7687 

0 

0 

0 

0 

0 

150 

150 

6 June  80 

4962 

0 

0 

0 

0 

0 

300 

300 

12  June  80 

8679 

0 

0 

0 

0 

0 

* 

270 

30  June  80 

10160 

531 

0 

0 

0 

0 

160 

180 

8 July  80 

3515 

470 

153 

0 

0 

0 

* 

260 

13  July  80 

2440 

784 

541 

0 

0 

0 

260 

260 

17  July  80 

1784 

609 

579 

66 

0 

0 

* 

140 

19  July  80 

9385 

1238 

767 

0 

368 

270 

* 

280 

22  July  80 

3619 

1977 

1 139 

143 

47 

94 

270 

280 

29  July  80 

5215 

1817 

1301 

570 

6 

66 

260 

260 

4 Aug  80 

4060 

850 

1 127 

314 

52 

554 

* 

300 

7 Aug  80 

4734 

1816 

1646 

794 

80 

271 

190 

250 

12  Aug  80 

2420 

559 

953 

1207 

1 

8 

1 10 

150 

13  Aug  80 

2362 

750 

484 

1355 

0 

0 

60 

230 

14  Aug  80 

3877 

2145 

1300 

1305 

0 

0 

270 

270 

23  Aug  80 

4168 

2301 

2062 

2524 

0 

0 

250 

280 

4 Sept  80 

6901 

1359 

1567 

1591 

2 

49 

270 

270 

Pogonomyrmex 

rugosus 

19  Nov  78 

4569 

0 

0 

0 

0 

0 

1 10 

200 

25  Mar  79 

3707 

0 

0 

0 

0 

0 

30 

130 

3 Apr  79 

3778 

0 

0 

0 

0 

0 

120 

240 

8 Apr  79 

14742 

0 

0 

0 

0 

0 

100 

360 

20  Apr  79 

3115 

0 

0 

0 

0 

0 

* 

270 

27  Apr  79 

11802 

0 

0 

0 

0 

0 

70 

125 

1 1 May  79 

7275 

0 

0 

0 

0 

0 

10 

160 

28  May  79 

10588 

0 

0 

0 

0 

0 

30 

160 

6 June  79 

8033 

532 

37 

0 

0 

0 

* 

300 

20  June  79 

9214 

141 1 

594 

22 

0 

0 

170 

190 

2 July  79 

2485 

1 136 

1057 

101 

5 

505 

240 

260 

10  July  79 

9374 

1522 

1368 

21 

0 

0 

270 

280 

24  July  79 

3086 

1753 

1693 

838 

418 

219 

160 

180 

6 Aug  79 

5648 

2440 

3072 

1501 

38 

181 

* 

300 

31  Aug  79 

7219 

1839 

1565 

365 

9 

342 

160 

270 

28  Sept  79 

1 1640 

2204 

2483 

1440 

0 

0 

350 

360 

27  Oct  79 

4655 

465 

633 

305 

0 

0 

280 

280 

7 Dec  79 

10538 

66 

0 

0 

0 

0 

370 

400 

1 Feb  80 

7503 

0 

0 

0 

0 

0 

40 

210 

23  Feb  80 

11239 

0 

0 

0 

0 

0 

* 

300 

74 


Psyche 


[Vol.  88 


Appendix  1.  cont. 

* Nest  queen  not  found. 

**  Nest  queen  found  but  level  not  recorded. 

(a)  Nest  received  extra  food  in  June  1979. 

(b)  Nest  received  extra  food  in  July  1979. 

(c)  Nest  received  less  food  throughout  1979  season. 

(d)  Control  nest. 

(e)  Nest  received  extra  food  throughout  1979  season. 
See  MacKay  (1981)  for  further  details. 

nr=  not  recorded. 


LABORATORY  EVALUATION  OF  WITHIN-SPECIES, 
BETWEEN-SPECIES,  AND  PARTHENOGENETIC 
REPRODUCTION  IN  RETICULITERMES  FLAVIPES 
AND  RETICU  LITE  RATES  VIRGINICUS 1 

By  Ralph  W.  Howard,2,3  Eldon  J.  Mallette,2 3 
Michael  I.  Haverty,4  and  Richard  V.  Smythe5 

Introduction 

Considerable  interest  currently  exists  regarding  the  reproductive 
strategies  of  social  insects  (Blum  and  Blum,  1979;  Crozier,  1979). 
Among  termites  (Order  Isoptera)  colony  foundation  by  alate  pairs, 
fusion  of  existing  colonies,  splitting  of  existing  colonies,  and 
parthenogenesis  have  all  been  reported  (Nutting,  1969).  Little 
information  is  available  regarding  the  relative  importance  of  each  of 
these  strategies. 

The  genus  Reticulitermes  (Rhinotermitidae)  contains  six  Nearctic 
and  twelve  Palearctic  species,  three  of  which  have  been  critically 
examined  for  reproductive  modes.  Pickens  (1932)  and  Weesner 
(1956)  studied  colony  foundation  of  R.  hesperus  Banks  by  male  + 
female  dealate  pairs,  as  well  as  by  parthenogenesis.  Buchli  (1950) 
studied  similar  strategies  for  R.  lucifugus  Rossi.  Clement  (1979) 
studied  interspecific  hybridization  of  R.  santonensis  Feytaud  and  R. 
lucifugus.  More  limited  studies  on  colony  foundation  by  male  + 
female  dealate  pairs  of  R.flavipes  (Kollar)  were  conducted  by  Beard 
(1974). 

Field  studies  with  R.  flavipes  (Howard  and  Haverty,  1980) 
suggest  that  an  important  reproductive  strategy  for  this  species  is 
colony  splitting  with  subsequent  production  of  numerous  (several 
hundred)  neotenic  reproductives.  However,  sizeable  alate  flights  are 
also  a prominent  feature  of  the  biology  of  Reticulitermes  spp.  and 


1 Manuscript  received  by  the  editor  July  6,  1981. 

2Forestry  Sciences  Laboratory,  Southern  Forest  Experiment  Station,  P.  O.  Box  2008 
GMF,  Gulfport,  MS  39503. 

3Author  to  whom  correspondence  should  be  addressed. 

4Pacific  Southwest  Forest  and  Range  Experiment  Station,  P.  O.  Box  245,  Berkeley, 
CA  94702. 

5North  Central  Forest  and  Range  Experiment  Station,  1992  Folwell  Avenue,  St. 
Paul,  MN  55108. 


75 


76 


Psvche 


[Vol.  88 


suggest  that  alate-based  reproductive  strategies  may  also  be  impor- 
tant. Field  studies  of  colony  foundation  by  alate  pairs  are  extremely 
difficult.  We  have  accordingly  chosen  to  investigate  by  laboratory 
studies  the  potential  of  alate  based  reproductive  modes.  We  have 
examined  incipient  colony  formation  with  males  and  females  of  the 
same  species,  males  and  females  of  different  species,  and  pairs  of 
conspecific  females.  Our  results  are  reported  here. 

Methods  and  Materials 


Termites 

Unflown  alates  were  collected  from  fallen  logs  in  the  De  Soto 
National  Forest  in  southern  Mississippi.  R.  flavipes  alates  were 
collected  from  mid-March  to  early  April  of  1968,  1969,  and  1970 
(one  source  colony  each  year)  and  from  mid-September  to  early 
October  of  1967,  1968,  and  1969  (one  source  colony  each  year).  R. 
virginicus  alates  were  collected  in  mid- April  to  mid-May  of  1968, 
1969,  and  1970  (one  source  colony  each  year).  At  least  500 
alates/ source  colony  were  anesthetized  with  CO2  (200  ml/ min), 
sexed,  placed  in  separate  petri  dishes  lined  with  moistened  filter 
paper,  and  transferred  to  a dark  incubator  at  25  +1°C  for  less  than 
one  week.  R.  flavipes  alates  to  be  paired  with  R.  virginicus  alates 
were  held  in  their  source  wood  for  up  to  30  days  in  an  incubator  at 
15  +1°C  until  R.  virginicus  alates  were  available.  All  alates  were 
allowed  to  lose  their  wings  naturally  before  pairing. 

Pairing  Procedures 

Each  experimental  unit  consisted  of  an  8.3-  X 12.7-cm  piece  of 
single  strength  window  glass  to  which  2.5  mm  X 1 cm  strips  of 
plexiglass  had  been  glued  to  form  a 6.3-  X 8.9-cm  rectangular  cell 
(Howard,  1980).  Washed  and  ovendried  sand  was  placed  in  the 
upper  third  of  the  cell  and  moistened  with  deionized  water.  Two  2- 
X 4-cm  X 1-  to  2-mm  weathered  strips  of  southern  pine  were  gently 
inserted  about  1 cm  apart  into  the  border  of  the  moistened  sand. 
Termites  were  placed  in  the  cell,  the  cell  was  closed  by  covering  the 
opening  with  four  2.5-  X 7.5-cm  microscopic  slides,  and  then  the 
cell  was  sealed  along  the  edges  with  hot  paraffin.  A small  opening 
(about  1 mm)  was  left  to  allow  for  air  exchange. 

Dealated  termites  were  randomly  selected,  paired,  and  placed  in 
an  experimental  unit.  Each  unit  was  examined  daily  during  the  first 
week  and  any  termites  caught  in  condensed  moisture  were  freed. 


1981]  Howard,  Mallet  te,  Haver  tv,  & S my  the — Reticulitermes  77 


Also  during  this  period  dead  dealates  were  replaced  with  live  ones. 
After  the  first  week  dead  dealates  were  removed  but  not  replaced. 
Subsequent  inspections  were  made  approximately  three  times  a 
week  for  the  following  month,  then  two  times  a week  for  3 months, 
once  a week  for  2 months,  and  once  every  2 weeks  thereafter  until 
the  dealates  were  dead  or  the  experiment  was  terminated.  At  each 
inspection  the  number  of  live  dealates,  eggs,  larvae  (by  stage), 
presoldiers,  and  soldiers  were  recorded. 

Seven  combinations  involving  the  following  pairings  were  exam- 
ined: (1)  R.  flavipes  male  + R.  flavipies  female  (spring);  (2)  R. 
flavipes  male  + R.  flavipes  female  (fall);  (3)  R.  virginicus  male  + 
R.  virginicus  female;  (4)  R.  flavipes  male  + R.  virginicus  female;  (5) 
R.  virginicus  male  + R.  flavipes  female;  (6)  R.  flavipes  female  + R. 
flavipes  female;  and  (7)  R.  virginicus  female  + R.  virginicus  female. 
Each  combination  was  replicated  a minimum  of  ten  times  per  year. 

Determination  of  Larval  Stage 

Larvae  from  six  of  the  seven  combinations  were  randomly 
selected  from  at  least  five  of  the  experimental  units  in  each 
combination  and  placed  in  Bouin’s  solution.  Measurements  were 
then  made  of  the  number  of  antennal  segments  and  head  width. 
These  data  were  used  for  determining  the  instars  of  live  larvae 
through  the  fourth  stage. 

Summarization  of  Data 

For  each  of  the  seven  combinations  the  following  two  kinds  of 
data  were  gathered:  (1)  mean  number  of  days  to  the  first  appearance 
of  an  egg,  1st-,  2nd-,  3rd-,  and  4th-stage  larva,  presoldier  and 
soldier;  and  (2)  mean  caste  composition  of  incipient  colonies  at 
selected  intervals.  Not  all  colonies  were  established  nor  observed  on 
the  same  days.  To  make  the  various  pairings  comparable,  dates  of 
establishment  and  observation  were  converted  to  Julian  dates. 
Observations  were  then  grouped  by  the  number  of  days  past  the 
data  of  pairing.  Results  of  pairings  replicated  over  2 or  3 years  were 
combined.  To  simplify  data  summarization,  observations  were 
grouped  in  3-day  intervals.  Because  of  the  different  times  of 
observation,  not  all  of  the  colonies  were  included  in  each  3-day 
interval,  giving  rise  to  fluctuations  in  the  number  of  colonies 
included  in  the  summarized  data. 

When  both  dealates  had  died,  colonies  were  removed  from  the 
experiment.  Some  of  the  colonies  remained  viable  for  more  than  2 


78 


Psyche 


[Vol.  88 


years.  However,  when  the  number  of  colonies  included  in  any  3-day 
interval  for  any  of  the  combinations  dropped  below  five,  we 
discontinued  summarization  of  the  data  for  that  combination,  since 
average  caste  compositions  based  on  less  than  five  colonies  would 
have  little  meaning. 

Examination  of  Symbiotic  Protozoa 

Fourth  stage  or  older  larvae  from  five  of  the  R.  virginicus  male  + 
R.  flavipes  female  pairing  and  from  four  of  the  R.  flavipes  male  + 
R.  viginicus  female  pairing  were  examined  for  the  presence  of 
species-characteristic  protozoa.  R.  flavipes  contains  mainly  Dine- 
nympha  fimbriata  Kirby,  D.  gracilis  Leidy,  Pyrsonympha  major 
Powell,  P.  vertens  Leidy,  Spirotrichonympha  flagellata  (Grassi)  and 
Trichonympha  agilis  Leidy.  R.  virginicus  contains  fewer  species  of 
protozoa,  with  the  primary  species  being  D.  fimbriata,  P.  minor,  S. 
flagellata,  and  T.  agilis  (Yamin,  1979). 

Results 

Egg  and  Larval  Development 

The  time  to  the  first  appearance  of  eggs  and  larval  instars  was 
rather  similar  for  four  of  the  seven  combinations  examined  ( R . 
flavipes  male  + R.  flavipes  female,  spring;  R.  virginicus  male  + R. 
virginicus  female;  R.  virginicus  male  + R.  flavipes  female;  and  R. 
virginicus  female  + R.  virginicus  female;  see  Table  1).  Egg 
production  began  in  these  combinations  about  8 to  15  days  after 
pairing,  and  the  first  larvae  were  produced  in  about  38  to  45  days. 
Fourth  stage  larvae  were  present  by  about  75  days  after  pairing. 

In  contrast,  the  fall  R.  flavipes  male  + R.  flavipes  female 
combination  required  almost  25  days  for  egg  production  to  begin 
and  almost  60  days  for  the  first  larva  to  be  produced.  Fourth  stage 
larvae  were  present  by  83  days  after  pairing.  The  combination  of  R. 
flavipes  male  + R.  virginicus  female  began  egg  production  within  15 
days  of  pairing,  but  only  one  larva  was  ever  produced  from  this 
cross.  This  larva  appeared  normal  and  successfully  molted  twice 
(Table  1).  Although  the  R.  flavipes  female  + R.  flavipes  female 
combination  was  prolific  in  egg  production,  none  of  these  eggs  ever 
hatched. 

Soldier  Production 

The  proportion  of  incipient  colonies  producing  soldiers  in  the 


1981]  Howard,  Mallette,  Haverty,  & Smythe—Reticulitermes  79 


Table  1.  Mean  number  of  days  until  the  appearance  of  the  first  egg  and  first 
individual  of  an  instar  in  incipient  colonies  of  different  parentage 


Instar 

Parentage1 

Egg 

1 

2 

3 

4 

Rf^  + Rf$  (spring) 

8.1 

38.2 

46.1 

52.9 

74.7 

Rf$  + Rf$  (fall) 

24.8 

59.7 

65.5 

73.1 

83.3 

RvS  + 

11.5 

40.2 

50.5 

58.6 

65.1 

Rf(5  + Rv9 

15.0 

48.5 

60. 42 

70. 62 

3 

R vS  + Rf? 

11.7 

42.2 

50.5 

57.0 

3 

Rf?  + Rf? 

7.0 

4 

4 

4 

4 

Rv?  + Rv? 

14.6 

45.2 

57.8 

64.6 

3 

1 Reticulitermes  flavipes;  Reticulitermes  virginicus. 
2Values  from  one  experimental  unit  only. 

3Not  measured. 

4No  larvae  ever  produced. 


various  combinations  varied  from  8 to  58.8  percent  (Table  2).  The 
R.  flavipes  male  + R.  flavipes  female  (spring)  combination  produced 
only  about  one-third  as  many  colonies  with  soldiers  as  did  the  R. 
flavipes  male  + R.  flavipes  female  (fall)  combination.  The  R. 
virginicus  male  T R.  virginicus  female  combination,  however, 
produced  approximately  the  same  proportion  of  colonies  with 
soldiers  as  did  the  R.  virginicus  female  + R.  virginicus  female 
combination. 

Temporal  Colony  Composition  and  Numbers 
The  relative  proportions  of  eggs  and  larvae  in  the  R.  flavipes  male 
+ R.  flavipes  female  (spring)  combination  (Fig.  1 A),  R.  virginicus 


Table  2.  Soldier  production  in  incipient  colonies  of  different  parentage 


Parentage1 

Number  of  colonies 
producing  soldiers 

Total  number 
of  colonies 

Percent  of  colonies 
producing  soldiers 

Rf<3  + Rf?  (spring) 

14 

69 

20.3 

Rf S + Rf?  (fall) 

20 

34 

58.8 

Rv<3  + Rv$ 

6 

75 

8.0 

RiS  + Rv2$ 

1 

85 

1.2 

R v$+  Rf$ 

22 

85 

25.9 

Rf S + Rf3$ 

0 

13 

0 

Rv$  + Rv$ 

2 

18 

11.1 

1 Reticulitermes  flavipes;  Reticulitermes  virginicus. 

2Only  one  pair  produced  a larva. 

3No  eggs  hatched  in  this  crossing. 


80 


Psyche 


[Vol.  88 


CM 


• * * 

• * * 


• ••• 


•••• 


•>/A* 

* * v*  • 

/* 


CD' 


oc 

LU 

m 


z 
< 
LU  CM 


• . 

o o 

COLONY  AGE.  DAYS 


O 

O 


Figure  1 . Mean  number  of  eggs  and  larvae  produced  by  incipient  colonies  headed 
by  male  and  female  dealates  of  the  same  species.  A.  R.  flavipes  (spring);  B.  R. 
flavipes  (fall);  C.  R.  virginicus.  Legend:  * eggs,  + larvae.  Each  point  is  a mean  of 
at  least  five  replicates. 


MEAN  NUMBER 


1981]  Howard,  Mallette,  Haverty,  & Smythe — Reticulitermes  81 


Figure  2.  Mean  number  of  eggs  and  larvae  produced  by  incipient  colonies  headed 
by  male  and  female  dealates  of  different  species.  A.  R.  virginicus  male  + R. 
flavipes  female;  B.  R.  flavipes  male  + R.  virginicus  female.  Legend:  * eggs, 
+ larvae.  Each  point  is  a mean  of  at  least  five  replicates. 


rOOL  T00Z 


MEAN  NUMBER 


82 


Psyche 


[Vol.  88 


o 

CM 


• • 


o o 

CD  CD 


CD  00  O CM 

COLONY  AGE,  DAYS~ 


Figure  3.  Mean  number  of  eggs  and  larvae  produced  by  incipient  colonies  headed 
by  two  female  dealates  of  the  same  species.  A.  R.  virginicus;  B.  R.  flavipes. 
Legend:  * eggs,  + larvae.  Each  point  is  a mean  of  at  least  five  replicates. 


100 


1981]  Howard,  Mallette,  Haverty,  & Smythe — Reticulitermes  83 

male  + R.  virginicus  female  combination  (Fig.  1C),  R.  virginicus 
male  + R.  flavipes  female  combination  (Fig.  2A),  and  R.  virginicus 
female  + R.  virginicus  female  combination  (Fig.  3 A)  were  similar  at 
all  time  intervals  examined.  A maximum  mean  number  of  six  to 
eight  eggs  were  present  by  day  40  and  a maximum  mean  number  of 
six  to  eight  larvae  were  present  by  60  to  100  days.  The  fall  R. 
flavipes  male  + R.  flavipes  female  combination  produced  similar 
numbers  of  eggs  and  larvae  (Fig.  1 B)  as  the  above  combinations,  but 
at  a slower  rate. 

In  contrast,  the  R.  flavipes  male  + R.  virginicus  female  combina- 
tion (Fig.  2B)  and  the  R.  flavipes  female  + R.  flavipes  female 
combination  (Fig.  3B)  each  produced  a mean  of  up  to  15  eggs  within 
the  first  40  to  60  days,  but  produced  essentially  no  larvae. 

Symbiotic  Protozoa  in  Progeny  of  Between-Species  Pairings 

R.  flavipes  male  + R.  virginicus  female:  Ten  larvae  beyond  the 
third  instar  from  four  experimental  units  were  examined  for 
protozoa.  Seven  larvae  contained  protozoa,  and  of  these,  two 
contained  protozoa  typical  of  R.  flavipes,  two  contained  protozoa 
typical  of  R.  virginicus,  and  the  remaining  three  contained  mixtures 
of  protozoa  characteristics  of  both  termite  species. 

R.  virginicus  male  + R.  flavipes  female:  Seventeen  larvae  from 
five  experimental  units  were  examined  for  protozoa.  All  larvae 
contained  protozoa.  Three  contained  protozoa  typical  of  R.  fla- 
vipes, five  contained  protozoa  typical  of  R.  virginicus,  and  the 
remaining  nine  contained  mixtures  of  protozoa  typical  of  both 
termite  species. 


Discussion 

Successful  incipient  colony  foundation  by  male  and  female 
dealates  of  R.  flavipes  and  R.  virginicus  occurred  readily  in  the 
laboratory.  The  young  colonies  were  provided  with  abundant  food, 
plentiful  water,  an  absence  of  predators,  and  near  optimum 
temperatures.  Despite  this,  the  growth  rate  of  all  colonies  was  slow, 
with  no  more  than  20  to  30  larvae  being  produced  within  the  first 
year.  These  results  agree  closely  with  published  laboratory  data  on 
several  other  rhinotermitids.  Buchli  (1950)  obtained  ca.  30  individ- 
uals from  R.  lucifugus  dealate  pairs  after  8 months,  Weesner  ( 1956) 
and  Pickens  (1932)  obtained  15  to  20  individuals  from  R.  hesperus 


84 


Psyche 


[Vol.  88 


dealate  pairs  after  one  year,  and  Beard  (1974)  obtained  ca.  53 
individuals  from  R.  flavipes  dealate  pairs  after  one  year. 

King  and  Spink  (1974)  and  Akhtar  (1978)  working  with  Copto- 
termes  formosanus  Shiraki  and  C.  heimi  (Wasmann),  respectively, 
both  obtained  ca.  30  individuals  from  dealate  pairs  the  first  year. 
Since  each  of  these  workers  utilized  different  temperature  and 
rearing  methods,  and  still  obtained  similar  results,  the  observed 
growth  rates  are  probably  a fair  approximation  of  that  to  be 
expected  from  field  colonies.  Such  a growth  rate  implies  that  R. 
flavipes  and  R.  virginicus  dealate  pairs  (and  probably  other 
rhinotermitids  as  well)  are  K strategists  (Matthew,  1976).  Incipient 
colonies  will  be  successful  only  if  the  dealate  pairs  establish  nests  in 
sites  that  are  sparsely  occupied  by  other  members  of  the  same 
species,  and  which  possess  adequate  food  and  defense  requirements 
necessary  for  slow,  long  term  colony  growth  (Oster  and  Wilson, 
1978). 

Our  laboratory  data  suggest  that  at  least  for  R.  virginicus,  it 
might  be  possible  for  female  dealates  alone  to  parthenogenetically 
establish  a colony  with  a reproductive  potential  equal  to  that  of  the 
normal  male  + female  dealate  combination.6  This  finding  raises  the 
question  of  whether  all  progeny  resulting  from  the  R.  virginicus 
male  + R.  virginicus  female  combinations  are  sexual  offspring,  or 
whether  some  fraction  might  have  been  of  parthenogenetic  origin. 

We  have  also  found  that  R.  virginicus  males  readily  mate  with  R. 
flavipes  females  in  the  laboratory,  producing  apparently  viable 
progeny  at  rates  comparable  to  those  from  same-species  pairings.  In 
contrast,  the  pairing  of  R.  flavipes  males  with  R.  virginicus  females 
results  in  nuptial  cell  construction,  but  only  a very  low  rate  of 
progeny  production.  We  infer  from  our  data  that  the  progeny  of  the 
R.  virginicus  male  + R.  flavipes  female  combinations  are  true 
interspecific  hybrids  rather  than  parthenogenetic  progeny,  since 
paired  R.  flavipes  females  laid  many  eggs,  but  only  one  of  them  ever 
hatched.  Since  the  larvae  resulting  from  these  mixed-species  mat- 

6None  of  the  females  were  dissected  to  verify  the  absence  of  sperm.  All  alates  however 
were  taken  from  the  logs  before  their  normal  flight  period,  and  had  fully  developed 
wings  which  presumably  rendered  them  incapable  of  copulation  within  the  confines 
of  the  galleries  of  the  logs.  Furthermore,  no  instances  are  known  of  any  termite 
species  that  copulate  until  they  have  flown,  shed  their  wings,  and  constructed  a 
nuptial  cell.  We  consider  it  extremely  unlikely  that  the  females  used  in  our 
experiments  had  been  inseminated. 


1981]  Howard,  Mallette,  Haverty,  & Smythe — Reticulitermes  85 


ings  contained  protozoa  typical  of  both  parents,  we  also  infer  that 
the  larvae  engage  in  proctodeal  feeding  with  both  parents.  We  do 
not  know  whether  such  mixed-species  pairing  occurs  in  the  field. 
The  main  flight  periods  of  R.  flavipes  and  R.  virginicus  are 
separated  by  about  one  month  (late  February  to  early  April  for  R. 
flavipes  and  mid-April  to  mid-May  for  R.  virginicus).  But  unpub- 
lished records  from  the  Forestry  Sciences  Laboratory  in  Gulfport, 
Mississippi,  indicate  that  R.  flavipes,  at  least,  may  have  flights  every 
month  of  the  year,  rendering  it  at  least  theoretically  possible  for 
interspecific  pairing  to  occur. 

Despite  the  success  of  incipient  colonies  in  the  laboratory,  their 
intrinsically  slow  growth  rates  raise  serious  questions  regarding  the 
importance  of  such  pairs  as  a major  means  of  population  expansion. 
As  noted  in  the  introduction,  our  field  studies  (Howard  and 
Haverty,  1980)  sugget  that  R.  flavipes  frequently  undergoes  popula- 
tion expansion  by  colony  fission  with  subsequent  production  of 
multiple  neotenic  reproductives.  Since  such  new  colonies  pre- 
sumably consist  of  several  thousand  individuals,  their  ability  to 
survive  should  be  markedly  greater  than  that  of  dealate  headed 
incipient  colonies.  It  is,  of  course,  possible  that  dealate  individuals 
or  pairs  could  be  adopted  by  an  established  colony,  but  we  know  of 
no  data  to  support  such  a position. 

Clearly,  considerably  more  work  should  be  done  to  verify  the 
findings  of  these  studies.  The  success  of  intraspecific  matings  is  not 
in  question.  The  confounding  results  of  the  interspecific  matings 
demand  further  cytological  and  experimental  evaluation.  The 
mechanisms  of  reproductive  isolation  should  be  clarified  as  well  as 
the  integrity  of  these  two  sympatric  species. 

Summary 

Incipient  colony  foundation  in  the  laboratory  by  dealates  of 
Reticulitermes  flavipes  (Kollar)  and  R.  virginicus  (Banks)  was  used 
to  examine  several  possible  reproductive  strategies  available  to  these 
sympatric  subterranean  termite  species.  Successful  colony  forma- 
tion and  progeny  production  occurred  with  pairings  of  R.  flavipes 
males  + R.  flavipes  females  (from  either  spring  or  fall  flights),  R. 
virginicus  males  + R.  virginicus  females,  R.  virginicus  males  + R. 
flavipes  females,  and  R.  virginicus  females  + R.  virginicus  females. 
Few  progeny  resulted  from  pairing  R.  flavipes  males  + R.  virginicus 


86 


Psyche 


[Vol.  88 


females,  or  from  pairing  R.  flavipes  females  + R.  flavipes  females. 
All  colony  growth  rates  were  slow,  producing  no  more  than  20  to  30 
individuals  within  the  first  year. 


Literature  Cited 


Akhtar,  M.  S. 

1978.  Some  observations  on  swarming  and  development  of  incipient  colonies 
of  termites  of  Pakistan.  Pakistan  J.  Zool.  10(2):  283-290. 

Beard,  R.  L. 

1974.  Termite  biology  and  bait-block  method  of  control.  Conn.  Agric.  Exp. 
Stn.  Bull.  748.  19  p. 

Blum,  M.  S.,  and  N.  A.  Blum. 

1979.  Sexual  Selection  and  Reproductive  Competition  in  Insects.  Academic 
Press,  New  York.  463  p. 

Buchli,  H. 

1950.  Recherche  sur  la  fondation  et  le  developpement  des  nouvelles  colonies 
chez  le  termite  Lucifuge  ( Reticulitermes  lucifugus  Ressi  [sic]).  Physiol. 
Comp.  Oecologia  2:  145-160. 

Clement,  J.  L. 

1979.  Hybridation  experimentale  entre  Reticulitermes  santonensis  Feytaud  et 
Reticulitermes  lucifugus  Rossi.  Ann.  Sci.  Nat.,  Zool.  1:  251-260. 

Crozier,  R.  H. 

1979.  Genetics  of  sociality.  P.  223-286.  In  Social  Insects,  Vol.  1 . H.  R.  Herman 
(ed.).  Academic  Press,  New  York,  San  Francisco,  and  London.  437  p. 

Howard,  R.  W. 

1980.  Effects  of  methoprene  on  colony  foundation  by  alates  of  Reticulitermes 
flavipes  (Kollar).  J.  Ga.  Entomol.  Soc.  15(3):  281-285. 

Howard,  R.  W.,  and  M.  I Haverty. 

1980.  Reproductives  in  mature  colonies  of  Reticulitermes  flavipes:  abundance, 
sex-ratio,  and  association  with  soldiers.  Environ.  Entomol.  9(4):  458-460. 

King,  E.  G.,  Jr.,  and  W.  T.  Spink. 

1974.  Laboratory  studies  on  the  biology  of  the  Formosan  subterranean  termite 
with  primary  emphasis  on  young  colony  development.  Ann.  Entomol. 
Soc.  Am.  67(6):  953-958. 

Matthews,  E.  G. 

1976.  Insect  Ecology.  Univ.  of  Queensland  Press,  St.  Lucia,  Queensland.  226 
P- 

Nutting,  W.  L. 

1969.  Flight  and  colony  foundation.  P.  233-282.  In  Biology  of  Termites,  Vol. 
1.  K.  Krishna  and  F.  M.  Weesner  (eds.).  Academic  Press,  New  York  and 
London.  598  p. 

Oster,  G.  F.,  and  E.  O.  Wilson. 

1978.  Caste  and  Ecology  in  the  Social  Insects.  Princeton  Univ.  Press, 
Princeton,  N.  J.  352  p. 


1981]  Howard,  Mallet te,  Haverty,  & Smythe — Reticulitermes  87 


Pickens,  A.  L. 

1932.  Distribution  and  life  histories  of  the  species  of  Reticulitermes  Holmgren 
in  California;  a study  of  the  subterranean  termites  with  reference  to(l) 
Zoogeography,  and  (2)  life  histories.  Ph.  D.  Thesis.  Univ.  Calif. 

Weesner,  F.  M. 

1956.  The  biology  of  colony  foundation  in  Reticulitermes  Hesperus  Banks. 
Univ.  Calif.  Pub.  Zool.  61(5):  253-314. 

Yamin,  M.  A. 

1979.  Flagellates  of  the  orders  Trichomonadida  Kirby,  Oxymonadida  Grasse, 
and  Hypermastigida  Grassi  and  Foa  reported  from  lower  termites 
(Isoptera  families  Mastotermitidae,  Kalotermitidae,  Hodotermitidae, 
Termopsidae,  Rhinotermitidae,  and  Serritermitidae)  and  from  the 
wood-feeding  roach  Cryptocercus  (Dictyoptera:  Cryptocercidae).  Socio- 
biology 4(1):  1-119. 


ECOLOGY  AND  LIFE  HISTORY  OF  THE 
RHYTIDOPONERA  IMPRESS  A GROUP 
(HYMENOPTERA:FORMICIDAE) 

I.  HABITATS,  NEST  SITES,  AND  FORAGING  BEHAVIOR 

By  Philip  S.  Ward1 

Department  of  Zoology,  University  of  Sydney, 

N.S.W.  2006,  Australia 

Introduction 

The  ponerine  ants  of  the  genus  Rhytidoponera  constitute  a rich 
assemblage  of  species,  widespread  throughout  Australia,  with  lesser 
representation  in  Melanesia  and  adjacent  regions  (Brown,  1958; 
Wilson,  1958).  On  the  Australian  mainland  they  have  collectively 
occupied  a broad  range  of  habitats,  and  often  rank  among  the  more 
abundant  members  of  an  ant  community.  Considerable  interest 
centers  on  the  unusual  habit,  apparently  widespread  in  the  genus,  of 
reproduction  by  mated  “workers”  in  lieu  of  a morphologically 
differentiated  dealate  queen  (Brown,  1953,  1954;  Whelden,  1957, 
1960;  Haskins  & Whelden,  1965). 

The  Rhytidoponera  impressa  group  consists  of  a small,  distinctive 
cluster  of  species  occurring  in  mesic  habitats  (mostly  rainforest  and 
wet  sclerophyll)  along  the  east  coast  of  Australia  and  in  New 
Guinea.  Until  recently,  the  impressa  group  was  thought  to  comprise 
no  more  than  three  species,  all  reproducing  by  means  of  distinct 
winged  queens  (Brown,  1953,  1954;  Haskins  & Whelden,  1965). 
However,  recent  studies  of  systematic  relationships  and  colony 
structure  in  the  impressa  group  have  revealed  the  presence  of  at  least 
5 close^  related  species  and  the  occurrence  of  reproduction  by  both 
queens  and  mated  workers  (Ward,  1978,  1980). 

There  is  a notable  paucity  of  detailed  ecological  studies  on 
rainforest  ponerines  in  general,  and  there  have  been  no  extensive 
field  studies  on  Rhytidoponera.  This  paper  summarizes  information 
on  habitat  and  nest  site  preferences,  colony  densities,  and  various 
aspects  of  foraging,  in  the  impressa  group.  A second  paper  describes 
life  cycle  and  reproductive  patterns  (Ward,  1981). 


'Present  address:  Department  of  Entomology,  University  of  California,  Davis, 
California  95616 

Manuscript  received  by  the  editor  April  15,  1981. 


89 


90 


Psyche 


[Vol.  88 


Methods 

Data  were  gathered  during  a survey  of  the  Rhytidoponera 
impressa  group  from  approximately  100  mesic  forest  sites  in  eastern 
Australia  and  New  Guinea.  A detailed  tabulation  of  these  collection 
sites  is  given  in  Ward  (1978).  Field  work  was  carried  out  from 
October,  1974  to  October,  1978,  with  a few  additional  collections  in 
May-July,  1980.  Voucher  specimens  from  these  collections  have 
been  deposited  in  the  Australian  National  Insect  Collection  (ANIC), 
CSIRO,  Canberra. 

In  rainforest  and  wet  sclerophyll  forest  the  collection  procedure 
was  as  follows:  colonies  of  the  impressa  group  were  sought  by 
examining  all  rotting  logs,  loose  stones  and  other  potential  nest  sites 
which  were  encountered  during  a more  or  less  random  (i.e. 
undirected)  walk  through  a tract  of  suitable  forest.  In  most  localities 
a tally  was  kept  of  the  number  of  “potential  nest  sites”  (logs  and 
stones)  sampled.  The  “rotting  log”  count  was  confined  to  moist 
rotten  logs  in  middle  to  late  stages  of  decay,  with  numerous 
preformed  cavities  (corresponding  roughly  to  the  “zorapteran”  and 
“passalid”  stages  of  Wilson,  1959),  since  field  observations  showed 
that  recently  fallen  or  dessicated  logs  were  rarely  inhabited.  If  a 
single  large  log  was  dissected  in  two  places  more  than  1 meter  apart 
it  was  counted  as  two  potential  nest  sites.  Records  from  rotting  logs 
include  a few  instances  where  ants  also  nested  in  soil  below  the  log. 
Stones  ranging  in  areal  size  from  about  100  to  1500  cm2  were 
recorded  as  potential  nest  sites  if  they  rested  completely  on  the 
ground  and  could  be  easily  overturned.  Fallen  epiphytic  fern  masses 
on  the  rainforest  floor  were  also  considered  potential  nest  sites  and 
were  examined  and  counted  in  areas  where  they  occurred.  Almost 
invariably,  a single  colony  occupied  only  one  nest  site,  so  the  terms 
“colony”  and  “nest”  are  used  in  equivalently  in  this  paper. 

When  an  impressa  group  colony  was  located,  an  attempt  was 
usually  made  to  collect  the  entire  colony  contents,  i.e.  all  workers, 
reproductives,  and  brood.  This  entailed  considerable  excavation  of 
rotting  wood  and/or  soil.  Where  only  colony  fragments  were 
believed  to  be  collected,  this  was  noted. 

Collected  colonies  were  returned  to  the  lab  and  their  contents 
enumerated.  A few  were  maintained  in  modified  Janet  or  Lubbock 
nests.  The  majority  were  frozen  for  electrophoresis. 

Field  observations  of  foraging  behavior,  colony  movement,  alate 


1981] 


Ward — Rhytidoponera  impressa.  I 


91 


dispersal,  and  mating  behavior  were  also  made.  In  addition,  field 
observations  and  collections  of  related  Rhytidoponera  species  from 
Australia,  New  Guinea,  and  New  Caledonia  provided  some  com- 
parative data. 


Results 


Habitat  Preferences 

The  known  members  of  the  Rhytidoponera  impressa  group  and 
their  respective  distributions  are  as  follows  (Ward,  1980):  chalybaea 
Emery  (=  cyrus  Forel),  New  South  Wales,  southern  Queensland, 
New  Zealand  (introduced);  confusa  Ward,  Victoria,  New  South 
Wales,  southern  Queensland;  enigmatica  Ward,  New  South  Wales; 
impressa  Mayr,  Queensland;  and  purpurea  Emery  (=  splendida 
Forel),  northern  Queensland,  New  Guinea. 

Most  species  in  the  impressa  group  occupy  a considerable  range 
of  latitude,  altitude  and  forest  types;  and  all  species  show  partial 
sympatry  with  at  least  one  other  species  (Table  1).  In  this  context,  a 
sympatric  association  is  defined  as  the  occurrence  of  two  (or  more) 
species  within  the  dispersal  range  of  their  alates.  In  all  cases  of 
sympatry,  non-conspecific  nests  were  located  within  several  hun- 
dred meters  of  one  another,  and  in  most  instances  within  50  meters. 
Despite  the  overlap  between  species,  differences  in  habitat  prefer- 
ences are  apparent. 

R.  confusa  is  essentially  a species  of  wet  sclerophyll  forest  and 
temperate  rainforest.  In  Victoria  and  southern  New  South  Wales  it 
is  principally  confined  to  lowland  wet  sclerophyll,  and  does  not 
occupy  cool  temperate  rainforest  of  the  type  dominated  by  such 
trees  as  Nothofagus,  Quintinia,  and/or  Atherosperma.  At  the 
northern  limit  of  its  range,  confusa  is  restricted  to  temperate  and 
subtropical  rainforest  at  moderate  to  high  elevations.  Thus,  there  is 
an  inverse  relationship  between  elevation  and  latitude  (Figure  1), 
and  the  regression  of  altitude  on  latitude  indicates  an  average  shift 
of  about  70m  per  degree  latitude. 

In  contrast  to  confusa,  chalybaea  is  common  in  subtropical 
rainforest  of  northern  New  South  Wales  and  southern  Queensland 
(where  confusa  is  rare  or  absent).  At  the  southern  limit  of  its 
distribution,  chalybaea  is  confined  to  disturbed  lowland  habitats. 
Thus,  in  the  Sydney  region,  it  occurs  commonly  in  well-watered 
parks  and  gardens,  and  only  penetrates  wet  sclerophyll  and 


92 


Psyche 


[Vol.  88 


C O T3 


£.2 


c > 

<L>  • = 

C/3  2>  3 

<D  O' 

b ^ 0 

til  D cn 

2 o r 

* o V 
*"2  D.  ^ ' — 

cuo  c/3  <L>  oo 

.2  <-  ‘ — ■ 

I-  4i 

3 CS  c 

^ o "3 


o £ 


u H „ 

D.  ° O 

e *->  'g. 

3-0 


•2  <» 
3 T3 


! I* 

c3  'S  - 

O >>  "*t 
a-  — -a 
a 8 e 


o i ^ w 

M 2 (N  ^ 

^ X)  C3 

^ c/3  '3  «-h 


S,  “ H 


cd 


5 ■= 


I » 
§ ° 

II 


* 2 
M r-~  3 

3 ON  J-; 

° 3 & 

<3  0 


3 o -3 
-3  <U  •§ 

g 2*  * 

o ^ - 


" ° r- 

n On 

£ 2 s c 

“ S 2 « 
>>  ^ 


!■§ 


<u 


O <U  ^ o 

° -C  3 ° 

— *7  E "S 

*i  O ■"  u 

3 +-*  X i- 

•3  i_  O O 
-o  OJ  <t3 

3 ^ D-’S 

•SS^ 
Jl  §.§  .. 

3 ~ p.  <*).*■* 

g c - « — 

«-  <D 


cd  o 
D.  o 


3 ^ - Z 

•-  .2  E jd 


1)  o 
b oo 
3 u 


-g  5b  _ 
h£  o 


$5 

%l 


pu^i^j^d 
imqjnqns 
pus  imqjft 

JS9J0JUIBJ 

Ajq 


JS3JOJUIBJ 

iBJo«n 


1S9J0JUIBJ 

psoidojx 


JS9JOJUIBJ 

jBDidojjqns 


JS9JOJUIBJ 

9JBJ9dui9X 


JS9J0J 

pAqdojgps 

PM 


9SUB>1 

lEuipmiqv 


ggiie-y 

jeuipmimi 


pgipnjs 

suoqBindoj 

•OM 


I ^ 


I I 


E E B S £ 

o o o o o 

o o oo  >n  <n 

° O — o ■ ■ 

in  in  i i o 

— cn 
m 


C/3  C/2  C/3  c/3  C/3 


On  On 
n m 


a 

>3  3 

5 -Cl 

^ -b 
sr  a 
o c: 


■~  a 

S ^ 

.£p  ci,  £ 

5 S ~ 


115(25)  7°-38°S  5- 1250m  24(11)  35(6)  30(3)  10(2) 


1981]  Ward — Rhytidoponera  impressa.  I 93 


DEGREES  LATITUDE  (South) 


Figure  1 . Altitude  and  latitude  of  57  populations  of  confusa  (open  circles)  and  34 
populations  of  chalybaea  (closed  circles).  Regressions  of  altitude  on  latitude  for 
confusa  (upper  line)  and  chalybaea  (lower  line)  are  highly  significant  (p  < .001). 

rainforest  gullies  which  are  ecologically  very  disturbed,  i.e.  heavily 
encroached  with  introduced  weeds  such  as  Lantana,  Ligustrum  and 
Tradescantia. 

Sympatric  associations  between  chalybaea  and  its  sibling  species, 
confusa,  occur  in  some  of  these  disturbed  gully  sites,  with  confusa 
preferentially  occupying  the  vegetationally  less  disturbed  portions 
of  the  gully.  These  two  species  also  occur  sympatrically  in  stands  of 
undisturbed  temperate  and  subtropical  rainforest  in  northern  New 
South  Wales  and  southern  Queensland.  In  this  region  chalybaea 
tends  to  occupy  more  xeric  microhabitats  than  confusa,  but  in  one 
locality  (an  isolated  patch  of  rainforest  at  Boonoo  Boonoo  Falls, 
N.S.W.)  no  obvious  nest  site  or  microhabitat  differences  were  found 
between  the  two  species,  which  nested  within  a few  meters  of  one 
another. 

R.  chalybaea  also  shows  an  altitudinal  shift  with  increasing 
latitude  (Figure  1)  and  tends  to  occur  at  lower  elevations  than 
confusa.  The  general  picture  is  one  of  partial  ecological  differentia- 
tion between  these  two  species  despite  their  very  close  morpho- 
logical resemblance  (cf.  Ward,  1980). 


94 


Psyche 


[Vol.  88 


Table  2.  Nest  site  records  for  the  Rhytidoponera  impressa  group,  excluding  small, 
incipient  colonies  (<  20  workers).  Figures  in  parentheses  represent  the  percentages 
(for  each  species)  of  colonies  occupying  a given  type  of  nest  site. 


Species 

Rotten 

Logs 

Stones 

Fallen 

Epiphytes 

Total 

confusa 

258 

143 

11 

412 

(62.6) 

(34.7) 

(2.7) 

chalybaea 

145 

19 

1 

165 

(87.9) 

(11.5) 

(0.6) 

impressa 

13 

1 

0 

14 

(92.9) 

(7.1) 

(0.0) 

purpurea 

34 

0 

0 

34 

(100.0) 

(0.0) 

(0.0) 

enigmatica 

0 

21 

0 

21 

(0.0) 

(100.0) 

(0.0) 

all  species 

450 

(69.7) 

184 

(28.5) 

12 

(1.9) 

646 

R.  enigmatica  is  a localized  species,  known  only  from  wet 
sclerophyll  vegetation  in  sandstone  gullies  (6  sites,  including  two 
ANIC  records)  and  urban  parkland  (1  site),  the  latter  record  coming 
from  an  area  where  the  original  habitat  would  have  been  sandstone 
gully  vegetation.  The  range  of  elevation  from  which  it  has  been 
recorded  is  10  to  180  meters.  Thus,  with  regard  to  habitat  preference 
enigmatica  is  the  most  stenotopic  species.  Most  of  the  known 
populations  are  in  sympatry  with,  or  in  close  proximity  to, 
populations  of  confusa  and/or  chalybaea. 

The  7 impressa  populations  studied  come  from  tropical  rainforest 
(1),  subtropical  rainforest  (5),  and  dry  rainforest  (1).  These  data, 
along  with  30  other  collection  records  in  the  ANIC,  indicate  that 
impressa  is  confined  to  Queensland  rainforest  at  altitudes  ranging 
from  30m  to  1050m. 

Based  on  the  12  populations  studied  here  plus  additional  records 
from  the  ANIC  and  from  Wilson  (1958),  purpurea  is  recorded  from 
subtropical  and  tropical  rainforest  (and  one  population  from  dry 
microphyll  rainforest  on  the  Mt.  Windsor  Tableland)  in  northern 
Queensland  (30m  to  1200m),  and  from  tropical  montane  rainforest 
(600m  to  1300m)  in  Papua  New  Guinea.  In  north  Queensland  it 
occurs  in  both  primary-growth  and  partially  disturbed  rainforest, 


1981] 


Ward — Rhytidoponera  impressa.  I 


95 


while  New  Guinea  records  indicate  a predilection  for  second-growth 
montane  rainforest. 

Nest  Site  Preferences  and  Densities 

Members  of  the  impressa  group  are  found  nesting  mostly  in 
rotten  logs  and  under  stones.  Nests  are  multi-chambered,  but  not 
highly  fragmented,  seldom  penetrating  deeper  than  15- 20cm  into 
soil,  or  occupying  more  than  lm  length  of  rotting  log.  Nest 
entrances  are  cryptic,  without  conspicuous  mounds  of  excavated 
material. 

Fallen  epiphytes  on  the  rainforest  floor  are  occasionally  utilized 
as  nest  sites  by  confusa  and  chalybaea.  Duringthe  present  study  no 
colonies  were  found  in  living  epiphytes  on  trees,  although  there  are 
single  records  of  a colony-founding  purpurea  queen  (Brown,  1954) 
and  a mature  purpurea  colony  (Wilson,  1958)  from  fern  epiphytes 
on  rainforest  trees. 

Nest  site  records  from  the  present  study  are  summarized  in  Table 
2 which  lists,  for  each  species,  the  number  of  colonies  collected  from 
rotten  logs,  under  stones,  and  in  fallen  epiphytes.  Excluded  from 
this  table  are  a small  number  of  single  records  from  other  nest  sites. 
Thus  confusa  was  also  found  nesting  in  a Banksia  lignotuber,  in  a 
rotting  bracket  fungus,  directly  in  the  soil,  and  (twice)  in  an 
abandoned  termite  mound  in  rainforest.  A chalybaea  colony  was 
located  under  the  bark  sheath  of  an  Archontophoenix  palm,  and  in 
urban  areas  this  species  occupied  less  orthodox  nest  sites  (e.g.  in  and 
under  rusting  metal,  under  concrete  slabs,  and  in  crevices  along  a 
stone  wall).  Three  purpurea  colonies  (two  in  north  Queensland,  one 
in  Papua  New  Guinea)  were  observed  nesting  in  cavities  in  the 
trunks  of  living  rainforest  trees,  and  in  New  Guinea  this  species  may 
be  primarily  an  arboreal  nester  (Wilson,  1958;  records  in  ANIC). 

Table  2 shows  that  there  is  a clear  trend  towards  greater 
specialization  in  the  rotten  log  nest  site  in  species  of  more  tropical 
latitudes.  The  difference  between  confusa  and  chalybaea  with 
respect  to  numbers  of  logs  and  stones  utilized  is  highly  significant 
(x?  = 33.0,  p < .001)  and  the  difference  between  chalybaea  and 
purpurea  is  also  significant  (x?  = 4.4,  p < .05).  In  contrast  to  all 
others,  enigmatica  (the  localized  species  of  wet  sclerophyll  gullies) 
appears  to  nest  exclusively  under  stones. 

In  70  populations  (from  63  localities,  due  to  some  sympatry)  a 
tally  was  kept  of  the  number  of  “potential”  nest  sites  (rotten  logs, 


96 


Psyche 


[Vol.  88 


00  I .0 

LlJ 


OO 

h- 

00 

LlJ 


.8 


Q 

m -b  h 


>- 

o 

z 

UJ 

=> 

o 

UJ 

CL 


A - 


.2- 


.2  .4  .6  .8  1.0 

FREQUENCY  OF  POTENTIAL  NEST  SITES 


Figure  2.  Within-species  frequencies  of  utilized  nest  sites  as  a function  of  potential 
nest  site  frequencies,  for  5 impressa  group  species.  Closed  circles  refer  to  log  nest  sites, 
open  circles  to  stones. 

stones,  fallen  epiphytes)  encountered  as  well  as  the  number  of  actual 
nest  site  occupancies  (Table  3).  It  seems  clear  that  nest  site 
availability  varies  from  species  to  species.  For  both  rotten  log  and 
stone  nest  sites  there  are  positive  correlations  (r  = 0.94,  p < .02,  in 
both  instances,  arcsine  transformed  data)  between  the  proportion  of 
a species’  colonies  found  in  a particular  nest  site  and  the  relative 
frequency  of  that  nest  site  for  the  species  (Figure  2).  This  suggests 
that  species-specific  preferences  are  partly  a function  of  nest  site 
availability.  (No  such  correlation  is  found  for  fallen  epiphytes — 
confusa  showns  the  highest  preference  for  this  nest  site  despite  its 
relative  rarity  in  the  southern  rainforests;  however,  the  numbers  are 
in  all  instances  rather  low.) 

The  relative  abundances  of  species  can  be  crudely  compared  by 


1981] 


Ward — Rhytidoponera  impressa.  I 


97 


Table  3.  Numbers  of  potential  nest  sites  (pns)  sampled  and  actual  nests  encoun- 
tered, for  70  impressa  group  populations. 


Species 

No. 

populations 

Logs 

Stones  Epiphytes 

Total 

confusa 

37 

no.  pns 

1838 

2984 

92 

4914 

no.  nests 

227 

98 

8 

333 

nests/ pns 

.124 

.033 

.087 

.068 

chalybaea 

22 

no.  pns 

1 164 

1136 

70 

2370 

no.  nests 

141 

17 

1 

159 

nests/ pns 

.121 

.015 

.014 

.067 

impressa 

4 

no.  pns 

260 

126 

7 

393 

no.  nests 

8 

1 

0 

9 

nests/  pns 

.031 

.008 

.000 

.023 

purpurea 

5 

no.  pns 

404 

109 

24 

537 

no.  nests 

21 

0 

0 

21 

nests/  pns 

.052 

.000 

.000 

.039 

enigmatica 

2 

no.  pns 

105 

561 

666 

no.  nests 

0 

15 

15 

nests/ pns 

.000 

.027 

.027 

all  species 

70 

no.  pns 

3771 

4916 

193 

8880 

no.  nests 

397 

131 

9 

537 

nests/ pns 

.105 

.270 

.047 

.060 

examining  the  proportion  of  potential  nest  sites  which  are  occupied. 
(The  desirable  complementary  data  on  absolute  densities  of  poten- 
tial nest  sites  for  different  geographical  regions  and  habitats  are  not 
available).  Comparing  the  density  figures  (Table  3)  for  confusa  and 
chalybaea,  the  former  occupies  a significantly  greater  proportion  of 
stone  nest  sites  than  chalybaea  (x?  = 9.7,  p < .01),  but  no  differences 
exist  in  the  proportion  of  suitable  rotten  logs  occupied,  and  the 
overall  nest  densities  (considering  all  potential  nest  sites)  are  the 
same  for  the  two  species.  Nest  densities  are  considerably  lower  for 
impressa,  purpurea,  and  enigmatica.  Rhytidoponera  confusa  and 
chalybaea  utilize  a significantly  greater  proportion  of  rotten  logs 
than  impressa  and  purpurea  (contingency  x\  p <.001,  for  all  four 
comparisons),  despite  the  greater  importance  of  rotting  logs  as  nest 
sites  in  the  more  northerly  (tropical)  species.  This  may  be  partly  the 
result  of  greater  competition  for  nest  sites  in  the  species-rich  tropical 
rainforests.  R.  confusa  and  chalybaea  are  often  common  and 
dominant  ants  in  temperate  and  subtropical  rainforests,  respec- 
tively, of  New  South  Wales  and  southern  Queensland  where  the 


98 


Psyche 


[Vol.  88 


numbers  of  sympatric  rainforest  ant  species  are  probably  about  one- 
quarter  to  one-half  that  experienced  by  purpurea  in  north  Queens- 
land rainforest. 

It  is  unclear  why  there  is  a disproportionate  decline  in  the 
utilization  of  stones  as  nest  sites  in  the  more  tropical  members  of  the 
impressa  group  (Table  3)  and  perhaps  for  tropical  rainforest  ants  in 
general  (cf.  Wilson,  1959,  p.  440).  One  possibility  is  that  in 
subtropical  and  tropical  rainforests  on  well-drained  soils,  stones 
frequently  lie  on  subsoil  below  the  thin  organic  horizon  and  offer  an 
environment  poorer  in  immediate  food  resources  and  more  de- 
manding for  nest  excavation  than  rotting  logs.  In  temperate  and 
some  subtropical  rainforests  of  New  South  Wales,  soil  horizons 
tend  to  be  less  sharply  stratified  and/or  litter  decomposition  is 
slower,  so  that  humic  material  extends  below  the  level  of  loose 
stones. 

Effects  of  Sympatry 

Nest  site  densities  for  sympatric  and  allopatric  populations  of 
confusa  and  chalybaea  are  given  in  Table  4.  Both  species  occupy  a 
significantly  greater  proportion  of  log  nest  sites  in  allopatric 
populations  (contingency  x2,  p < .01  and  p < .001,  for  confusa  and 
chalybaea  respectively)  and  confusa  inhabits  a greater  proportion  of 
stone  nests  sites  allopatrically  (x2  = 5.4,  p < .05).  The  lower 
sympatric  densities  of  confusa  and  chalybaea  could  be  a result  of 
sympatric  associations  occurring  in  more  marginal  environments. 
However,  the  combined  sympatric  nest  densities  are  very  similar  to 
the  allopatric  densities  of  both  species.  There  are  no  significant 
differences  between  the  total  proportion  of  rotting  logs  occupied 
sympatrically  and  the  proportion  utilized  allopatrically  by  either 
confusa  (x2  = 0.7)  or  chalybaea  (x2  = 1.8).  The  combined  sympatric 
nest  density  under  stones  is  the  same  as  that  for  allopatric  confusa 
populations.  While  these  results  could  be  coincindental,  it  seems 
more  reasonable  to  conclude  that  sympatry  has  a depressant  effect 
on  relative  abundance,  and  that  competition  for  nest  sites,  food,  or 
foraging  space  is  important. 

Other  Sympatric  Congeners 

Other,  more  distantly  related  Rhytidoponera  species  also  co- 
occur with  members  of  the  impressa  group.  R.  victoriae  (s.l.)  is  a 
common  species  (or  complex  of  species)  present  in  rainforest  and 
other  mesic  habitats  along  the  entire  east  coast  of  Australia.  R. 


1981] 


Ward — Rhytidoponera  impressa.  I 


99 


03 

a, 

_o 

~0 

c 

03 


?3 

O (3 
<1 

<u  «£> 

a 

£ 

3 ^ 

Z.  ’2 


- 

XI  > 


I 

b s. 

oo  r-  cn  Tt 

o >r 

& * 
o 3 

£ ° 

© © © © 

2 

c 

o 

£5 

No.  potential 
nest  sites 

650 

2334 

650 

486 

.0  ^3 

Sc 

o 

Proport 

occupit 

.071 

.133 

.043 

.146 

c 

3 

o 

oc, 

No.  potential 
nest  sites 

280 

1558 

280 

884 

*< 

st 

o 

£ 

< 3 

1 

5 

32 

5 

17 

53  Js 

t > 

1 8 

3 -G 

5 1 

2 

ic  wi 
c) 

itric 

trie) 

3 

3 

£ 

o3  ~ & a, 

si  * = 

« o y\  os 

■Ss  ,» 

a <3  33 

5 S5  <)  <> 

e 3 3 
O O -3  C 

3 3 3 3 

100 


Psyche 


[Vol.  88 


victoriae  is  considerably  smaller  than  the  impressa  group  species, 
and  nests  preferentially  under  stones. 

In  some  north  Queensland  localities,  purpurea  or  impressa 
coexist  with  one  of  several  small  Rhytidoponera  species  (e.g. 
chnoopyx  and  kurandensis  nesting  in  logs  and  under  stones)  and 
with  one  of  several  larger  species  (scaberimma  and  related  species, 
nesting  in  logs  and  directly  in  the  soil).  There  are  no  rainforest 
Rhytidoponera  of  comparable  size  to  the  impressa  group  species 
that  regularly  coexist  with  the  latter  with  the  exception  of  croesus 
(s.l.),  which  nests  in  rotten  logs  and  in  tree  trunks  in  rainforest  and 
wet  sclerophyll  of  New  South  Wales  and  southern  Queensland.  R. 
croesus  appears  to  be  generally  uncommon,  and  in  fact  averages 
slightly  smaller  than  chalybaea,  confusa  and  impressa  to  a degree 
which  may  significantly  reduce  prey  size  overlap  (see  below). 

Colonies  of  other  Rhytidoponera  species  are  virtually  never 
found  occupying  the  same  nest  site  as  an  impressa  group  colony 
even  though  other  medium  to  large  ponerines  such  as  Amblyopone 
australis,  Leptogenys  hackeri  and  Prionogenys  podenzanai  are 
occasionally  found  nesting  in  close  proximity  to  an  impressa  group 
colony  (e.g.  under  the  same  stone,  or  in  adjacent  cavities  in  a log). 

Colony  Movement 

It  appears  that  species  in  the  impressa  group  are  prone  to  move 
colonies  from  one  nest  site  to  another  rather  frequently.  For 
example,  in  one  rainforest  population  of  confusa  (Royal  National 
Park,  N.S.W.)  eight  stones  under  which  colonies  had  been  briefly 
located  and  otherwise  left  undisturbed  were  examined  one  week 
later:  half  were  unoccupied.  Three  weeks  later,  only  two  colonies 
remained  under  the  stones.  While  the  censussing  no  doubt  consti- 
tuted a disturbance  conducive  to  nest-movement,  it  demonstrates 
nevertheless  the  readiness  with  which  colony  movement  is  carried 
out. 

During  the  course  of  field  collections,  vacated  nest  chambers  were 
occasionally  encountered  (under  stones  or  in  rotten  logs)  whose 
previous  occupants  could  be  traced  to  an  impressa  group  species  on 
the  basis  of  cocoon  remains  in  the  middens.  Moreover,  colony 
movement  involving  transport  of  brood  and  other  workers  was 
observed  several  times  in  chalybaea  (and  in  other  Rhytidoponera 
species  outside  the  impressa  group)  (Ward,  1981). 


1981] 


Ward — Rhytidoponera  impressa.  I 


101 


Foraging  and  Food- Retrieval 

Members  of  the  Rhytidoponera  impressa  group  are  partly 
predacious  on  other  arthropods,  but  also  scavenge  for  dead  insects, 
seeds,  animal  feces,  etc.  Capture  of  live  prey  is  achieved  by  a short 
lunge  forward,  coincindent  with  rapid  closure  of  the  outstretched 
mandibles.  Prey  thus  captured  are  subdued  by  stinging. 

In  most  species,  foraging  occurs  principally  on  the  ground, 
among  leaf  litter  and  rotting  logs.  However, purpurea  workers  were 
frequently  observed  foraging  on  low  foliage  of  understorey  plants, 
as  well  as  on  the  rainforest  floor,  in  north  Queensland.  In  Papua 
New  Guinea  this  species  nests  (at  least  partly)  arboreally,  but  limited 
observations  (Wau;  September,  1975)  suggests  that  it  tends  to 
forage  downward  from  the  nest  entrance.  Urban  and  suburban 
populations  of  chalybaea,  noted  for  their  unusual  nest  sites  (above), 
usually  forage  on  the  ground  and  on  low  vegetation,  in  damp  tree- 
shaded  situations.  On  one  occasion  chalybaea  workers  were  ob- 
served foraging  in  a house  in  an  urban  residential  area  of  Sydney. 

Foraging  is  not  restricted  to  any  particular  time  of  the  day  or 
season,  although  activity  decreases  noticeably  towards  the  middle  of 
the  day  (and  in  the  winter).  Periods  of  clear  warm  weather  after  rain 
seem  particularly  conducive  to  high  levels  of  foraging  activity. 

Field  observations  indicate  that  workers  are  usually  lone  foragers, 
although  occasionally  several  individuals  co-operatively  transport  a 
large  food  item  back  to  the  nest.  Sometimes  this  occurs  close  to  the 
nest  entrance,  seemingly  as  a result  of  fortuitous  encounters  of  a 
heavily-laden  forager  with  other  workers.  In  lab  colonies  of 
chalybaea,  single  workers  struggling  with  a large  prey  item  in  a food 
arena  were  observed  to  make  movements  of  the  gaster  suggesting 
stridulation.  On  the  other  hand,  chemical  recruitment  to  food 
sources  does  occur,  although  this  behavior  is  rudimentary  in 
comparison  to  the  mass-recruitment  patterns  of  some  higher  ants.  It 
is  readily  demonstrated  by  placing  large  food  baits  (e.g.  chunks  of 
tuna  fish  or  large  insects)  close  to  a nest.  Workers  which  discover 
the  food  and  return  to  the  nest  with  a portion  of  the  bait  can  be 
observed  dragging  the  tips  of  their  gasters  along  the  ground,  and 
subsequent  outward-bound  foragers  follow  the  same  path  to  the 
food  (field  observations  on  chalybaea  and  purpurea).  Large  pieces 
of  the  bait  are  retrieved  co-operatively  by  several  workers;  smaller 
portions  are  carried  by  single  foragers. 


102 


Psyche 


[Vol.  88 


When  such  baiting  experiments  are  carried  out,  there  appears  to 
be  little  active  defense  of  the  food  by  Rhytidoponera  workers.  When 
baits  are  partially  occupied  by  other  smaller  but  mass-recruiting  ant 
species,  such  as  Pheidole,  Rhytidoponera  workers  adopt  a “grab- 
and-run”  strategy.  This  is  illustrated  by  the  following  observations 
on  purpurea  in  rainforest  near  Cape  Tribulation,  north  Queensland 
(5  June  1980). 

A purpurea  colony  was  located  in  the  trunk  of  a living  palm  tree, 
in  a cavity  60cm  above  ground.  Workers  were  foraging  down  the 
palm  trunk  and  on  the  adjacent  rainforest  floor.  A small  chunk  of 
tuna  fish  was  placed  on  a stone,  1.5m  from  the  palm  tree,  and  close 
to  a purpurea  forager  which  soon  located  the  bait.  It  grasped  a small 
piece  of  the  tuna  and  returned  to  the  nest,  dragging  the  end  of  its 
gaster  along  the  ground.  A few  minutes  later,  a worker  (possibly  the 
same  individual)  emerged  from  the  nest  entrance  and  returned  to  the 
bait  by  exactly  the  same  trail.  By  this  time,  the  remaining  tuna  bait 
was  in  two  pieces,  each  attended  by  2-3  workers  of  a Meranoplus 
sp.  The  purpurea  worker  carefully  circled  around  one  piece  of  tuna 
to  an  unoccupied  corner  and  grabbed  it,  inadvertently  getting  a 
Meranoplus  worker  at  the  same  time.  The  two  briefly  grappled,  and 
the  purpurea  worker  dropped  the  food  and  retreated  several 
centimeters.  It  then  approached  the  second  piece  of  tuna,  edged  in 
towards  another  exposed  corner,  swiftly  grabbed  it  (this  time 
without  a Meranoplus  worker),  and  hurriedly  departed  for  the  nest 
by  a different  route. 

Unrecruited  workers  of  the  impressa  group  apparently  forage 
randomly,  without  laying  a continuous  odour  trail,  but  upon 
locating  food  they  return  directly  to  the  nest.  It  is  unclear  what 
method(s)  of  orientation  are  utilized.  Any  explanation  must  take 
into  account  the  observation  that  foraging  occurs  nocturnally  as 
well  as  diurnally  (at  least  in  confusa  and  chalybaea). 

Food  Diversity  and  Size 

The  great  majority  of  food  items  collected  by  impressa  group 
workers  are  small,  individual  objects  brought  in  by  single  foragers. 
Eighty-one  food  items  were  returned  to  a single  chalybaea  nest 
observed  over  a total  of  8 hours  (Table  5).  Of  these,  one  item  (an 
earthworm)  was  transported  by  four  workers;  the  remaining  food 
items  (encompassing  56  arthropods,  17  Ficus  seeds  or  pieces  of  fruit, 
and  7 pieces  of  miscellaneous  organic  material)  were  carried  by 


1981] 


Ward — Rhytidoponera  impressa.  I 


103 


T3 

i 

Q 

-C 

a 

■5 

5 

cu 

ft 

O 

1 

* 


.©  00 

O O 
o <U 

t 6 

oo  '■£ 

<— > c 

o .2 

.s  5 

j e: 

a> 

CO 

?•§ 


H 2 


2^ 
cd  u 
D.^ 

a, 

< 


OO— hOOOO  — <N  — <N  — — < <N  © m m <N  <N  — — H CN  — * I I I v© 


— I — . >/->  <N  — coo— • — « O O <N  © © © ~ O — © ro  <N  CN  © © I 


i-,  u o 

<D  <U  > 

Xj  M > 

Uh  U-  L_ 

O O o 

£ £ c 

dd  loo 


O 

£ £ « 
i-  i-  cd 
2 .2,  * 

Sij 

E 


<x 

II 

II 


(N  <N 


-Si  -Si 

I :§ 

CL,  ft, 


O, 

o 

C i 
<u 
8 
>% 

X 


<u  <u 

Cd  03 

t & 


<u 
cd 
-a 

•e 

.2  £ ~ 

5 0712  2 2 
. . . . <D  <U 


cd  cd 


3 
^ T3 
*S  cd 
3 w 

■a  2 
cd  ,2 
' — ' JO 

flj 

cd  « 

*2  8 
:c  o 

S £3 

X U .3 


§•3 
8 T3 
>>  cd 


2 e 

|1 


4> 
cd 
T3  2 
■ — cd 

^.*2 
x>  £ 
o 2 

6 I 

O g 
e ** 

W O 


a «2  _ 

0 aa 
8 ° ° 
g'2’2  « 
8 'E/o,  « 

D U 

1 J J 


a&a 

o o o 

4)  o 4> 

B'.&’o  o o 
QQUUU 


4)  <U 
0,0.. 
o 2 

6 x 

o 6 

X w . 


c«  jrt  cd 
t;  •§  T3 

S3  © © 

o.-g  2- 

w C j- 
O O D, 
<o  =3  e 
2 0 0 
£ U < 


Isopoda:  Oniscidae  . . . 

Diplopoda  

Annelida:  Lumbricidae 

Ficus  seeds 

Ficus  fruit  (pulp) 


104 


Psyche 


[Vol.  88 


Table  6.  List  of  1 9 food  items  returned  to  a single  nest  of  Rhytidoponera  croesus 
s.l.  (Royal  National  Park,  N.S.W.,  26  January,  1976)  over  a three-hour  observa- 


tion period. 

Hymenoptera:  Formicidae:  Paratrechina  sp.  (worker) 1 

" " Solenopsis  sp.  (worker) 1 

" " Chelaner  sp.  (worker) 1 

" " Myrmicinae  (male  alate) 1 

Hymenoptera:  Pteromalidae  (adult)  1 

Lepidoptera:  adult  microlepidopteran 1 

Lepidoptera:  larvae 1 

Diptera:  Nematocera  (adults) 2 

Diptera:  Brachycera  (adults) 2 

Coleoptera:  Chrysomelidae  (adult)  1 

Homoptera:  Coccoidea  (nymph) 1 

Homoptera:  Cicadellidae  (nymph) 1 

Insect  larva,  unidentified 1 

Unidentified  insect  legs  2 

Acarina  (small  mite) 1 

Mammalian  (?)  excrement,  with  veg.  matter  and  insect  parts 1 

19 


single  workers.  Thirty-one  (55%)  of  the  56  arthropod  items  were 
alive  when  retrieved  from  their  captors  (near  the  nest  entrance). 
Some  of  the  remaining  items  may  have  been  killed  or  paralyzed 
during  capture;  others  were  clearly  scavenged  as  dead  material. 

It  is  of  some  interest  to  note  that  19  (34%)  of  the  56  arthropod 
items  consisted  of  other  ant  species  (including  alates).  Some  of  these 
ants,  particularly  alates,  may  have  been  injured  or  dying  when 
collected.  On  the  other  hand,  predation  on  healthy,  active  worker 
ants  was  observed  first-hand  in  the  field:  chalybaea  workers  from 
the  Sydney  University  population  were  seen  preying  at  the  soil 
entrances  of  Pheidole  nests,  grabbing  workers  as  they  emerged. 

For  comparison  with  another  similar-sized,  rainforest  species  of 
Rhytidoponera  outside  the  impressa  group,  Table  6 lists  the  food 
items  returned  to  a Rhytidoponera  croesus  nest  over  a three-hour 
observation  period.  The  mean  head  widths  for  workers  of  croesus 
and  chalybaea  are  1.25  ± 0.03  s.d.  (n=8)  and  1.36  mm  ± 0.08  s.d. 
(n=80),  respectively.  Although  there  is  considerable  similarity  in 
food  items  taken  by  the  two  species  as  measured  by  ordinal 
taxonomic  categories,  an  analysis  of  food  size  (Figure  3)  reveals  that 
the  mean  food  item  length  of  croesus  (2.5  mm)  is  significantly  less 
than  that  of  chalybaea  (3.5  mm)  (t-test,  p<  .02).  However,  the  food 


NO.  OF  ITEMS 


1981] 


Ward — Rhytidoponera  impressa.  I 


105 


Figure  3.  Frequency  distributions  of  the  lengths  of  food  items  taken  from  80 
chalybaea  foragers  and  19  croesus  foragers  (see  Tables  5 and  6).  Each  distribution  is 
based  on  workers  from  one  colony  only. 


size  distributions  are  based  on  limited  single-nest  samples,  and  there 
is  likely  to  be  significant  temporal  and  spatial  heterogeneity  within, 
as  well  as  between,  species. 

Additional  studies  on  food  item  diversity  and  overlap  in  Rhy- 
tidoponera are  desirable.  Such  studies  are  feasible  for  ants  which  are 
primarily  lone-foraging  predators  and  scavengers,  because  of  the 
discrete,  visible  nature  of  most  foraged  items.  However,  difficulties 
remain  in  assessing  the  importance  of  honeydew  and  other  liquid 
foods,  which  may  be  carried  in  the  crop  as  well  as  between  the 
mandibles. 

Two  species  in  the  impressa  group  were  recorded  collecting 
honeydew  from  homopterans.  Workers  of  chalybaea  were  seen 
tending  coccids  on  a fresh  shoot  emerging  from  the  trunk  of  a 
camphor  laurel  tree  ( Cinnamomum  eamphora),  in  the  Sydney 
region.  R.  purpurea  workers  were  observed  tending  aphids  on 
ginger  plants  ( Alpinia  caerulaea ) in  several  places  at  Lake  Eacham, 
Qld. 


106 


Psyche 


[Vol.  88 


In  one  of  the  latter  instances,  observations  were  made  inter- 
mittently over  a period  of  2 days,  during  which  time  a force  of  about 
15  workers  was  regularly  maintained  on  the  plant.  These  workers 
gave  outward-facing  aggressive  displays  (mandibles  barred)  when 
the  plant  was  disturbed.  A small  contingency  of  workers  was  also 
clustered  among  leaf  litter  at  the  base  of  the  plant,  apparently 
controlling  access  to  the  plant  and  aphids.  Detense  of  “tending 
rights”  may  be  important  since  other  aggressive,  aphid-tending  ants 
such  as  Pheidole  were  present  in  the  same  locality.  The  colony  of  the 
purpurea  workers  was  located  in  a rotten  log  5m  distant.  Workers 
returning  to  the  colony  from  the  ginger  plant  showed  high  fidelity  to 
a particular  route  which  involved  following  the  ground  for  half  the 
distance  and  then  proceeding  along  a decumbent  liana  (one  of 
many)  which  led  back  to  the  log. 

Thus,  despite  the  “lone  forager”  status  of  most  impressa  group 
workers,  short-term  recruitment,  co-operative  food  retrieval,  and 
(in  at  least  one  species)  persistent,  long-range  trails,  may  be  used. 
Excepting  persistent  trails,  species  in  the  impressa  group  appear  to 
show  a level  of  individual  foraging  and  recruitment  similar  to  that 
described  for  the  myrmicine  ant,  Novomessor  (Holldobler,  et  al., 
1978). 

The  species  in  the  impressa  group  with  the  most  sophisticated 
foraging  and  recruitment  behavior  {purpurea ) is  the  only  member 
whose  colonies  are  entirely  monogynous  and  queenright.  It  is 
tempting  to  speculate  that  widespread  polygyny  and  worker  repro- 
duction in  other  Rhytidoponera  species  may  have  constrained 
ergonomic  improvements  because  of  a reduction  in  the  efficacy  of 
colony-level  selection  (cf.  Oster  & Wilson,  1978). 

Summary 

The  five  known  species  of  the  Rhytidoponera  impressa  group 
collectively  inhabit  a variety  of  mesic  forest  habitats  (from  wet 
sclerophyll  to  tropical  rainforest)  along  the  east  coast  of  Australia, 
with  one  species  {purpurea ) also  occurring  in  montane  rainforest  of 
New  Guinea.  R.  chalybaea  has  invaded  mesic  anthropogenic 
habitats  (parks  and  gardens)  in  the  Sydney  region.  All  species  show 
partial  sympatry  with  at  least  one  other  species. 

Most  colonies  are  located  in  rotten  logs  or  under  stones.  There 
are  significant  differences  between  species  in  the  frequencies  with 


1981] 


Ward — Rhytidoponera  impressa.  I 


107 


which  different  nest  sites  are  utilized,  and  these  preferences  are 
correlated  with  the  availability  of  potential  nest  sites.  The  more 
tropical  species  ( impressa  and  purpurea)  show  a stronger  preference 
for  rotten  logs,  but  occur  at  lower  nest  densities,  than  inhabitants  of 
temperate  and  subtropical  rainforest  (< confusa  and  chalybaea). 
Where  confusa  and  chalybaea  occur  sympatrically,  they  have 
significantly  lower  nest  densities  than  allopatrically. 

Workers  of  the  impressa  group  are  generally  lone-foraging 
predators  and  scavengers,  but  co-operative  food  retrieval  and 
recruitment  to  food  sources  occur  to  a limited  degree.  The  majority 
of  food  items  are  small  arthropods:  other  ant  species  may  be  a 
significant  component  of  the  diet.  Foraging  usually  occurs  among 
leaf  litter  and  logs  on  the  ground  but  at  least  two  species  ( chalybaea 
and  purpurea ) also  forage  on  low  foliage  and  tend  homopterans. 

References 

Brown,  W.  L. 

1953.  Characters  and  synonymies  among  the  genera  of  ants.  Part  I.  Breviora, 
11,  1-13. 

1954.  Systematic  and  other  notes  on  some  of  the  smaller  species  of  the  ant 
genus  Rhytidoponera  Mayr.  Breviora,  33,  1 11. 

1958.  Contributions  toward  a reclassification  of  the  Formicidae.  II.  Tribe 
Ectatommini  (Hymenoptera).  Bull.  Mus.  Comp.  Zool.  Harvard,  118, 
175-362. 

Haskins,  C.  P.  and  W.  M.  Whelden. 

1965.  “Queenlessness”,  worker  sibship,  and  colony  versus  population  structure 
in  the  formicid  genus  Rhytidoponera.  Psyche,  72,  87-112. 
Holldobler,  B.,  R.  C.  Stanton  and  H.  Markl. 

1978.  Recruitment  and  food-retrieving  behavior  in  Novomessor  (Formicidae, 
Hymenoptera).  I.  Chemical  signals.  Behav.  Ecol.  Sociobiol.,  4,  163-181. 
Oster,  G.  F.  and  E.  O.  Wilson 

1978.  Caste  and  ecology  in  the  social  insects.  Princeton  University  Press, 
Princeton,  N.J. 

Specht,  R.  L.,  E.  M.  Roe  and  V.  H.  Boughton 

1974.  Conservation  of  major  plant  communities  in  Australia  and  Papua  New 
Guinea.  Aust.  J.  Bot.  Suppl.  No.  7. 

Ward,  P.  S. 

1978.  Genetic  variation,  colony  structure,  and  social  behaviour  in  the  Rhy- 
tidoponera impressa  group,  a species  complex  of  ponerine  ants.  Ph.D. 
thesis.  University  of  Sydney. 

1980.  A systematic  revision  of  the  Rhytidoponera  impressa  group  (Hymenop- 
tera: Formicidae)  in  Australia  and  New  Guinea.  Aust.  .1.  Zool.  28, 
475-498. 


108 


Psyche 


[Vol.  88 


1981.  Ecology  and  life  history  of  the  Rhytidoponera  impressa  group  (Hymen- 
optera:  Formicidae).  II.  Colony  origin,  seasonal  cycles,  and  reproduc- 
tion. Psyche,  88:  109-126. 

Webb,  L.  S. 

1978.  A structural  comparison  of  New  Zealand  and  south-east  Australian  rain 
forests  and  their  tropical  affinities.  Aust.  J.  Ecol.  , 7-21. 

Whelden,  W.  M. 

1957.  Anatomy  of  Rhytidoponera  convexa.  Ann.  Ent.  Soc.  Am.,  50,  271-282. 
1960.  Anatomy  of  Rhytidoponera  metalliea.  Ann.  Ent.  Soc.  Am.,  53,  793-808. 

Wilson,  E.  O. 

1958.  Studies  on  the  ant  fauna  of  Melanesia.  III.  Rhytidoponera  in  western 
Melanesia  and  the  Moluccas.  IV.  The  tribe  Ponerini.  Bull.  Mus.  Comp. 
Zool.  Harvard,  119,  303-371. 

1959.  Some  ecological  characteristics  of  ants  in  New  Guinea  rain  forests. 
Ecology,  40,  437-447. 


ECOLOGY  AND  LIFE  HISTORY  OF  THE 
RHYTIDOPONERA  IMPRESS  A GROUP 
(HYMENOPTERA:FORMICIDAE) 

II.  COLONY  ORIGIN,  SEASONAL  CYCLES, 

AND  REPRODUCTION 

By  Philip  S.  Ward1 

Department  of  Zoology,  University  of  Sydney, 

N.S.W.  2006,  Australia 

Introduction 

This  paper  is  concerned  with  colony  foundation  and  with 
seasonal  cycles  in  brood  composition  and  alate  production  in  the 
Rhytidoponera  impressa  group,  a species  complex  of  ponerine  ants 
restricted  to  rainforest  and  other  mesic  habitats  in  eastern  Australia 
and  New  Guinea. 

Life  cycle  information  is  most  complete  for  confusa  and  chaly- 
baea,  and  most  of  what  follows  refers  to  those  species.  Relevant 
data  on  the  other  three  members  of  the  impressa  group  ( enigmatica , 
impressa,  and  purpurea ) are  given  where  available.  When  pertinent 
to  the  discussion,  some  observations  on  related  Rhytidoponera 
species  outside  the  impressa  group  are  also  included. 

Methods 

Collection  methods  are  described  in  Ward  (1981).  Most  of  the 
data  are  based  on  field  observations  and  collections.  Where  ap- 
propriate, suspected  reproductive  females  were  dissected  to  ascer- 
tain the  condition  of  the  ovaries  and  spermatheca. 

Results 

Colony  origin 

In  the  Rhytidoponera  impressa  group  there  are  two  methods  by 
which  colonies  can  originate: 

(i)  from  lone,  colony-founding  winged  females  (queens),  in  the 
manner  characteristic  of  many  ants;  or 


‘Present  address:  Department  of  Entomology,  University  of  California,  Davis, 
California  95616 

Manuscript  received  by  the  editor  April  15,  1981. 


109 


110 


Psyche 


[Vol.  88 


(ii)  as  a result  of  colony  fission  or  budding  (hesmosis),  in  which 
one  or  more  mated  “workers”,  accompanied  by  uninsemi- 
nated nest-mates,  leave  the  parent  colony  to  found  a new 
daughter  nest. 

As  the  foregoing  remarks  imply,  there  are  two  kinds  of  reproduc- 
tive females:  queens  and  ergatoid  (worker-like)  gynes,  the  latter 
indistinguishable  morphologically  from  unmated  workers.  This  is 
the  first  record  of  reproductive  workers  in  the  impressa  group  (they 
are  common  and  well-documented  in  some  other  Rhytidoponera ) 
where  previous  reports  suggested  that  the  only  functional  reproduc- 
tives  were  winged  queens  (cf.  Brown,  1953,  1954;  Haskins  & 
Whelden,  1965).  Mated  queens  and  ergatoid  gynes  never  coexist  in 
the  same  nest,  but  they  often  occur  in  different  nests  in  the  same 
population  (in  confusa,  chalybaea  and  impressa ).  This  rather 
remarkable  dimorphism  of  female  reproductives  in  the  impressa 
group  and  the  resulting  differences  in  colony  structure  and  genetic 
relatedness  will  be  examined  in  more  detail  elsewhere  (Ward,  in 
prep.). 

There  is  little  information  on  the  frequency  of  colony  fission  in 
worker-reproductive  colonies  or  on  the  size  of  newly-budded 
daughter  colonies.  Occasionally  small  isolated  clusters  of  workers 
and  brood  are  seen  in  the  field,  under  stones  or  in  rotten  log  cavities. 
Table  1 summarizes  the  composition  of  four  such  clusters  in  the 
impressa  group,  and  two  from  other  Rhytidoponera  species  ( tas - 
maniensis  and  fulgens).  Similar  observations  were  made  by  Haskins 
& Whelden  (1965)  on  R.  metallica.  Note  that  in  the  two  cases  where 
workers  were  dissected  (Table  1),  only  one  individual  in  each  cluster 
was  found  to  be  inseminated.  In  no  instance  in  the  impressa  group 
(or  in  any  other  Rhytidoponera  species)  was  a single  isolated 
worker,  with  brood,  located  in  the  field,  in  contrast  to  the  frequent 
occurrence  of  single  colony-founding  queens  (see  below). 

The  process  of  colony  fission  is  observationally  difficult  to 
distinguish  from  the  movement  of  a colony  from  one  nest  site  to 
another,  and  the  two  events  may  be  inter-related.  Table  2 summar- 
izes observations  made  on  colony  movement  in  chalybaea  and  in 
three  other  Rhytidoponera  species  (outside  the  impressa  group).  In 
only  one  instance  ( maniae ) was  a single  colony  observed  splitting 
into  two  nests,  but  the  same  event  may  have  been  occurring  during 
the  other  observations,  if  some  workers  remained  at  the  original 
nest  site. 


1981] 


Ward — Rhytidoponera  impressa.  II 


Table  1.  Composition  of  small,  isolated  clusters  of  workers  and  brood  (incipient 
worker-reproductive  colonies?)  in  Rhytidoponera  confusa,  chalybaea,  tasmaniensis, 
and  fulgens. 


Species 

Locality 

Date 

No. 

workers 

Brood 

confusa 

Royal  Natl.  Park, 

5.xi.  1974 

5 

eggs,  larvae 

N.S.W. 

" 

Pearl  Beach,  N.S.W. 

9.iv.  1977 

25* 

9 eggs,  8 larvae 

" 

Seal  Rocks,  N.S.W. 

14.  vi.  1977 

13 

several  larvae 

chalybaea 

Whian  Whian  State 

14. v.  1977 

9 

several  larvae 

Forest,  N.S.W. 

tasmaniensis 

nr.  Wonboyn  Lake, 

25.x.  1975 

6 

none  seen 

N.S.W. 

fulgens 

Mt.  Koghis, 

18. ii.  1977 

4* 

several  larvae. 

New  Caledonia 

one  worker  cocoon 

♦workers  dissected,  one  inseminated. 


In  view  of  the  apparent  scarcity  of  very  small  isolated  clusters  of 
workers  and  brood  (of  the  size  documented  in  Table  1,  i.e.  5-25 
workers),  it  seems  likely  that  colony  fission  in  the  impressa  group 
often  produces  daughter  colonies  larger  in  size.  (Mature  worker- 
reproductive  colonies,  i.e.  those  with  alates,  contain,  on  average, 
about  150  workers.)  More  field  observations  on  budding  are 
needed;  the  small  amount  of  information  accumulated  thus  far 
suggests  that  nocturnal  observations  might  be  rewarding.  It  is  also 
possible  that  some  worker-reproductive  colonies  develop  from 
former  queen-right  colonies  in  which  the  queen  has  died. 

The  origin  and  development  of  queen-founded  colonies  in  the 
impressa  group  has  been  more  extensively  documented.  Incipient 
queen-right  colonies  have  been  observed  repeatedly  in  the  field 
(Table  3).  Mated  queens  apparently  disperse  for  some  distance, 
undergo  dealation,  and  search  for  a suitable  nest  site  (under  stones, 
rotting  logs,  etc.).  Having  located  shelter,  the  queen  excavates  a 
small  cavity,  lays  several  eggs,  and  rears  a small  brood  of  workers, 
the  first  of  these  appearing  within  about  6 months  (3-4  months  in 
lab  colonies).  Unlike  the  claustral  colony  foundation  typical  of 
higher  ants,  queens  forage  outside  the  nest  for  food,  and  feed  their 
larvae  partly  on  insect  prey. 

The  available  field  information  on  incipient,  queenright  colonies 
suggests  that  they  are  usually  founded  in  the  spring  and  early 


Table  2.  Field  observations  on  colony  movement  and  worker  transport  in  Rhytidoponera. 

Time 

Species  Date  Locality  (EDT)  Weather Observations 


112 


Psyche 


[Vol.  88 


s 

8 .2 

5 ^ 

cz  <& 

t— I 

c '-5 

<D 

(U 

t/5  g 
a ea 

C 1/3 

O ID 

c/3  'l“' 

D 2 

-3  o 
O <2 
CO  D 

g 2 

V>  3 

o ^ 

&1  73 

3 . , 


£ I 
I ."2 

o & 
£ c 


« 3 

> t- 

d o 


>3  M 

"O  .2 
3 

_o 
73 


d 

c o 

0 +J 

c « 

1 -g 


n.  2 


c/3 


cZ  cd 

£ (-HH 

^ O 

<D  Qh 
U 0 

% 2 

o 


O 

S'  3 


>> 

3 

"O 

D 

2 a 
2 


D 3 
■£ 

o O 

T3  u 

2 

3 _o 
2 d 

o 5 

o ^ 

8 - 

o d 

D g 
cd  f3 
> £2 


T3 

s 

3 

0 £ 

1 d 
cd  g 

3*  — 

£ c 
d 2 


° <u 

! ? 


cd  c 


,3  '-3 


^ 2 


§ s 

3 D 


2 Id  cd 

Si  £ ~ 

£ is  S 

>3 

3 

H . . ”3 


53  c ~ 

rv  '3  i+h 
2 O 

>> 

§ £ 3 

3 'td  3 

c/3  3 3 


O O 
£ £ 


§ 2 
3 


S 3 

2 ^ 
c/3  DO 
<U  60 

3 ^ 

O DO 

I 'I 


C c/5  O 

> <5  ^ 

S oji 

£ * O 

O 3 7D 
8^3 


3 

o 

S3  2 

o S 


<u  — - 

2 © 
O c/3 

2 .2 
is 

3 C 
3 


D.  2 


G J-h  O 

(U  iu  Z, 


I 

c/3  ^ 


1 § 

DO  £ 


c/3  3 

3 P 

3 C 


C/3  D 
3 U 
D 3 
-*  3 

O C/3 

£ .2 


T3 

O 

DO  'C 
-1  * 
■§ ! i 

£ O 3 


8 e 

o d 
m 3 

<A 

3- 

2 

o 

>T3  m c 

o E. 

VO  O-, 

vd  ci. 

vd 

vd 

cx 

Ov  — 3 

>. 

d 

3 

co 

s- 

<u 

3 

GO  ‘g 

J3 

3 

"i 

>3 

<D 

3 

~o 

0 5? 

• o 

1 3 U 
. . DO 

c;  d 

-GJ  flJ 
^ > 

o 

_ap 

CO 

^ «» 

^ C ^ 

C/5  3 

CO  3 

CO 

"3 

co  3 t: 

Z D 

£ D 

Z 

oa 

3 

z ^ u 

976 

vO 

r- 

Ov 

1977 

VO 

r- 

OV 

O O d 

m r*d  G 

ri  3 d 


cu 

-Ci 

-C 


.xi.1975  NSW:  15  km  c.9.00  cloudy  workers  carrying  brood  and  other  workers  to  one  of 

N Coombah  a.m.  two  new  nest  sites,  0.6  and  2.5  meters,  respectively, 

from  old  nest  site,  the  directions  at  right  angles  to  one 
another;  all  nests  directly  in  soil 


Table  3.  Field  data  on  43  incipient,  queenright  colonies  (with  < 20  workers).  All  Rhytidoponera  confusa  except  the  following 
accessions:  2006  ( chalybaea ),  2620  ( chalybaea ) and  2580  ( impressa ). 

Brood*  Probable 

Accession  Population  Dealate  year  of 

no.  code  no.  Date  Female(s)  Workers  Eggs  Larvae  Cocoons  origin 


1981] 


Ward — Rhytidoponera  impressa.  II 


113 


in 

r- 

o 


m r*~)  m 
r-  r- 
® O'  O' 


't  t 't 

r-  r-~  r- 

Ov  O'  O' 


Tf  Tt  Tt  Tt  Tf 

r — t — r — i~-~ 

O O'  O'  O'  O' 


r-- 

O' 


r-~ 

O' 


r-~ 

O' 


o 

O'  O'  O'  O'  O'  O'  O'  O'  O' 


O-j 1-  OOOOOO  IOOO  I l©H h°  1+0  + 


+ + + OOOOOO  I+  + + I ,+  + + + I + O + 


oo  + oo©H — | — |-  i-| — (-  + i ioH — |-o  i o + o 


m<NTf  OOOOOOOOOOOOOOOO(0-OOrn 


* * ■* 


—I  'O 
ON  (N 


_a.  . t M xj  xf  xl 

+ + + + + + + 

*.  * '*  * * 'y  '*  'y 


Tj-  rf  Tj- 

t"  f-  ^ 

'y  'y  'y  + 

O'  O'  oi  O' 

— — ON 


m «/->  JC? 

r — r~~ 


r- 


<N 

ON 


ON 

ON 


r-~  r-~ 
> > 
o o 


ON 


r-~  r-'  r-~ 


(N  OO  DO  00 
- (N  (N  (N 


'O  rn  OO  OO 
— — rr)  r<-> 


m on  O' 
r-  o r- 

m Tt  Tt 


or^ooo— ■Tj-r'— 'ONmcn— — oo»^> 
r-r-mr-fN>roiAovnoO'Ooo>o«ri«ro«r)t^-ocoo 
fNfNfnm^t^tTfTtTtTtTtrtiniy-iiomiromoooo 


Table  3 (continued). 


114 


Psyche 


[Vol.  88 


(U 

3 

2 ° .s 

o S .2P 

<U  t- 

CU  O 


^3  ed 

O > 

5 s 

to  J 


cd  « 

3 g 
« 6 
Q <U 


a o 

S c 

II 


■d  d1  d-  Tf 

r-  r-  r-  r- 
O'  O'  O'  O' 


r-  r-  r- 
r-  r-" 


NO  NO  *0  NO  NO  NO  NO  NO  NO  'O  NO 
r-r-r-r-r-r-r-r-r-r-r- 
O'  O'  O'  O'  O'  O'  O'  O'  O'  O'  O' 


m 

r- 

m 

>n 

«n 

NO 

NO 

(v 

NO 

NO 

NO 

NO 

r-~ 

r-~ 

t — 

r» 

r- 

t" 

I — 

’> 

X 

X 

x 

’x 

X 

X 

X 

X 

'x 

rn 

in 

O' 

r-‘ 

in 

oo 

CN 

ro 

>n 

[T;  £;  r-  r-  r'- 

r':  r-  i"  r- 

.>  .►  > > -• 

^ ® £ ~ o; 


X X 
oo  on 


oo  — < o m 
© oo  t~-  (N 
oo  oo  O'  m 


no  oo  o 

no  Tf  rsi 
m >n  © 
— — <N 


— 00'0'0(NmTTt~-<NOO 
hXO'OMf'l'J'thaOM 
O'O'O'OOTtTj-i^Tj-m'o 

— — - MM(N(N(N(N(N(N 


O'  o 
Tf  «n 
00  oo 

<N  <N 


brood:  +,  present;  0,  absent;  -,  no  information 


1981] 


Ward — Rhytidoponera  impressa.  II 


115 


Table  4.  Composition  of  322  confusa  colonies,  with  respect  to  numbers  of  cocoons 
and  alates,  and  month  of  collection.  Large  standard  deviations  are  due  to  variation  in 
colony  size,  and  to  the  fact  that  not  all  nests  produce  alates  in  a given  season. 


Month 

Sample  Size 
<#colonies) 

# worker  cocoons 
mean  ± S.D. 

# alate  cocoons 
mean  ± S.D. 

# alates 
mean  ± S.D. 

Sept. 

13 

0.0±  0.0 

0.0+  0.0 

10.8+13.9 

Oct. 

47 

0.2±  0.5 

0.0+  0.0 

5.3+12.7 

Nov. 

42 

3.6+  6.5 

0.0+  0.0 

0.2+  1.5 

Dec. 

13 

16.3+27.7 

5.7+14.6 

0.1+  0.3 

Jan. 

24 

43.0+55.3 

16.9+38.0 

0.0+  0.0 

Feb. 

10 

62.8+40.1 

15.9+15.7 

1.0+  2.8 

Mar. 

15 

20.5+25.2 

3.9+  5.2 

6.9+  8.3 

Apr. 

26 

27.5+32.5 

1.8+  4.4 

23.1+29.9 

May 

41 

0.4+  1.2 

0.0+  0.0 

12.7+21.4 

June 

50 

0.1+  0.3 

0.0+  0.0 

25.5+39.0 

July 

32 

0.0+  0.0 

0.0+  0.0 

25.0+29.7 

Aug. 

9 

0.0+  0.0 

0.0+  0.0 

21.9+34.1 

summer,  and  that  development  proceeds  rather  slowly.  Of  the  17 
dealate  females  collected  with  eggs  or  no  brood  at  all,  13  came  from 
spring  and  early  summer  months  (October-December)  and  only  4 
from  the  fall  (April-May)  (Table  3).  These  findings  are  consistent 
with  the  observation  that  virgin  alates  usually  remain  in  the  nests 
throughout  the  winter,  and  fly  in  the  spring.  Nevertheless  the 
occurrence  of  a few  incipient  colonies  in  apparently  early  stages  of 
development  in  April  and  May  requires  some  explanation:  it  seems 
likely  that  either  development  was  hindered  in  these  colonies  or  that 
occasional  fall  mating  flights  occur. 

Colony  foundation  in  the  spring  and  early  summer  appears  to  be 
the  pattern  followed  \n  purpurea:  Brown  (1954)  noted  many  colony- 
founding dealate  females  of  this  species  in  October  and  November 
on  the  Atherton  Tableland,  north  Queensland. 

Forty-one  of  the  43  incipient  colonies  listed  in  Table  3 contained 
only  a single  queen.  The  two  instances  of  primary  pleometrosis 
(colony  foundation  by  more  than  one  queen)  both  involved  colonies 
in  a very  early  stage  of  development,  without  brood.  One  of  these 
pairs  (acc.  no.  1996)  was  brought  into  the  lab,  and  colony 
development  was  monitored.  The  two  queens  cohabited  peacefully 
from  3 October,  1976  until  the  end  of  December,  at  which  time  the 
colony  contained  12  eggs,  8 larvae  and  8 worker  cocoons.  The  first 
worker  emerged  2 January,  1977;  four  days  later  (after  a second 


116 


Psyche 


[Vol.  88 


ALATES 


ALATE 

COCOONS 


WORKER 

COCOONS 


LARVAE 


EGGS 


i i i l l i i l l I i i i 

SEP  OCT  NOV  DEC  JAN  FEB  MAR  APR  MAY  JUN  JUL  AUG 

(13)  (47)  (42)  (30)  (25)  (13)  (28)  (30)  (72)  (107)  (32)  (39) 

Figure  1.  Seasonal  changes  in  brood  and  alate  composition  in  colonies  of  confusa 
and  chalybaea,  as  measured  by  the  proportion  of  colonies  with  various  life  stages. 
Maximum  width  (shown  at  either  side  of  figure)  indicates  that  100%  of  colonies 
contain  the  particular  stage.  Because  there  were  no  obvious  differences  between  species 
or  between  years,  data  covering  both  species  over  3‘/2  seasons  have  been  combined. 
Figures  in  parentheses  refer  to  the  number  of  colonies  sampled  in  each  month.  Total 
sample  size:  479  colonies. 


worker  had  eclosed)  one  queen  was  found  ousted  from  the  nest  and 
almost  dead.  The  colony  (with  one  remaining  queen)  continued  to 
develop  until  artificially  terminated  15  months  later.  Subsequent 
spermathecal  dissections  and  electrophoretic  analysis  using  allo- 
zyme  markers  confirmed  that  both  females  were  inseminated,  and 
that  both  had  contributed  worker  offspring  to  the  incipient  colony. 

Seasonal  cycle  in  mature  colonies 

There  are  consistent  seasonal  patterns  in  the  occurrence  of  brood 
and  alates  in  mature  colonies  of  the  impressa  group.  These  seasonal 
patterns  are  essentially  the  same  for  both  queenright  and  worker- 
reproductive  colonies,  except  that  alates  in  the  latter  are  pre- 


1981] 


Ward — Rhytidoponera  impressa.  II 


117 


MAR  APR  MAY  JUN  JUL  AUG  SEP  OCT  NOV  DEC  JAN 

(7)  (19)  (20)  (24)  (27)  (6)  (13)  (39)  (20)  (4)  (4) 

Figure  2.  Maximum  larval  length  per  nest  (measured  in  millimeters,  with  larva  in 
natural  resting  position),  in  relation  to  time  of  year,  for  1 83  nests  of  confusa.  One  larva 
(the  largest)  was  measured  in  each  nest.  Figures  in  parentheses  indicate  the  number  of 
nests  sampled.  No  data  available  for  February. 

dominantly  male  only.  Figure  1 summarizes  the  seasonal  changes 
for  confusa  and  cha/ybaea,  in  terms  of  the  proportion  of  colonies 
containing  various  stages  of  brood  or  alates,  for  each  month.  Data 
on  absolute  numbers  of  worker  cocoons,  alate  cocoons,  and  alates 
are  given  in  Table  4 for  322  confusa  colonies  (similar  patterns  are 
shown  by  chalybaea,  but  numbers  average  higher). 

In  confusa  and  chalybaea  there  appear  to  be  two  peaks  of  egg 
production — one  in  the  spring  (September-October)  and  another  in 
the  late  fall  (February- April).  Larvae  are  continually  present;  those 
overwintering  are  small  to  medium-sized  and  show  little  growth 
until  the  spring,  when  development  proceeds  rapidly  (Figure  2). 
Worker  cocoons  first  appear  in  October-November,  and  not  until 


118 


Psyche 


[Vol.  88 


approximately  two  months  later  do  the  first  alate  cocoons  appear 
(December-January).  Adult  workers  emerge  from  January  until 
June,  while  alates  eclose  over  a shorter  time  period  (February- 
April).  Most,  or  all,  of  the  alates  overwinter  in  the  nest,  and  are 
released  in  the  spring,  possibly  in  several  bursts,  since  some  nests 
have  been  found  with  alates  as  late  as  November  (and  one  nest  with 
a single  male  in  early  December). 

Several  points  of  interest  emerge  from  the  foregoing: 

(1)  There  is  only  one  crop  of  adults  produced  each  year,  and 
alate  production  is  restricted  to  a limited  period  of  the  total  time 
that  new  offspring  are  produced.  At  any  given  time,  the  standing 
crop  of  new  cocoons  consists  on  average  of  no  more  than  about  30% 
alates  (Table  5).  These  facts  may  be  relevant  to  a consideration  of 
control  of  the  sex  ratio  of  investment. 

(2)  No  cocoons  are  overwintered,  and  there  is  a period  of  6 
months  (July-December)  when  no  new  individuals  are  added  to  the 
workforce.  At  first  glance,  this  would  seem  detrimental  to  the 
increased  foraging  requirements  during  rapid  larval  growth  in  the 
spring  and  early  summer.  However,  because  of  a time  lag  between 
worker  eclosion  and  foraging  (callow  workers  remain  in  the  nest)  it 
may  in  fact  produce  an  effective  increase  in  the  foraging  force  when 
it  is  most  needed. 

(3)  In  the  absence  of  data  on  sex-  and  caste-specific  growth 
rates,  it  is  difficult  to  know  whether  alates  arise  from  the  overwin- 
tering larvae  (hence,  from  eggs  laid  the  previous  season)  or  from 
eggs  laid  in  the  spring.  However  one  piece  of  evidence  suggests  the 
latter:  the  discrepancy  between  the  appearances  of  the  first  worker 
and  first  alate  cocoons  (two  summer  months)  seems  to  be  too  great 
to  be  explained  by  assuming  that  equivalent-sized  overwintering 
larvae  require  that  extra  period  of  time  (and  quantity  of  food)  to 
develop  into  reproductives.  Rather,  it  would  seem  more  likely  that 
the  reproductives  develop  from  spring-laid  eggs  or  alternatively 
from  smaller  overwintering  larvae. 

Not  all  nests  of  confusa  and  chalybaea  contain  alates  in  a given 
season,  alate  production  being  associated  with  larger  colony  sizes 
(Table  5).  Nevertheless,  there  is  considerable  overlap  in  colony  size 
between  nests  with  and  without  alates,  partly  due  to  the  fact  that 
worker-reproductive  colonies  produce  alates  at  a smaller  size  (and 
probably  younger  age)  than  queenright  colonies  (Ward,  1978).  It 
seems  likely  that  a variety  of  genetic,  environmental,  and  develop- 


1981] 


Ward — Rhytidoponera  impressa.  II 


119 


Table  5.  Mean  colony  size  (number  of  workers)  for  nests  with  alates  and  for  those 
without  alates  at  the  time  of  year  (February- September)  when  winged  reproductives 
are  normally  present. 

Mean  no.  workers  (±  S.D.) 

Species  Alates  present  Alates  absent 

confusa  203. 1 + 179.9  (n=  132)  83.9+  65.7  (n=41) 

chalybaea  270.7+206.2  (n=68)  146.7+122.4  (n=41) 

mental  (ergonomic)  factors  influence  the  production  of  alates. 

The  available  information  on  impressa  and  purpurea  indicates  a 
seasonal  brood  cycle  similar  to  that  of  confusa  and  chalybaea.  Nests 
of  the  two  former  species  collected  in  the  winter  in  Queensland 
generally  had  small  larvae  (sometimes  eggs),  alates,  and  few  or  no 
cocoons  (sample  of  8 impressa  colonies,  16  purpurea  colonies).  A 
lowland  population  of  purpurea  from  near  Cape  Tribulation,  north 
Queensland,  was  exceptional  in  overwintering  with  mature  larvae, 
as  well  as  worker  cocoons  and  adult  alates.  Nothing  is  known  of  the 
brood  cycle  in  New  Guinea  populations  of  purpurea  which  inhabit 
much  less  seasonal  environments. 

Thus,  in  Australia  at  least,  four  species  in  the  impressa  group 
produce  one  brood  of  sexuals  a year,  most  or  all  of  which  are 
overwintered  in  the  nest  and  released  in  the  spring.  This  occurs 
despite  contrasting  climatic  regimes  at  the  north-south  extremes  of 
range  (summer  rainy  season  in  the  north,  and  winter  rains  in  the 
south)  (cf.  Brown,  1954). 

Collections  of  enigmatica  suggest  a similar  brood  cycle  (i.e.  small, 
overwintering  larvae;  cocoons  present  only  in  summer),  with  one 
important  distinction:  alates  are  usually  absent  from  nests  in  the 
winter.  Of  13  nests  collected  in  the  early  winter  (April  30- July  1) 
only  one  contained  alates  (all  males);  on  the  other  hand,  four  out  of 
five  nests  collected  in  the  summer  (January  12- March  7)  contained 
alate  pupae  (also  all  males).  The  differences  are  significant  (p  < .02, 
two-tailed  Fisher’s  exact  test),  and  suggest  that  alates  fly  principally 
in  the  fall.  If  this  is  so,  there  would  appear  to  be  considerable 
temporal  isolation  between  enigmatica  and  its  two  sympatric 
congeners  (< chalybaea  and  confusa). 

Mating  Flights 

Two  pieces  of  indirect  evidence  suggest  that  reproductives  of 
confusa  and  chalybaea  normally  mate  in  the  spring: 


120 


Psyche 


[Vol.  88 


(i)  the  proportion  of  nests  containing  alates  is  more  or  less 
constant  throughout  the  late  summer,  fall  and  winter, 
dropping  rapidly  in  the  spring;  and 

(ii)  there  is  a flush  of  colony-founding  females  in  the  early  to 
mid-summer.  To  the  extent  that  impressa  and  purpurea 
share  the  same  seasonal  cycle,  it  may  be  supposed  that  their 
nuptial  flights  also  occur  in  the  spring. 

Alates  of  confusa  and  chalybaea  were  observed  actively  dispers- 
ing or  swarming  on  several  occasions  in  rainforest  and  urban 
parkland  in  the  Sydney  region.  All  observations  but  one  (out  of  15) 
were  made  in  the  spring  (September  15-November  10),  and  the  only 
large-scale  mating  swarms  were  seen  at  this  time.  Most  observations 
involved  congregations  of  males  around  nest  entrances.  On  six 
occasions,  isolated  male  or  female  alates  were  observed  away  from 
the  nest,  apparently  in  a dispersing  phase.  Spring  mating  flights 
were  observed  for  3 consecutive  years  (1976-78)  in  the  chalybaea 
population  occurring  on  the  University  of  Sydney  campus.  Because 
of  the  scarcity  of  information  on  this  important  stage  of  the  life 
cycle,  the  1976  mating  swarm  is  described  in  detail. 

This  flight  took  place  on  4 October  1976,  a mild  overcast  day  with 
brief  periods  of  sunshine  and  light  rain.  At  the  time  observations 
were  begun  (10:15  a.m.  EST)  large  numbers  of  chalybaea  alates, 
mostly  males,  were  observed  flying  in  parts  of  the  University 
campus.  Alates  were  distinctly  concentrated  into  clusters  in  tree- 
shaded  areas.  Three  of  these  concentrations  were  examined  in  detail 
(Sites  A,  B and  C in  Figure  3). 

Site  A.  This  cluster  was  centered  about  a chalybaea  nest  entrance 
between  two  slabs  of  sandstone  which  formed  part  of  a stone  wall. 
Between  10:45  and  11:45  a.m.  there  were  several  hundred  males 
within  2 meters  of  the  nest  entrance.  No  alate  females  were  seen. 
Although  males  spent  most  of  the  time  on  the  ground  chasing  other 
individuals,  the  congregation  appeared  to  be  formed  by  males  flying 
into  the  site.  There  were  large  numbers  of  workers  milling  around 
the  nest  entrance  and  most  behaved  aggressively  towards  the  males, 
but  this  did  not  deter  the  latter  from  making  repeated  attempts  to 
mate  with  workers  (and  with  other  males).  Three  apparently 
successful  male-worker  matings  were  observed;  in  each  instance  the 
pair  was  already  in  copulation  when  discovered,  in  a position 
similar  to  that  described  by  Holldobler  & Haskins  (1977)  for  R. 
metallica.  The  worker  dragged  the  male  on  the  ground  for  about  30 


1981] 


Ward — Rhytidoponera  impressa.  II 


121 


Figure  3.  Section  of  University  of  Sydney  campus  where  chalybaea  nuptial  flights 
were  observed.  A,  B,  and  C represent  sites  of  large  clusters  of  alates,  described  in  text. 
Lesser  numbers  of  alates  were  observed  at  locations  U3,  and  elsewhere. 

seconds,  after  which  separation  occurred.  Wings  were  vibrated 
rapidly  during  attempts  by  males  to  mount  workers.  A mating 
attempt  by  one  male  often  attracted  others,  resulting  in  a buzzing, 
frolicking  ball  of  males.  Males  were  also  observed  to  enter  (and 
leave)  the  nest,  and  may  perhaps  have  mated  with  workers  within 
the  nest.  Several  instances  were  noted  of  workers  forcibly  evicting 
males  from  the  nest,  dragging  them  to  a distance  of  1 meter  from  the 
nest  entrance.  Workers  were  still  foraging  during  these  events:  two 
which  were  observed  returning  with  a dead  honey-bee,  and  another 
with  a seed,  were  unmolested  by  males. 

Sites  B and  C.  Similar  observations  were  made  at  these  sites 
(Figure  3),  with  large  numbers  (>  100)  of  alates  and  workers 
clustered  in  the  vicinity  of  nest  entrances,  along  sandstone  walls.  A 
few  alate  females  were  also  seen  among  these  swarms.  Despite 
persistent  attempts  by  males,  no  successful  worker-male  or  queen- 
male  matings  were  recorded.  There  was  a noticeable  decline  in 
swarming  activity  by  early  afternoon. 

Male  alates  were  observed  in  smaller  numbers  at  several  other 
places  on  campus,  particularly  at  Sites  1,  2,  and  3 (Figure  3).  A 
single,  inseminated  dealate  female  was  encountered  at  Site  3 in  mid- 
afternoon, apparently  searching  for  a nest  site.  During  the  day, 
samples  of  alates  were  collected  from  each  observation  site.  Out  of  a 
total  of  293  alates,  279  (95.2%)  were  males,  and  14  (4.8%)  were 
females. 


122 


Psyche 


[Vol.  88 


Later  in  the  evening  (10:45  p.m.),  lesser  numbers  of  alate  males 
were  present,  but  inactive,  on  the  ground  at  various  locations.  Over 
the  next  3 weeks  small  congregations  of  males  were  seen  around  nest 
entrances,  but  never  in  such  numbers  or  frenzied  activity  as  during 
the  large-scale  swarm  of  4 October. 

The  Sydney  University  population  of  chalybaea  consists  princi- 
pally of  worker-reproductive  colonies,  so  the  preponderance  of 
males  among  alates  is  not  surprising.  The  1976  mating  swarms 
apparently  involved  insemination  of  both  workers  and  queens.  Only 
one  mated  queen  was  found,  however,  and  it  remains  unclear  if 
queens  mate  predominantly  in  the  vicinity  of  nest  entrances  or  in 
separate  rendevous  sites. 

Given  the  limited  number  of  successful  matings  observed,  it  is 
conceivable  that  the  mating  swarm  had  already  passed  a peak  of 
activity  at  the  time  that  observations  began  (10:15  a.m.).  This  is  also 
suggested  by  the  absence  of  workers  in  a sex  pheromone-releasing 
posture  (as  described  by  Holldobler  and  Haskins  (1977)  for  R. 
metallica).  Such  “calling”  workers  were  observed  in  lab  colonies  of 
chalybaea , where  the  behavior  occurred  both  inside  and  outside  the 
artificial  nest.  The  posture  adopted  was  similar  to  that  described  for 
metallica  (i.e.,  head  and  mesosoma  lowered,  gaster  raised  and 
arched,  with  tergites  exposed).  In  addition,  workers  repeatedly 
rubbed  the  sides  of  the  gaster  with  their  hind  tibiae,  presumably 
facilitating  release  of  pygidial  (=tergal)  gland  pheromone.  Such 
rubbing  movements  have  been  reported  in  Amblyopone  pallipes 
queens  (Haskins,  1979)  but  not  previously  in  Rhytidoponera. 

A mated  worker  from  one  of  the  copulating  pairs  observed  at  Site 
A was  isolated  in  a modified  Janet  (plaster-of-Paris)  nest  in  the  lab 
and  fed  on  honey  and  Drosophila.  On  1 1 November  the  first  egg 
was  seen,  and  by  21  December  there  were  2 eggs,  1 larva  and  1 
worker  cocoon.  Just  before  the  colony  was  terminated,  in  March, 
1977,  this  mated  worker  had  produced  three  worker  offspring  (the 
first  had  appeared  on  19  January  1977).  This  is  perhaps  the  first 
record  among  the  Formicidae  of  colony-foundation  by  a lone 
worker.  However,  as  mentioned  previously,  there  is  no  evidence  that 
single  workers  found  colonies  in  the  field  and  it  appears  that  they 
are  always  accompanied  by  an  entourage  of  uninseminated  workers. 

The  inseminated  dealate  female,  also  collected  on  4 October  1976, 
was  kept  under  similar  lab  conditions  for  five  months.  The  first 
worker  appeared  on  7 January  1977.  At  time  of  termination 


1981] 


Ward—  Rhytidoponera  impressa.  II 


123 


(March,  1977)  the  colony  consisted  of  1 queen,  1 worker,  1 worker 
^cocoon,  1 larva  and  several  eggs. 

The  following  spring,  in  the  morning  and  early  afternoon  of  1 
October  1977  another  large  mating  swarm  of  chalybaea  occurred  on 
the  University  of  Sydney  campus.  As  before,  this  consisted  mostly 
of  male  alates,  concentrated  into  more  or  less  discrete  clusters 
around  several  nest  entrances.  Large  clusters  were  situated  at  Sites 
A and  C (Figure  3),  at  exactly  the  same  places  observed  in  1976.  No 
matings  were  directly  observed,  but  a timid  worker  which  was  being 
mobbed  by  males  was  later  found  to  be  inseminated.  Workers  were 
generally  very  aggressive  towards  males,  but  the  latter  persisted  in 
attempts  to  mate.  Once  again,  samples  of  alates  were  collected  from 
various  sites,  of  which  97.0%  (195)  were  males  and  3.0%  (6)  were 
females.  These  figures  are  not  significantly  different  from  those  of 
1976. 

On  October  12,  1978  small  swarms  (20-30  individuals)  of 
Rhytidoponera  chalybaea  males  were  observed  at  Site  C and  at 
several  other  locations  on  campus  (but  not  Site  A).  At  10:15  a.m. 
males  were  mostly  at  nest  entrances,  apparently  in  the  process  of 
emerging.  One  alate  female  was  observed;  this  individual  emerged 
from  a nest  entrance,  and  flew  off  into  open  sky,  ascending  rapidly. 
Similar  behavior  was  observed  in  males.  By  1 1:30  a.m.  many  males 
appeared  to  be  flying  into  the  area,  congregations  had  formed 
outside  nest  entrances,  and  males  made  repeated  attempts  to  mate 
with  workers. 

On  the  afternoon  of  the  same  day  two  chalybaea  queens  (one 
alate,  one  partially  dealate)  were  seen  floundering  on  the  sidewalk  in 
a heavily  built-up  section  of  downtown  Sydney.  Both  were  unin- 
seminated. This  suggests  that  alate  females  may  disperse  a con- 
siderable distance  before  mating. 

Colonly  Structure  and  Life  Cycle 

In  most  populations  of  confusa,  chalybaea,  and  impressa,  queen- 
right  and  worker-reproductive  colonies  coexist,  in  intermediate 
proportions.  Despite  the  likely  disparity  between  mating  sites  of 
winged  queens  and  workers,  genetic  data  from  electrophoretic 
studies  (Ward,  1978,  1980)  reveal  no  indication  of  extensive 
inbreeding  or  assortative  mating  with  respect  to  colony  type.  This  is 
consistent  with  the  observation  that  brood  development  and  alate 
production  proceed  at  similar  rates  in  the  two  colony  types,  and  that 


124 


Psyche 


[Vol.  88 


release  of  alates  in  worker-reproductive  colonies  occurs  syncronous- 
ly  with  (or  at  least  in  the  same  season  as)  queenright  colonies. 

As  for  the  remaining  species  in  the  impressa  group,  only 
queenright  nests  are  known  in  purpurea  and  this  species  shows  a 
brood  development  pattern  similar  to  the  three  others.  By  contrast, 
distinct  winged  queens  are  unknown  in  enigmatica  (all  recorded 
colonies  worker-reproductive),  and  this  species  diverges  from  its 
closely  related  congeners  by  releasing  most  alates  in  the  fall, 
although  males  were  found  overwintering  in  one  nest.  The  limited 
information  indicates  a possible  relaxation  of  synchrony  in  the 
release  of  ergatoid-seeking  male  alates,  a pattern  which  would  be 
predicted  with  the  loss  of  the  winged  queen  caste,  especially  if  the 
sexual  calling  behaviour  of  ergatoid  gynes  is  temporally  dispersed. 
This  trend  is  continued  in  some  other  Rhytidoponera  species 
outside  the  impressa  group,  in  which  functional  queens  are  rare  or 
absent,  and  flights  of  alates  (males)  are  reported  to  be  highly  non- 
specific with  respect  to  season  (Brown,  1958;  Haskins  & Whelden, 
1965;  Haskins,  1979).  However,  since  most  of  the  data  come  from 
lab  colonies  of  one  species  ( metalliea ) additional  field  observations 
are  desirable. 

Scattered  collections  of  colonies  from  different  times  of  the  year 
may  give  a misleading  impression  of  patterns  of  alate  production.  In 
at  least  two  species  of  the  impressa  group,  alates  can  be  found  in 
some  nests  from  February  to  November.  Although  this  superficially 
suggests  aseasonal  production  of  alates,  a detailed  examination  of 
brood  development  demonstrated  that  only  one  crop  of  alates  is 
produced  each  year  and  that  alates  are  released  over  a limited  time 
period.  Additional  field  studies  are  necessary  to  determine  whether 
brood  development  in  Rhytidoponera  species  without  queenright 
colonies  is  less  constrained  by  the  need  for  synchronous  alate 
release.  For  comparison  with  the  impressa  group,  such  studies 
would  be  most  appropriately  directed  towards  other  species  of  east 
Australian  mesic  forests,  in  order  to  minimize  climatic  and  other 
environmental  differences. 

Summary 

In  the  Rhytidoponera  impressa  group  there  are  two  kinds  of 
colonies,  which  are  distinguished  by  the  type  of  reproductive  female 
present:  queenright  colonies  with  a single  dealate  queen,  and 


1981] 


Ward — Rhytidoponera  impressa.  II 


125 


worker-reproductive  colonies  in  which  one  or  more  mated  “work- 
ers” occur  in  lieu  of  a queen.  It  appears  that  worker-reproductive 
colonies  normally  reproduce  by  colony  fission  or  budding,  although 
information  on  this  process  is  fragmentary.  Queenright  colonies  are 
founded  by  lone  queens.  Colony-founding  queens  are  most  fre- 
quently encountered  in  the  spring  and  early  summer;  such  queens 
leave  the  brood  chamber  to  forage  for  food. 

In  mature  colonies  of  confusa  and  chalybaea,  the  development  of 
brood  and  production  of  alates  is  highly  seasonal  (and  essentially 
similar  for  both  queenright  and  worker-reproductive  colonies).  One 
crop  of  workers  and  alates  is  produced  each  year,  the  former 
eclosing  from  cocoons  between  January  and  June,  the  latter 
between  February  and  April.  Most  or  all  alates  overwinter  in  the 
nest  (along  with  small  to  medium-sized  larvae),  and  are  released  in 
the  spring  (September-November).  Similar  seasonal  patterns  are 
shown  by  impressa,  purpurea  (in  Australia),  and  enigmatica,  except 
that  colonies  of  enigmatica  generally  do  not  retain  alates  over  the 
winter. 

In  the  population  of  chalybaea  on  the  University  of  Sydney 
campus,  mating  flights  took  place  in  early  October  for  3 consecutive 
years.  During  these  flights,  flying  males  became  concentrated  into 
clusters  around  nest  entrances  where  they  attempted  to  mate  with 
workers,  with  males,  and  with  the  occasional  alate  female.  Several 
worker-male  but  no  queen-male  matings  were  observed  in  these 
nest-associated  swarms.  Like  males,  queens  appear  to  disperse  some 
distance  before  mating,  and  possibly  utilize  mating  sites  other  than 
nest  entrances. 


Acknowledgements 

This  work  was  supported  by  an  Australian  Commonwealth 
Scholarship.  Additional  support  from  L.  C.  Birch  and  the  Univers- 
ity of  Sydney  is  gratefully  acknowledged.  1 thank  D.  Feener  and  A. 
Forsyth  for  comments  on  the  manuscript. 

References 


Brown,  W.  L. 

1953.  Characters  and  synonymies  among  the  genera  of  ants.  Part  I.  Breviora, 

11,  1 13. 


126 


Psyche 


[Vol.  88 


1954.  Systematic  and  other  notes  on  some  of  the  smaller  species  of  the  ant 
genus  Rhytidoponera  Mayr.  Breviora,  33,  1 11. 

1958.  Contributions  toward  a reclassification  of  the  Formicidae.  II.  Tribe 
Ectatommini  (Hymenoptera).  Bull.  Mus.  Comp.  Zool.  Harvard,  118, 
175-362. 

Haskins,  C.  P. 

1979.  Sexual  calling  behavior  in  highly  primitive  ants.  Psyche,  85,  407-415. 

Haskins,  C.  P.  and  W.  M.  Whelden 

1965.  “Queenlessness”,  worker  sibship,  and  colony  versus  population  structure 
in  the  formicid  genus  Rhytidoponera.  Psyche,  72,  87-1  12. 

Holldobler,  B.  and  C.  P.  Haskins 

1977.  Sexual  calling  behavior  in  primitive  ants.  Science,  195,  793-794. 

Ward,  P.  S. 

1978.  Genetic  variation,  colony  structure,  and  social  behaviour  in  the  Rhyti- 
doponera impressa  group,  a species  complex  of  ponerine  ants.  Ph.D. 
Thesis,  University  of  Sydney. 

1980.  Genetic  variation  and  population  differentiation  in  the  Rhytidoponera 
impressa  group,  a species  complex  of  ponerine  ants  (Hymenoptera: 
Formicidae).  Evolution,  34,  1060-1076. 

1981.  Ecology  and  life  history  of  the  Rhytidoponera  impressa  group  (Hymen- 
optera: Formicidae).  I.  Habitats,  nest  sites,  and  foraging  behavior. 
Psyche,  88:  89-108. 


THE  ONTOGENY  OF  LYSSOMANES  VIRIDIS 
(WALCKENAER)  (ARANEAE:  SALTICIDAE) 
ON  MAGNOLIA  GRANDIFLORA  LJ 

By  David  B.  Richman  and  Willard  H.  Whitcomb 

Department  of  Entomology  and  Nematology 
University  of  Florida,  Gainesville,  Florida,  32611 


Introduction 

Lyssomanes  viridis  (Walckenaer)  is  a translucent  green  spider 
found  in  the  southeastern  United  States  from  North  Carolina  to 
Florida  and  Texas  (Kaston  1978).  It  has  sometimes  been  placed  in  a 
separate  family,  Lyssomanidae,  but  the  most  recent  taxonomic 
study  (Galiano  1976)  includes  it  in  the  Salticidae.  This  species 
commonly  lives  on  the  tree  Magnolia  grandiflora  L.  in  mesic 
situations,  on  palmettoes  in  various  habitats,  and  on  Lyonia  sp.  and 
other  shrubs  in  the  sand  pine  scrub  of  central  Florida.  No  complete 
life  cycle  has  been  published  for  any  Lyssomanes  species.  Crane 
(1950)  did  present  descriptions  of  the  early  stages,  including  second 
postembryo  (her  first  instar)  and  first  instar  (her  second  instar)  of  L. 
bradyspilus  Crane.  The  current  paper  is  the  result  of  a total  of  two 
and  one  half  years  of  collection  and  observation  of  a natural 
population  of  L.  viridis  for  the  purpose  of  learning  about  the 
ontogeny  of  this  spider  in  the  wild. 

Methods 

Eggs,  immatures,  and  adults  of  L . viridis  were  collected  on  the 
undersides  of  leaves  by  turning  the  leaves  and  catching  the  spiders  in 
vials  in  a stand  of  Magnolia  grandiflora  at  Tall  Timbers  Research 
Station,  Leon  County,  Florida  (Figure  1).  These  were  collected 
monthly  from  August  1971  to  February  1973  and  usually  twice  a 
month  from  June  1977  through  June  1978.  The  population  densities 
were  measured  by  counting  the  number  of  spiders  collected  per  1000 
leaves  from  September  1977  through  the  death  of  the  adults  and  the 


'Florida  Agricultural  Experiment  Station  Journal  Series  No.  3059. 
Manuscript  received  by  the  editor  July  10,  1981 


127 


128 


Psyche 


[Vol.  88 


Figure  1.  Magnolia  stand  at  Tall  Timbers  Research  Station,  Leon  County, 
Florida.  Figure  2.  Gravid  female  of  Lyssomanes  viridis  (Walckenaer).  Figure 
3.  Eggs  of  Lyssomanes  viridis  on  underside  of  magnolia  leaf.  Figure  4.  First 
postembryos  of  Lyssomanes  viridis  on  magnolia  leaf. 

rise  of  immatures  in  June  1978.  Leaves  were  counted  arbitrarily  as 
the  stand  was  circled.  Collections  were  made  from  0 to  2 m above 
ground  level  ,on  both  the  outside  edge  and  within  the  stand.  All 
spiders  with  prey  were  preserved  separately  and  identifications  of 
the  prey  obtained  from  various  specialists.  Carapace  widths  were 
measured  using  a dissecting  microscope  equipped  with  an  ocular 
micrometer  for  20  specimens  per  sample,  if  available.  The  number 
of  instars  was  calculated  by  using  a method  of  simple  regression  of 
carapace  widths  developed  by  Hagstrum  (1971).  It  was  assumed  that 
salticids  exhibit  a similar  mean  relationship  between  logarithms  of 
carapace  width  and  stadium  (log  y=0.0871x  — 0.2692  where 
x=stadium  and  y=carapace  width)  as  Lycosids,  Loxoscelids, 
Clubionids,  Oxyopids,  and  Ctenizids.  Egg  masses  collected  on  the 
magnolia  trees  were  allowed  to  hatch  and  the  time  spent  in  the  first 


1980] 


Richman  & Whitcomb — Lyssomanes  viridis 


129 


and  second  postembryonic  stages  was  measured  in  the  laboratory; 
however,  these  were  not  raised  through  the  various  instars. 

Weather  information  for  Tall  Timbers  was  obtained  from  the 
research  station.  Rainfall  during  August  1971  to  February  1973 
averaged  11.4  cm  (SD  = 7.6  cm)  with  33.7  cm  falling  in  June  1972 
(hurricane  Agnes)  and  only  1.6  cm  falling  in  September  1972. 
Rainfall  during  July  1977  through  June  1978  averaged  10.9  cm(SD 
= 4.3  cm)  per  month  with  a maximum  of  16.9  cm  in  August  1977 
and  a minimum  of  3.7  cm  in  October  1977.  Relative  humidity 
almost  always  reached  100%  at  some  time  during  the  day  except  for 
a few  days  during  the  winters. 

Results  and  Discussion 

We  found  that  mating  took  place  in  May  and  that  the  males 
disappeared  by  mid-June.  Some  females  lingered  on  at  least  until 
August.  Gravid  females  (Figure  2)  laid  25-70  eggs  (mean  42.7, 
SD=11.6,  no. =24)  at  a height  of  33-131  cm  (mean  87.9,  SD=28.0, 
no. =12)  on  the  magnolia  stand  from  May  31  to  July  6.  Second 
clutches  may  have  been  produced  only  occasionally  as  females 
usually  guarded  the  eggs  until  first  instar  and  females  laid  second 
clutches  only  twice  (infertile)  in  the  laboratory  (not  included  in  egg 
counts).  The  bright  green  eggs  (Figure  3)  were  ca.  1 mm  in  diameter 
and  were  loosely  covered  by  silk  (there  was  no  distinct  cocoon).  The 
first  postembryonic  stage  (Figure  4 — chorion  molted)  lasted  32-35.5 
hours  (no.  of  egg  masses  = 4)  and  the  second  postembryonic  stage 
(legs  free  of  vitelline  membrane)  lasted  7 days  (no.  of  egg  masses  = 
5).  The  carapace  widths  (Figure  5)  indicated  that  there  were 
probably  7 instars  including  adult  female  after  second  postembryo, 
based  on  Hagstrum’s  (1971  Figure  1)  data  for  laboratory  reared 
Lycosidae,  Loxoscelidae,  Clubionidae  and  Oxyopidae  and  for  field 
collected  Ctenizidae.  Males  may  have  one  less  instar  than  females. 
The  immature  stage  individuals  lasted  from  June  to  the  next  May 
when  most  matured  (Figure  5).  Spiderling  first  instars  occurred 
from  June  to  July,  most  individuals  reaching  second  instar  by  the 
first  of  August.  The  majority  reached  third  to  fourth  instar  in 
September  and  passed  through  the  winter  as  third  to  fifth  instars. 
The  female’s  sixth,  or  penultimate,  instar  started  to  be  evident  in 
March.  Courtship  was  observed  by  Richman  (in  press). 

Immature  spiders,  especially  early  instars,  fed  primarily  on 


130 


Psyche 


[Vol.  88 


2.0 


6 

□ 


9 9 9 
<3 


o 


o° 


_ 1.5- 

E 

E 


O 

< 

< l.Oh 

CL 

< 
o 


o 


□ 


.5- 


jjasondjfmamjj 

Figure  5.  Mean  carapace  width  of  immature  and  adult  Lyssomanes  viridis 
(Walckenaer)  at  Tall  Timbers,  Leon  County,  Florida  1971-1973  and  1977-1978. 
Open  circles  = 1971-1972,  closed  circles  = 1972-1973,  open  squares  = 1977-1978, 
closed  square  (1)  = 1978  broods.  Sex  symbols  indicate  males  and  females  for 
1971-1972  and  1977-1978. 


midges  of  the  family  Chironomidae.  Adults  and  large  immatures 
tended  to  take  larger  prey,  such  as  syrphid  and  dolichopodid  flies. 
Of  12  prey  records,  immatures  were  found  with  3 chironomids,  (one 
identified  as  Orthocladiini  by  A.  R.  Soponis),  1 chaoborid  fly,  1 
syrphid  fly  (genus  Toxomerus  identified  by  H.  V.  Weems),  1 


1980] 


Richman  & Whitcomb — Lyssomanes  viridis 


131 


Figure  6.  Population  densities  of  Lyssomanes  viridis  (Walckenaer)  in  magnolia 
stand  at  Tall  Timbers  fall  1977  to  spring  1978.  Data  points  are  means  between 
surveys  made  both  inside  and  outside  the  stand,  during  the  last  half  of  each  month. 

encyrtid  wasp  and  1 aphid  (genus  Macrosiphon  identified  by  H.  A. 
Denmark).  Adult  males  were  collected  with  an  unknown  dipteran 
and  a salticid  spider  of  the  genus  Hentzia.  Adult  females  were 
collected  with  a dolichopodid  fly,  an  unknown  dipteran  and  a 
psocid. 

The  population  density  (Figure  6)  dropped  during  the  winter,  but 
rose  in  the  spring  nearly  to  that  of  the  previous  fall,  probably 
reflecting  inactivity  during  the  winter,  rather  than  a significant 
mortality.  The  population  drop  during  May  is  probably  a result  of 
the  death  of  adults.  Adults  were  only  found  during  the  spring  and 
early  summer. 

Some  adult  spiders  were  found  in  the  nest  ol  a mud  dauber  of  the 
genus  Trypoxylon  by  G.  B.  Edwards  at  Newnan’s  Lake,  Alachua 
County,  Florida.  A large  Trypoxylon  was  observed  during  June  at 
Tall  Timbers  and  a fresh  nest  was  found  on  the  underside  of  a 
magnolia  leaf.  The  nest  in  this  case  was  filled  with  Araneidae.  One 
adult  female  L.  viridis  was  collected  and  found  to  have  a large 
mirmithid  nematode  in  its  abdomen.  No  egg  parasites  were  seen. 

Complete  life  histories  have  been  published  for  several  salticids. 


132 


Psyche 


[Vol.  88 


notably  Philaeus  chrysops  Poda  (Bonnet  1933),  three  species  of 
Corythalia  (Crane  1948)  and  Phidippus  johnsoni  Peckham  and 
Peckham  (Jackson  1978).  Female  P.  chrysops  generally  had  seven 
molts  (six  instars)  before  maturity,  and  this  was  also  true  of  the 
three  species  of  Corythalia  observed  by  Crane  (1948).  Jackson 
(1978)  reported  6-9  molts  for  P.  johnsoni.  Thus,  the  life  cycle  of  L. 
viridis  seems  to  compare  well  with  those  of  other  salticids. 

Summary 

A population  of  the  salticid  spider  Lyssomanes  viridis  (Walcke- 
naer)  was  sampled  for  two  and  one  half  years  on  a stand  of 
Magnolia  grandiflora  L.  trees  in  North  Florida.  Mating  took  place 
in  May  and  adult  males  disappeared  by  mid-June.  Females  laid 
25-70  eggs  per  clutch  mostly  during  June.  These  hatched  from  June 
to  July  and  the  immatures  overwintered  in  middle  instars.  After 
temperatures  increased  in  the  spring  the  spiders  rapidly  developed 
to  adults.  Simple  linear  regression  of  the  carapace  widths  indicated 
that  this  species  has  a total  of  seven  instars  from  the  end  of  second 
postembryo  through  adult  female.  Males  may  have  one  fewer  instar. 
L.  viridis  feeds  primarily  on  Diptera  in  this  habitat. 


Acknowledgements 

We  would  especially  like  to  thank  Dr.  Bruce  Means  and  Ed  and 
Roy  Komarek  of  Tall  Timbers  Research  Station  for  their  help  and 
support.  Also  we  would  like  to  thank  Dr.  Robert  C.  Hemenway,  Jr., 
Dr.  G.  B.  Edwards,  and  Dr.  Barbara  Saffer  for  their  help  in 
collecting  specimens  in  the  magnolia  stand. 

Literature  Cited 


Bonnet 

1933.  Cycle  vital  de  Philaeus  chrysops  Poda  (Araneide,  Salticide).  Arch.  Zool. 
Exp.  Gen.  75:  129-44. 

Crane,  J. 

1948.  Comparative  biology  of  salticid  spiders  at  Rancho  Grande,  Venezuela. 
Part  I.  Systematics  and  life  history  in  Corythalia.  Zoologia  33:  1-38. 
Comparative  biology  of  salticid  spiders  at  Rancho  Grande,  Venezuela. 
Part  V.  Postembryonological  development  of  color  and  pattern.  Zoo- 
logica  35:  253-61. 


1950. 


1980] 


Richman  & Whitcomb — Lyssomanes  viridis 


133 


Galiano,  M.E. 

1976.  Comentarios  sobre  la  categoria  sistematica  del  taxon  “Lyssomanidae” 
(Araneae).  Rev.  Mus.  Argentino  Cienc.  Natur.  Entomol.  5:  59-70. 

Hagstrum,  D.W. 

1971.  Carapace  width  as  a tool  for  evaluating  the  rate  of  development  of 
spiders  in  the  laboratory  and  the  field.  Ann.  Entomol.  Soc.  Amer.  64: 
757-60. 

Jackson,  R.R. 

1978.  Life  history  of  Phidippus  johnsoni  (Araneae,  Salticidae).  J.  Arachnol.  6: 
1-29. 

Kaston,  B.J. 

1978.  How  to  know  the  spiders.  3rd  Ed.  Wm.  C.  Brown,  Dubeque.  272  p. 

Richman,  D.B. 

(in  press).  Epigamic  display  in  jumping  spiders  (Araneae,  Salticidae)  and  its  use  in 
systematics.  J.  Arach. 


THE  EMIGRATION  BEHAVIOR  OF  TWO  SPECIES  OF 
THE  GENUS  PHE1DOLE  (FORMICIDAE:  MYRMICINAE). 

By  Robert  droual1  and  Howard  Topoff2 


Introduction 

Colony  emigrations  are  common  among  ants  (Wilson  1971)  and 
occur  for  a diversity  of  reasons.  However,  except  for  the  legionary 
ants,  in  which  colony  emigrations  are  an  inherent  part  of  the 
foraging  ecology  (Wilson  1971),  and  species  which  inhabit  delicate 
and  easily  disturbed  nests  (Holldobler  and  Wilson  1977,  Moglich 
1979),  emigrations  are  thought  to  occur  infrequently.  Here  we 
present  evidence  that  two  species  of  the  genus  Pheidole,  P.  deser- 
torum  Wheeler  and  P.  hyatti  Emery,  emigrate  frequently  under 
environmentally  stable  conditions.  We  further  advance  the  hy- 
pothesis that  the  surplus  nests  resulting  from  these  emigrations, 
reduce  the  secondary  losses  which  occur  as  a consequence  of  the 
panic-alarm  defense  these  species  employ  against  army  ants  of  the 
genus  Neivamyrmex,  by  serving  as  temporary  shelters  and  centers 
for  colony  reorganization. 


Methods 

This  investigation  was  conducted  during  the  months  of  June,  July 
and  August,  1980,  at  two  different  study  sites.  One  site  was  an  oak- 
juniper  woodland  located  on  the  grounds  of  the  Southwestern 
Research  Station  of  the  American  Museum  of  Natural  History  near 
Portal,  Arizona  (elev.  1646  m).  The  other  site  was  a desert-grassland 
located  8 km  N.W.  of  Rodeo,  Hidalgo  Co.,  New  Mexico  (elev.  1250 
m).  In  both  habitats  a winter  (Dec.,  Jan.,  Feb.  and  March)  and  a 
summer  (July,  June  and  August)  rainy  season  occur.  On  the  oak- 
juniper  woodland  site  colonies  of  both  P.  desertorum  and  P.  hyatti 
were  located  and  marked;  on  the  desert-grassland  site  only  colonies 
of  P.  desertorum  were  located  and  marked. 

'Biology  Program,  City  College  of  C.U.N.Y.,  New  York,  N.Y.  10031 
Psychology  Department,  Hunter  College  of  C.U.N.Y.,  New  York,  N.Y.  10021  and 
Department  of  Entomology,  The  American  Museum  of  Natural  History,  New  York, 
N.Y.  10024 

Manuscript  received  by  the  editor  June  8,  1981 


135 


136 


Psyche 


[Vol.  88 


Table  1.  Emigration  characteristics  of  P.  hyatti. 


Colony 

Days 

Observed 

Number  of 
Emigrations 

Returns  to  a 
Former  Nest 

Distance  Between  First 
and  Last  Observed 
Nests  (m) 

H-Jnl4-1 

66 

6 

3 

4.2 

H-Jnl4-2 

66 

16 

1 1 

1.1 

H-Jnl4-3 

66 

8 

2 

5.4 

H-JnI5-l 

23 

1 

0 

1.5 

H-Jnl5-2 

63 

7 

3 

0.0 

H-Jnl7-1 

63 

0 

0 

- 

H-Jnl7-2 

63 

6 

2 

0.8 

H-Jnl8-1 

62 

10 

5 

3.2 

H-Jnl8-2 

55 

7 

2 

6.8 

H-Jnl9-1 

61 

7 

4 

0.0 

H-Jnl9-2 

57 

4 

1 

0.0 

H-Jnl9-3 

61 

6 

2 

1.5 

H-Jnl9-4 

60 

5 

2 

0.0 

H-Jnl9-5 

61 

7 

5 

2.0 

H-Jn21-1 

59 

3 

0 

2.6 

H-Jn21-2 

58 

6 

4 

1.9 

H-Jn21-3 

57 

4 

2 

0.0 

H-Jn21-4 

57 

6 

3 

3.0 

H-Jn23-1 

53 

2 

0 

0.8 

H-Jn24-1 

56 

7 

1 

0.8 

H-Jn26-1 

51 

5 

1 

2.5 

H-Jn26-2 

16 

2 

0 

1.1 

H-Jn27-1 

46 

8 

2 

0.4 

H-Jn28-1 

52 

4 

2 

2.4 

Total 

137 

57 

Colony  designations  were  based  on  the  species  (D  - desertorum, 
H - hyatti ),  the  date  when  the  colony  was  found  (Jn  - June,  J1  - July, 
A - August)  and  the  order  in  which  it  was  found  on  that  date.  For 
example,  H-Jnl8-2  is  the  second  P.  hyatti  colony  found  on  June  18. 

Activity  for  both  species  began  at  approximately  2000  hr  (MST) 
and  ceased  0500  hr.  To  determine  emigration  frequency  all  colonies 
were  inspected  at  least  once  daily  between  2000  and  2400  hr.  In 
order  to  avoid  disturbing  the  colony  any  prolonged  observations 
were  made  using  red  light.  About  two-thirds  of  the  emigrations  for 
each  species  were  documented  indirectly  when  a colony  occupying  a 
nest  one  night  was  found  at  another  nest  the  following  night.  A 
colony  was  assumed  to  be  occupying  a nest  if  10  foragers  were 


1981] 


Droual  & Topoff—  Genus  Pheidole 


137 


Table  2.  Emigration  characteristics  of  P.  desertorum. 


Colony 

Days 

Observed 

Number  of 
Emigrations 

Returns  to  a 
Former  Nest 

Distance  Between  First 
and  Last  Observed 
Nests  (m) 

D-JnlO-la* 

70 

4 

2 

0.5 

D-Jnll-la 

63 

7 

3 

4.8 

D-Jnl  l-2a 

68 

0 

0 

- 

D-Jnl2-la 

60 

4 

1 

15.6 

D-Jnl2-2a 

68 

1 

0 

2.5 

D-Jnl2-3a 

68 

2 

1 

0.0 

D-Jnl2-4a 

64 

6 

3 

3.0 

D-Jnl2-5a 

21 

1 

0 

4.2 

D-Jnl3-la 

57 

2 

1 

0.0 

D-Jnl4-la 

61 

3 

2 

0.0 

D-Jnl5-la 

65 

6 

3 

6.6 

D-Jnl5-2a 

62 

3 

1 

0.0 

D-Jnl6-1  b 

64 

4 

3 

0.0 

D-Jnl7-1  b 

57 

4 

2 

0.0 

D-Jnl7-2b 

63 

5 

3 

2.4 

D-Jnl8-la 

49 

3 

1 

3.1 

D-Jn20-1  b 

57 

8 

6 

0.0 

D-Jn25-la 

55 

7 

4 

1.5 

D-Jn28-la 

50 

5 

3 

3.7 

D-Jl  1-la 

42 

7 

3 

2.4 

D-Jl  13-lb 

37 

2 

1 

0.0 

D-Jl  13-2b 

37 

3 

1 

1.2 

D-Jl  15-lb 

33 

4 

1 

0.0 

D-Jl  30-lb 

20 

6 

3 

0.0 

D-Jl  30-2a 

19 

2 

1 

0.0 

D-A  1-la 

18 

2 

0 

4.0 

Total 

101 

49 

*a:  desert-grassland;  b:  oak-juniper  woodland. 


observed  leaving  the  nest  during  a 1 min  period.  If  this  criterion  was 
not  met,  or  if  there  was  some  other  reason  to  doubt  the  location  of 
the  colony,  a small  peanut  butter  bait  was  used  to  locate  the  colony. 
To  avoid  confusion  when  using  this  indirect  method,  an  attempt  was 
made  to  locate  and  mark  any  neighboring  conspecific  colonies.  The 
remainder  of  the  emigrations  were  observed  directly  when  an 
emigration  was  discovered  in  progress.  The  nests  were  marked  and 
the  distance  between  the  old  and  the  new  nests  measured. 

With  the  statistical  tests  employed  in  this  paper  probabilities  of 
.05  or  less  were  accepted  as  significant. 


138 


Psyche 


[Vol.  88 


Results 

Colonies  of  both  P.  desertorum  and  P.  hyatti  showed  consider- 
able variation  in  their  frequencies  of  emigration  (see  Tables  1 and  2). 
One  colony  of  each  species  (D-Jnll-2  and  H-Jnl7-1)  did  not 
emigrate  at  all,  while  one  P.  desertorum  colony  (D-Jn20-1)  emi- 
grated 8 times,  and  one  P.  hyatti  colony  (H-Jnl4-2)  emigrated  16 
times.  Despite  this  variability,  both  species  showed  a clear  tendency 
to  emigrate  frequently:  over  one-half  of  the  P.  desertorum  colonies 
emigrated  at  least  4 times,  and  over  one  half  of  the  P.  hyatti  colonies 
emigrated  at  least  6 times.  To  statistically  compare  the  emigration 
frequency  of  the  two  species,  the  percentage  days  for  which  an 
emigration  was  documented  was  calculated  for  each  colony,  and  the 
percentages  for  each  species  were  compared  using  the  Wilcoxon 
two-sample  test  (Sokal  and  Rohlf  1969).  No  significant  difference 
was  found  in  the  emigration  frequency  between  the  two  species 
(.1  > P > .05). 

This  similarity  between  species  in  emigration  frequency  can  also 
be  seen  if  the  frequency  of  time  interval  between  emigrations  is 
compared.  Figures  1 and  2 show  the  frequency  of  emigration 
interval  for  P.  hyatti  and  P.  desertorum,  respectively.  Both  distribu- 
tions are  strongly  skewed  to  the  right  with  a surprisingly  high 
number  of  emigrations  occurring  1 to  2 days  after  the  previous 
emigration.  No  significant  difference  was  found  in  the  frequency 
distribution  between  the  two  species  (Wilcoxon  two-sample  test: 
.4  > P > .2) 

The  daily  occurrence  of  emigrations  among  all  colonies  is  shown 
in  the  form  of  frequency  histograms  in  Figs.  3,  4 and  5.  The  upper 
line  in  the  graphs  outlines  the  number  of  colonies  which  were 
included  in  the  sample  size  each  night.  Excluded  from  the  sample 
were  colonies  which  were  raided  by  army  ants,  or  were  still  suffering 
from  the  effects  of  an  army  ant  raid  (see  Discussion).  Superimposed 
over  the  graphs  is  a bar  diagram  showing  the  daily  rainfall. 

A positive  correlation  was  found  to  exist  between  emigration 
frequency  and  daily  rainfall  in  all  three  cases  (Spearman  rank 
correlation  coefficient:  P.  hyatti : rs  = .28,  N = 66;  P.  desertorum  in 
oak-juniper  woodland  : rs  = .25,  N = 64;  P.  desertorum  in  desert- 
grassland  : rs  = .32,  N = 70).  The  effect  of  rainfall  on  emigration 
frequency  is  most  clearly  seen  in  P.  desertorum  in  the  desert- 
grassland  habitat  (Fig.  5).  During  the  29  days  before  the  first  heavy 


1981] 


Droual  & Topoff—  Genus  Pheidole 


139 


rainfall  on  July  9 only  three  emigrations  occurred,  but  within  9 days 
after  this  rainfall  29  emigrations  occurred.  During  this  9 day  period 
13  of  the  15  colonies  being  observed  on  this  site  emigrated  at  least 
once. 

The  emigration  distance  for  both  species  was  variable.  Mean 
emigration  distance  for  P.  hvatti  was  1.8  ± 1.0  m (N=137;  range  0.3 
— 4.9)  and  mean  emigration  distance  for  P.  desertorum  was  2.5  ± 
1.4  m (N=102;  range  0.4  — 6.9).  The  larger  emigration  distance  of 
P.  desertorum  over  that  of  P.  hyatti  correlates  with  the  larger  size  of 
this  species  (mean  length  of  P.  hyatti  minor  = 2.64  ± 0.04  mm, 
N=50;  mean  length  of  P.  desertorum  minor  = 3.14  ± 0.03  mm, 
N=57). 

Despite  the  high  emigration  frequency  of  both  species,  colonies  of 
neither  species  tended  to  move  far  from  the  nests  at  which  they  were 
first  discovered.  Tables  1 and  2 show  the  number  of  times  each 
colony  returned  to  a former  nest,  and  the  distance  between  the  first 
and  last  nests.  As  can  be  seen,  49%  of  P.  desertorum ’s  emigrations, 
and  42%  of  P.  hyatti" s emigrations  were  to  former  nest  sites,  and  1 1 
P.  desertorum  colonies  and  5 P.  hyatti  colonies  at  the  end  of  the 
study  were  at  the  nest  at  which  they  were  first  discovered.  This 
crisscrossing  pattern  of  emigrations  is  illustrated  for  three  colonies 
of  each  species  in  Figs.  6 and  7.  The  relative  location  of  the  nests 
reveal  a clumped  rather  than  a linear  arrangement  which  would  be 
expected  if  the  colony  were  emigrating  out  of  an  area.  The  dates  of 
nest  movements  for  each  colony  show  that  the  variability  of 
emigration  interval  within  each  colony  was  considerable.  This  can 
be  readily  seen  by  examining  the  ranges  of  emigration  intervals  for 
the  colonies  shown  in  Figs.  6 and  7:  for  the  P.  hyatti  colonies  the 
ranges  are,  H-Jnl4-2:  1-17  days;  H-Jnl4-3:  3-18  days;  H-Jnl8-1: 
1-8  days;  for  the  P.  desertorum  colonies  the  ranges  are  D-Jn25-1: 
1-19  days;  D-Jnl2-4:  1-21  days;  D-J130-1:  1-4  days. 

Because  P.  desertorum  and  P.  hyatti  emigrated  so  frequently 
about  33%  of  the  emigrations  of  both  species  were  discovered  in 
progress.  These  emigrations  were  readily  noticed  as  hundreds  to 
thousands  of  workers,  most  carrying  brood,  formed  a column 
connecting  the  old  nest  to  the  new  nest.  The  width  of  this  column  for 
P.  hyatti  was  about  3 cm,  while  for  P.  desertorum  the  column 
tended  to  be  wider  (on  one  occasion  reaching  a width  of  15  cm). 
Laboratory  experiments  have  revealed  that  P.  hyatti's  emigrations 


140 


Psyche 


[Vol.  88 


EMIGRATION  INTERVAL  (DAYS) 

Figure  1 . Frequency  of  the  time  interval  between  emigrations  for  Pheidole  hvatti. 

are  totally  organized  by  a substance  secreted  by  the  poison  gland 
(Droual  et  al.,  in  prep.)-  The  queen  of  both  species  moved  inde- 
pendently in  the  emigrations  although  she  was  usually  surrounded 
by  a retinue  consisting  mostly  of  minor  workers  (workers  of  the 
genus  Pheidole  are  dimorphic)  who  tugged  her  by  the  mandibles  or 
antennae  if  she  hesitated  en  route  to  the  new  nest.  During  June  and 
the  early  part  of  July  alates  were  frequently  seen  in  the  column  also 
moving  independently.  However,  on  one  occasion,  during  a P. 
desertorum  emigration,  workers  were  observed  carrying  some  of  the 
males. 

A number  of  phenomena  related  to  these  species’  high  emigration 
frequencies  were  observed.  One  colony  of  each  species  (D-Jn20-1 
and  H-Jnl9-2)  performed  what  we  call  an  aborted  emigration.  In 
these  cases  the  colony  was  observed  emigrating  to  a new  nest  but  on 
the  following  night  was  found  to  be  back  at  its  old  nest.  One  P. 
desertorum  colony  (D-Jn-25-1)  appeared  to  perform  two  emigra- 
tions in  one  night.  On  August  17  the  colony  was  observed  emigrat- 
ing from  nest  2 to  nest  1 (see  Fig.  6).  However  on  the  following  night 
the  colony  was  found  at  nest  3.  On  a number  of  occasions  an 


1981] 


Droual  & Topoff — Genus  Pheidole 


141 


EMIGRATION  INTERVAL  (DAYS) 


Figure  2.  Frequency  of  the  time  interval  between  emigrations  for  Pheidole 
desertorum. 


emigration  could  be  predicted  in  advance  by  the  colony’s  excavation 
activity  at  another  site.  For  example,  before  colony  D-Jnl2-2 
emigrated  to  its  second  nest  site  on  8/17,  workers  from  the  colony 
were  observed  excavating  at  the  site  on  8/4,  8/ 5,  8/7  and  8/ 10-8/ 17. 
However  two  colonies  of  both  P.  desertorum  and  P.  hyatti  were 
observed  excavating  at  sites  to  which  they  did  not  emigrate  even 
though  they  emigrated  later  to  other  nests. 


Discussion 

In  this  paper  we  have  shown  that  P.  desertorum  and  P.  hyatti 
emigrate  frequently  and  that  the  emigration  frequencies  of  the  two 
species  are  similar.  This  similarity  in  emigration  frequency  becomes 
even  more  marked  when  it  is  taken  into  account  that  most  of  P. 
desertorum ’s  emigrations  in  the  desert-grassland  occurred  after  the 
first  rainfall.  The  sharp  increase  in  emigration  activity  after  the  rain 
can  possibly  be  explained  by  the  affect  of  the  rainfall  upon  the  soil. 
Before  the  rains  began  the  soil  was  very  hard  and  compact,  but  after 


142 


Psyche 


[Vol.  88 


5 

o 


6/15  6/20  6/25  6/30  7/5  7/10  7/15  7/20  7/25  7/30  8/5  8/10  8/15 

DATE 

Figure  3.  Daily  occurrence  of  emigrations  for  Pheidole  hvatti.  Black  bars 
indicate  the  number  of  colonies  which  emigrated  each  night.  Upper  line  outlines  the 
number  colonies  included  in  the  sample  each  night.  Right  ordinate  indicates  rainfall 
for  the  superimposed  bar  diagram  showing  daily  rainfall. 


the  first  heavy  rainfall  the  soil  loosened  considerably.  This  un- 
doubtedly made  the  excavation  of  new  nests  by  the  desert-grassland 
dwelling  colonies  much  easier.  The  same  reasoning  can  be  applied  to 
explain  the  positive  correlation  between  emigration  frequency  and 
daily  rainfall  in  both  habitats.  However,  in  the  oak-juniper  wood- 
land, the  greater  amount  of  vegetation,  the  rockier  soil  and  the 
generally  moister  conditions  probably  account  for  the  relatively 
higher  emigration  activity  before  the  beginning  of  the  rainy  season 
in  this  habitat. 

The  need  to  perform  a colony  emigration  is  a contingency  almost 
all  species  of  ants  can  be  expected  to  face  (Wilson  1971).  However 
some  species  emigrate  more  than  others.  Among  the  legionary  ants, 
particularly  the  Ecitoninae  and  Dorylinae,  colony  emigrations  are 
an  integral  part  of  the  foraging  ecology  (Wilson  1971).  Oppor- 
tunistic nesters  such  as  Tapinoma  melanocephalum,  T.  sessile, 
Paratrechina  bourbonica  and  P.  longicornis  occupy  ready-made 
nests  such  as  the  tufts  of  dead  grass  and  hollow  plant  stems  which 


NO  OF  EMIGRATIONS  NO  OF  EMIGRATIONS 


1981] 


Droual  & Topoff— Genus  Pheidole 


143 


20-| 


15 


10- 


P DESERTORUM 

(OAK -JUNIPER  WOODLAND  HABITAT) 


4U4 


on 


j 


"U  L 


iLI 


i ■ ■ j ■ 


-30 


■2  0 


10 


6/15  6/20  6/25  6/30  7/5  7/10  7/15  7/20  7/25  7/30  8/5  8/10  8/15 

DATE 

Figure  4.  Daily  occurrence  of  emigrations  for  Pheidole  desertorum  in  oat 
miper  woodland  (See  Fig.  3). 


P.  DESERTORUM 

(DESERT- GRASSLAND  HABITAT) 


i . . ..  i j 

6/10  6/15  6/20  6/25  6/30  7/5  7/10  7/15  7/20  7/25  7/30  8/5  8/10  8/15 


DATE 

Figure  5.  Daily  occurrence  of  emigrations  for  Pheidole  desertorum  in  desert- 
grassland  (see  Fig.  3). 


RAINFALL  (CM)  ' RAINFALL  (CM) 


144 


H - Jn 14  - 2 


/N 


I METER 


H- Jnl4-3 


✓ N 


Psyche 


NEST  7 


[Vol.  88 


NEST 

MOVEMENT  DATE 

1- 2  6/15 

2- 3  7/2 

3- 1  7/7 

1-4  7/9 

4- 3  7/15 

3- 4  7/18 

4- 5  7/22 

5- 1  7/27 

1- 4  7/30 

4- 6  7/31 

6- 5  8/3 

5- 4  8/5 

4-6  8/6 

6- 2  8/11 

2- 1  8/13 

1-4  8/15 

4-6  8/17 


NEST 

MOVEMENT  DATE 

1-2  6/18 

2- 3  6/22 

3- 4  7/10 

4- 5  7/20 

5- 1  7/27 

'1-6  8/1 

6- 7  8/11 

7 4 8/14 


I METER 


H - Jn  18-1 


/ N 


i 1 

I METER 


NEST  4 


NEST  6 


NEST  2*^ 


NEST  I 


NEST 

MOVEMENT  DATE 

1- 2  7/1 

2- 3  7/5 

3- 4  7/11 

4- 2  7/17 

2- 3  7/21 

3- 5  7/29 

5- 2  8/4 

2- 1  8/5 

1-3  8/8 

3- 6  8/13 


Figure  6.  Patterns  of  emigrations  for  three  colonies  of  Pheidole  hyatti.  Dates  of 
nest  movements  are  shown  on  the  right.  *This  nest  had  two  entrances. 


are  short-lived.  When  these  nests  are  disturbed,  the  colonies  quickly 
organize  emigrations  to  other  such  nests  (Holldobler  and  Wilson 
1977).  Leptothorax  acervorum  in  oak-juniper  woodland,  construct 
delicate  nests  under  stones  which  can  be  easily  dislodged  by  large 
vertebrates,  and  are  prone  to  emigrate  when  their  nest  is  disturbed 
(Moglich  1979). 

Most  species  build  or  choose  nest  sites  which  are  longer-lived  and 
less  easily  disturbed  and  are  thought  to  emigrate  infrequently. 
Among  these  species  emigrations  can  be  due  to  a local  factor  such  as 


1981] 


Droual  & Topoff— Genus  Pheidole 


145 


D-Jn25-I 


NEST 

MOVEMENT  DATE 

1- 2  7/12 

2- 3  7/15 

3- 4  7/16 

4-  I 8/5 

1- 4  8/6 

4-2  8/15 

2- 3  8/17 


I METER 


NEST 

MOVEMENT 

DATE 

1-2 

7/10 

2-1 

7/31 

1 -3 

8/1 

3-4 

8/4 

4-2 

8/5 

2-4 

8/14 

D-JI  30-1 


/ N 


NEST  I NEST  4 


NEST 

MOVEMENT 

DATE 

1-2 

7/31 

2-3 

8/3 

3-1 

8/7 

1 -3 

8/11 

3-4 

8/15 

4-1 

8/16 

I METER 


Figure  7.  Patterns  of  emigrations  for  three  colonies  of  Pheidole  desertorum. 
Dates  of  nest  movements  are  shown  on  the  right. 


shading  (Brian  1956,  Carlson  and  Gentry  1973),  or  climatic  ad- 
versity such  as  drought  or  frost  (Brian  1952).  A colony  may  also  be 
forced  to  emigrate  because  of  some  biotic  factor  such  as  inter-  and 
intra-specific  competition  (Holldobler  1976,  Waloff  and  Blackith 
1962,  Brian  1952,  Brian  et.  al.  1965)  and  predation  (Gentry  1974). 

The  view  that  emigrations  occur  infrequently  among  most  ants 


146 


Psyche 


[Vol.  88 


was  recently  challenged  by  Smallwood  and  Culver  (1979).  These 
investigators  conducted  a study  in  which  they  found  that  Tapinoma 
sessile  and  Aphaenogaster  rudis  emigrated  frequently.  Their  study 
differs  from  ours  in  that  colonies  were  marked  and  rechecked  only 
after  intervals  of  11-21  days  and  no  attempt  was  made  to  follow  the 
behavior  of  individual  colonies.  Because  T.  sessile  and  A.  rudis 
choose  different  nesting  sites  and  have  different  life  styles  these 
investigators  deduced  that  emigrations  occur  more  frequently 
among  ants  than  had  been  previously  thought.  However,  the  fact 
that  T.  sessile,  as  mentioned  above,  is  an  opportunistic  nester  which 
is  expected  to  emigrate  frequently  weakens  their  argument. 

It  is  difficult  to  apply  any  of  the  known  or  previously  hypothe- 
sized causes  of  colony  emigrations  to  explain  the  frequent  emigra- 
tions of  P.  desertorum  and  P.  hyatti.  The  nests  of  both  species  are 
excavated  in  the  soil  to  a depth  of  30  to  40  cm  (based  on  excavations 
in  oak-juniper  woodland),  and  hence  are  not  easily  disturbed. 
Shading  is  obviously  not  a factor  in  the  desert-grassland,  and  is 
negligible  in  the  oak-juniper  woodland  where  the  canopy  is  not 
extensive.  Permanent  deterioration  of  the  nest  as  a cause  is  elimi- 
nated by  the  fact  that  colonies  return  to  former  nests.  Indeed  almost 
half  of  all  emigrations  for  both  species  resulted  in  a return  to  a 
former  nest.  This  fact,  and  the  patterns  of  emigrations  which  tended 
to  keep  a colony  in  the  same  area,  argue  against  any  hypothesis 
which  involves  a deterioration  of  some  local  condition  such  as  might 
be  due  to  interference  competition  or  to  a decrease  in  the  local  food 
supply. 

However,  in  one  instance  the  possibility  that  an  emigration  may 
have  been  the  result  of  intraspecific  competition  should  be  men- 
tioned. This  involved  colony  H-Jnl4-1  which  was  observed  emigra- 
ting a distance  of  4.9  m the  day  after  it  was  found.  This  distance  is 
considerably  larger  than  the  mean  emigration  distance  of  1.8  m 
found  for  P.  hyatti.  Four  days  later,  colony  H-Jnl9-1  was  dis- 
covered emigrating  2.3  m into  the  nest  vacated  by  colony  H-Jnl4-1. 
Although  large-scale  conflicts  between  these  species  were  never 
observed,  workers  will  attack  any  alien  workers  of  either  species 
discovered  near  their  nest.  Frequent  encounters  of  this  sort  may 
have  caused  colony  H-Jnl4-1  to  make  its  unusually  long  emigra- 
tion. 

In  discussing  the  causation  of  any  behavior  a distinction  should 
be  made  between  those  hypotheses  that  invoke  a proximate  cause 


1981] 


Droual  & Topoff— Genus  Pheidole 


147 


and  those  that  invoke  an  ultimate  cause  (Wilson  1971).  For 
example,  it  has  been  hypothesized  that  the  ultimate  cause  of  army 
ant  emigrations  is  to  prevent  a local  depletion  of  food  resources 
(Wilson  1971).  The  proximate  cause  of  these  emigrations,  at  least 
among  the  Ecitoninae,  was  discovered  to  be  recruitment  to  a new 
nest  under  periods  of  high  colony  arousal  due  to  brood  stimulation 
(Schneirla  1938).  However,  it  has  been  recently  shown  that  food 
supply  may  also  be  a proximate  factor  (Topoff  and  Mirenda  1980, 
Mirenda  and  Topoff  1980).  The  hypothesis  we  are  advancing  to 
explain  the  frequent  emigrations  of  P.  hyatti  and  P.  desertorum 
concerns  the  ultimate  cause  of  these  emigrations  although  both  the 
ultimate  and  proximate  causes  are  the  subject  of  further  investiga- 
tion by  us. 

Both  P.  desertorum  and  P.  hyatti , which  are  small  and  lack 
potent  stings,  are  easy  prey  for  army  ants  of  the  genus  Neivamyrmex. 
Mirenda  et.  al.  (1980)  found,  in  the  same  desert -grassland  site 
employed  in  this  study,  that  P.  desertorum  was  the  species  most 
frequently  raided  by  N.  nigrescens.  Our  own  observations  also  show 
that  both  P.  desertorum  and  P.  hyatti  are  heavily  preyed  upon  by 
members  of  the  genus  Neivamyrmex  (Tables  3 & 4).  Some  P. 
desertorum  colonies  were  raided  repeatedly  by  the  same  army  ant 
colony  which  entered  the  statary  phase  in  a nearby  bivouac.  On  two 
occasions  an  army  ant  colony  actually  bivouacked  in  the  evacuated 
nest  of  a P.  desertorum  colony.  One  P.  hyatti  colony  was  raided  by 
two  species  of  Neivamyrmex.  Of  these  colonies  only  five  appeared 
to  be  completely  eliminated  by  the  army  ants.  Part  of  the  reason  for 


Table  3.  Observed  army  ant  raids  on  colonies  of  P.  hyatti. 


Colony 

Dates  of  Raids 

Species  Raiding 

H-Jnl5-1 

7/7,  7/8* 

Neivamyrmex  nigrescens 

H-Jnl5-2 

7/8 

N.  nigrescens 

H-Jnl9-4 

7/28 

N.  texanus 

H-Jn21-3 

8/15 

N.  opacithorax 

H-Jn21-4 

8/17 

N.  nigrescens 

H-Jn21-5 

8/15 

N.  opacithorax 

8/18 

N.  nigrescens 

H-Jn23-1 

8/15* 

N.  opacithorax 

H-Jn26-1 

8/12 

N.  nigrescens 

H-Jn27-1 

8/12* 

N.  nigrescens 

Colony  was  not  seen  afterwards. 


148 


Psyche 


[Vol.  88 


this  is  the  panic-alarm  defense  employed  by  these  species  against  the 
army  ants. 

That  defense  behavior  in  ants  can  be  both  enemy  specific  and 
complex  was  established  with  the  discovery  of  the  alarm-recruit- 
ment defense  of  Pheidole  dentata  against  the  fire  ant  Solenopsis 
geminata  (Wilson  1975  and  1976).  Although  more  evidence  is 
necessary,  the  defense  behavior  of  P.  hyatti  and  P.  desertorum 
appears  to  be  both  enemy  specific  and  complex.  The  defense,  which 
begins  when  a Pheidole  forager  contacts  an  army  ant  and  runs  back 
into  the  nest  raising  an  alarm,  occurs  in  two  phases.  In  the  first,  or 
“milling”,  phase,  workers  carrying  brood  well  out  of  the  nest  but 
remain  in  close  contact  near  the  nest’s  entrance.  In  the  second,  or 
absconding,  phase,  the  workers  flee  from  the  nest.  P.  desertorum's 
flight  is  protean  in  nature  (Humphries  and  Driver  1970)  with 
workers  scattering  in  all  directions.  In  P.  hyatti  the  exodus  is  more 
organized  with  the  workers  fleeing  in  columns  which  appear  to 
follow  recently-laid  chemical  trails. 

After  evacuating  from  their  nest  the  fleeing  workers  tend  to 
concentrate  at  temporary  shelters  such  as  that  provided  by  leaf 
litter,  fallen  branches,  rotting  logs  and  tufts  of  grass.  Some  workers 
eventually  find  some  or  all  of  the  former  nests  and  begin  to  recruit 
other  workers  to  them.  After  the  raid  is  over  workers  will  also  start 
to  return  to  the  evacuated  nest.  In  this  manner  the  colony  becomes 
fragmented  with  various  proportions  of  the  colony  in  some  or  all  of 
the  available  nests.  The  colony  then  begins  the  process  of  reorgan- 
izing with  segments  in  one  nest  emigrating  to  join  segments  in 
another  nest  until  the  colony  becomes  reunited  in  one  nest.  Hence  it 
appears  that  the  surplus  nests  resulting  from  the  frequent  emigra- 
tions of  these  species  serve  a dual  purpose  after  an  evacuation:  they 
provide  shelter  and  centers  for  reorganization. 

After  a nest  evacuation,  finding  a place  of  suitable  moisture 
before  the  lethal  surface  temperatures  and  low  surface  humidity  of 
the  approaching  day  is  undoubtedly  of  vital  importance  for  these 
nocturnal  species.  This  problem  becomes  particularly  severe  in  the 
desert-grassland  where  the  lack  of  ground  cover  makes  nests 
excavated  in  the  ground  the  only  suitable  shelters.  Having  alternate 
nests  becomes  a necessity  when  an  army  ant  colony  bivouacs  in  the 
evacuated  nest.  The  hypothesis  we  are  proposing  then  is  that  the 
surplus  nests  which  result  from  the  emigrations  of  these  species 
increases  the  effectiveness  of  the  panic-alarm  defense  by  reducing 


1981] 


Droual  & Topoff— Genus  Pheidole 


149 


Table  4.  Observed  army  ant  raids  on  colonies  of  P.  desertorum. 


Colony 

Dates  of  Raids 

Species  Raiding 

H-Jnll-1 

8/ 1-8/5 

Neivamyrmex  nigrescens 

H-Jnll-2 

7/9 

N. 

nigrescens 

H-Jnl2-1 

8/4,  8/6,  8/8 

N. 

nigrescens 

H-Jnl2-4 

7/4 

N. 

nigrescens 

H-Jnl2-5 

7/2* 

N. 

nigrescens 

H-Jnl3-1 

7/12,  8/ 10* 

N. 

nigrescens 

H-Jnl4-1 

8/5,  8/6,  8/12-8/15 

N. 

nigrescens 

H-Jnl8-1 

7/11,  7/12,  7/16, 
7/25,  7/26,  7/28,  8/1 

N. 

nigrescens 

H-Jnl7-1 

8/18 

N. 

nigrescens 

H-Jn25-1 

8/28 

N. 

nigrescens 

H-Jl  4-1 

8/3,  8 / 5—8 / 7,  8/9 

N. 

nigrescens 

H-Jl  30-2 

8/13 

N. 

nigrescens 

H-Jl  15-1 

7/28 

N. 

texanus 

*Colony  not  seen  afterwards. 


the  secondary  losses  which  result  from  the  disorganization  which 
follows  the  defense.  If  this  hypothesis  proves  to  be  correct,  the 
possibility  that  the  frequent  emigrations  of  these  species  have 
evolved  to  serve  as  part  of  a defense  system  against  the  army  ants 
has  to  be  entertained. 

Acknowlegements 

This  research  was  supported  by  a N.I.M.H.  training  grant  (MH 
15341),  a Theodore  Roosevelt  Memorial  Fund  Grant,  PSY-CUNY 
Grant  13492  and  a NSF  Grant  BNS-  8004565.  We  thank  Dr.  G. 
Turkewitz  for  his  statistical  advice  and  Roy  R.  Snelling  for  identify- 
ing the  Pheidole  specimens. 

Literature  Cited 

Brian,  M.  V. 

1952.  The  structure  of  a dense  natural  ant  population.  J.  Anim.  Ecol.  21: 
12-24. 

1956.  Segregation  of  species  of  the  ant  genus  Myrmica.  J.  Anim.  Ecol.  25: 
319-337. 

Brian,  M.  V.,  J.  Hibble  and  D.  J.  Stradling. 

1965.  Ant  pattern  and  density  in  a Southern  English  heath.  J.  Anim.  Ecol.  34: 
545-555. 

Carlson,  D.  M.  and  J.  B.  Gentry. 

1973.  Effects  of  shading  on  the  migratory  behavior  of  the  Florida  harvester 
ant,  Pogonomyrmex  badius.  Ecology  54:  452-453. 


150 


Psyche 


[Vol.  88 


Gentry,  J.  B. 

1974.  Response  to  predation  by  colonies  of  the  Florida  harvester  ant,  Pogono- 
myrmex  badius.  Ecology  55:  1328-1338. 

Holldobler,  B. 

1976.  Recruitment  behavior,  home  range  orientation  and  territoriality  in 
harvester  ants,  Pogonomyrmex.  Behav.  Ecol.  Sociobiol.  1(1):  3-44. 

Holldobler,  B.  and  E.  O.  Wilson. 

1977.  The  number  of  queens:  an  important  trait  in  ant  evolution.  Natur- 
wissenshaften  64:  8-15. 

Mirenda,  J.  T.,  D.  G.  Eakins,  K.  Gravelle  and  H.  Topoff. 

1980.  Predatory  behavior  and  prey  selection  by  army  ants  in  a desert-grassland 
habitat.  Behav.  Ecol.  Sociobiol.  7:  119  127. 

Mirenda,  J.  T.  and  H.  Topoff. 

1980.  Nomadic  behavior  of  army  ants  in  a desert-grassland  habitat.  Behav. 
Ecol.  Sociobiol.  7:  129-135. 

Moglich,  M. 

1978.  Social  organization  of  nest  emigration  in  Leptothorax.  (Hym.  Form.). 
Ins.  Soc.  25(3):  205-225. 

SCHNEIRLA,  T.  C. 

1938.  A theory  of  army-ant  behavior  based  upon  the  analysis  of  activities  in  a 
representative  species.  J.  Comp.  Psychol.  25:  51-90. 

Smallwood,  J.  and  D.  C.  Culver. 

1979.  Colony  movements  of  some  North  American  ants.  J.  Anim.  Ecol.  48: 
373-382. 

SOKAL,  R.  R.  AND  F.  J.  ROHLF. 

1969.  Biometry.  W.  H.  Freeman  and  company,  San  Francisco,  xxi  + 776  p. 

Topoff,  H.  and  J.  Mirenda. 

1980.  Army  ants  of  the  move:  relation  between  food  availability  and  emigra- 
tion frequency  in  Neivamyrmex  nigrescens.  Science  207:  1099-1100. 

Waloff,  N.  and  R.  E.  Blackith. 

1962.  The  growth  and  distribution  of  the  mounds  of  Lasius  flavus  (Fabricius) 
(Hym:Formicidae)  in  Silwood  Park,  Berkshire.  J.  Anim.  Ecol.  31: 
421-437. 

Wilson,  E.  O. 

1971.  The  Insect  Societies.  Belknap,  Harvard  Univ.  Press,  Cambridge,  x + 548 
P- 

1975.  Enemy  specification  in  the  alarm-recruitment  system  of  an  ant.  Science 
190:  798-800. 

1976.  The  organization  of  colony  defense  in  the  ant  Pheidole  dentata  Mayr 
(Hymenoptera:  Formicidae).  Behav.  Ecol.  Sociobiol.  1:  63-81. 


STATARY  BEHAVIOR  IN  NOMADIC  COLONIES 
OF  ARMY  ANTS: 

THE  EFFECT  OF  OVERFEEDING 


By  Howard  Topoff,  Aron  Rothstein*,  Susan  Pujdak,  and 
Tina  Dahlstrom** 

Department  of  Psychology,  Hunter  College  of  CUNY,  New  York, 

N.Y.  10021,  and  The  American  Museum  of  Natural  History 

Introduction 

Nearctic  colonies  of  the  army  ant  Neivamyrmex  nigrescens 
Cresson  (subfamily  Ecitoninae)  exhibit  behavioral  cycles  consisting 
of  alternating  nomadic  and  statary  phases.  During  the  statary 
phase,  a colony  remains  at  the  same  nesting  site  and  forages 
irregularly  for  food.  The  nomadic  phase,  by  contrast,  is  character- 
ized by  night-long  raids  and  frequent  emigrations  to  new  bivouacs. 
According  to  Schneirla  (1957,  1958),  the  nomadic  phase  is  triggered 
by  stimulation  arising  from  newly-eclosed  callows,  and  is  subse- 
quently maintained  by  comparable  excitation  from  the  developing 
larvae.  Experimental  support  for  brood-stimulation  theory  stems 
from  studies  showing:  (1)  an  abrupt  reduction  in  nomadism  after 
removing  a portion  of  a larval  brood  (Schneirla  and  Brown,  1950); 
and  (2)  the  eclosion  of  a pupal  brood  (in  the  absence  of  newly- 
hatched  larvae)  is  indeed  sufficient  to  initiate  a nomadic  phase 
(Topoff  et  al.,  1980a).  Recent  studies  have  suggested,  however,  that 
brood  stimulation  may  in  turn  depend  on  the  degree  of  brood 
satiation.  Thus,  in  a preliminary  field  study  involving  food  aug- 
mentation, Mirenda  et  al.  (in  press)  was  able  to  halt  the  occurrence 
of  emigrations  during  a portion  of  the  nomadic  phase  in  colonies  of 
N.  nigrescens.  This  was  followed  by  more  prolonged  laboratory 
studies  (Topoff  and  Mirenda,  1980  a,b)  showing  that  the  frequency 
and  direction  of  nomadic  emigrations  are  indeed  influenced  by  the 
amount  and  location  of  food. 

This  paper  reports  findings  from  our  continued  studies  of  food 
augmentation  for  colonies  of  N.  nigrescens.  In  previous  studies, 
larval  stimulation  was  reduced  by  artificially  feeding  colonies  early 
in  the  nomadic  phase,  after  callow  eclosion.  Because  an  additional 

*Department  of  Biology,  City  College  of  CUNY,  New  York,  N.Y.  10031 

**Department  of  Zoology,  University  of  California,  Riverside,  CA  92521 


151 


152 


Psyche 


[Vol.  88 


goal  of  the  present  study  was  to  reduce  callow  stimulation  and 
thereby  delay  the  onset  of  the  nomadic  phase,  overfeeding  com- 
menced late  in  the  statary  period.  For  the  first  colony,  emigrations 
were  delayed  approximately  6 days.  In  the  second  overfed  colony, 
we  were  able  to  virtually  eliminate  the  nomadic  phase,  together  with 
all  associated  patterns  of  raiding  and  emigration  behavior. 

Methods  and  Procedures 

This  study  was  conducted  during  July  and  August,  1980,  in  a 
desert-grassland  habitat,  8 km  east  of  Portal,  Arizona.  The  site  was 
chosen  because  the  pavement-like  substrate  and  patchy  vegetation 
provided  us  with  an  excellent  view  of  the  ants’  raiding  and 
emigration  activities.  Surface  soil  temperatures  averaged  50°  C at 
1500  hr  (MST)  and  17° C at  0200  hr  throughout  the  summer 
(Mirenda  et  al.  1980).  As  a result  of  the  severe  daytime  temperatures 
and  aridity,  colonies  of  N.  nigrescens  were  usually  active  on  the 
surface  only  between  1900-0500  hr. 

Colonies  were  located  by  walking  through  the  study  area  with 
gasoline  lanterns  or  miner’s  cap  lamps.  Colony  no.  1 was  found  on 
July  11,  at  the  end  of  a nomadic  phase,  and  observed  nightly 
throughout  its  next  statary  period.  During  the  subsequent  nomadic 
phase,  the  colony  was  estimated  to  contain  approximately  80,000 
adults  and  50,000  larvae.  Colony  no.  2 was  collected  on  July  14, 
during  its  last  nomadic  emigration,  and  maintained  in  the  labora- 
tory (see  Topoff  et  al.  1980b  for  details  of  the  rearing  procedure) 
until  the  pupal  brood  was  fully  pigmented.  Prior  to  release  in  the 
field,  the  colony  was  culled  to  contain  4,000  adults  and  4,000  pupae. 
By  the  next  nomadic  phase,  approximately  4,500  larvae  were  also 
present  in  the  colony.  This  small  colony  size  was  chosen  for  two 
reasons:  (1)  to  increase  our  ability  to  appreciably  overfeed  the 
colony;  and  (2)  a laboratory  colony  of  comparable  size  had 
previously  been  released  without  food  supplementation,  as  part  of  a 
study  designed  to  show  that  laboratory  rearing  and  population 
reduction  do  not  alter  qualitative  aspects  of  nomadic  behavior.  This 
colony  could  therefore  serve  as  a convenient  control  for  our 
artificially-fed  colony. 

Food  for  both  experimental  colonies  consisted  of  adult  and  brood 
individuals  of  the  myrmicine  ant  Novomessor  cockerelli,  and 
workers  of  the  termite  genus  Gnathamitermes.  To  collect  Novo- 


1981]  Topoff,  Rothstein,  Pujdak,  & Dahlstrom — Army  Ants  153 


messor  brood  we  made  use  of  the  panic-alarm  behavior  that  this 
prey  species  exhibits  when  raided  by  army  ants.  Accordingly,  we 
released  several  hundred  adult  i V.  nigrescens  into  the  nest  entrance 
of  Novomessor,  and  aspirated  the  larvae  and  pupae  that  were 
removed  from  the  nest  by  their  own  adult  workers.  Whenever 
colonies  were  artificially  fed,  food  was  given  at  the  start  of  raiding  in 
the  evening,  while  the  column  was  within  2 m from  their  bivouac.  If 
columns  emerged  from  more  than  one  exit  hole,  booty  was  placed  at 
the  front  of  each  ant  column. 

Results 

The  raiding  and  emigration  activities  of  colony  no.  1 are  summar- 
ized in  Table  1.  This  colony  was  found  on  July  11,  late  in  its 
nomadic  phase.  It  became  statary  on  July  13,  after  settling  into  a 
kangaroo  rat  mound  ( Dipodomys  spectabilis).  On  the  third  statary 
night,  the  colony  conducted  a 3-m  long  shift  to  the  other  side  of  the 
mound.  A statary  shift  differs  from  a nomadic  emigration  in  that  it 
is  neither  preceded  nor  followed  by  raiding.  It  consists  instead  of  a 
single,  unbranched  column,  and  is  presumably  caused  by  a dis- 
turbance at  the  old  site.  For  the  next  13  statary  days,  the  colony 
remained  at  the  same  bivouac,  and  staged  either  brief  (1-3  hr)  or  no 
predatory  raids.  On  statary  day  17,  however,  the  colony  conducted  a 
longer  shift  to  an  adjacent  mound.  During  the  move,  we  observed 
that  all  of  the  pupae  were  deeply  pigmented,  and  that  a few  callows 
were  being  transported  by  mature  adults  to  the  new  site.  As  a result 
of  detecting  the  onset  of  eclosion,  we  started  artificial  feeding  of  the 
colony  on  the  next  night  (July  30),  and  continued  to  supply  food  for 
a total  of  six  consecutive  nights  (Table  1). 

Each  evening,  a basal  column  "appeared  on  the  surface  shortly 
after  sunset  (1800-1900  hr).  As  soon  as  the  ants  contacted  the  food, 
the  process  of  mass  recruitment  resulted  in  a sharp  increase  in  ant 
traffic  out  of  the  nest.  On  the  days  of  heaviest  feeding,  when  more 
than  30  g of  booty  were  provided,  the  army  ants  required  several 
hours  to  transport  it  back  to  the  bivouac.  The  colony  occasionally 
put  out  additional  raiding  columns  later  each  night,  but  all  captured 
booty  was  promptly  brought  back  to  the  original  bivouac,  and  no 
emigrations  occurred.  On  the  afternoon  of  August  5,  the  study  area 
received  14  mm  of  rainfall  between  1400-1550  hr.  The  overcast  sky, 
coupled  with  cool  temperatures  late  in  the  afternoon,  enabled  the 


154 


Psyche 


[Vol.  88 


Table  1.  Activity  schedule  for  Neivamyrmex  nigrescens  colony  no.  1 


Date 

Activity 

Raid  Emigrate 

Food 

Provided  (g) 

Proposed 

Phase-Day 

7/11 

+ 

+ 

- 

N-? 

7/12 

+ 

+ 

- 

N-? 

7/13 

+ 

- 

- 

S-l 

7/14 

+ 

- 

- 

S-2 

7/15 

- 

ss* 

- 

S-3 

7/16 

+ 

- 

- 

S-4 

7/17 

+ 

- 

- 

S-5 

7/18 

+ 

- 

- 

S-6 

7/19 

+ 

- 

- 

S-7 

7/20 

- 

- 

- 

S-8 

7/21 

- 

- 

- 

S-9 

7/22 

- 

- 

- 

S-10 

7/23 

+ 

- 

- 

S-l  1 

7/24 

- 

- 

- 

S-12 

7/25 

- 

- 

- 

S-l  3 

7/26 

- 

- 

- 

S-14 

7/27 

- 

- 

- 

S-l  5 

7/28 

- 

- 

- 

S-l  6 

7/29 

- 

ss* 

- 

S-l  7 

7/30 

+ 

- 

9.1 

S-18 

7/31 

+ 

- 

34.7 

N-l 

8/1 

+ 

- 

18.5 

N-2 

8/2 

+ 

- 

32.6 

N-3 

8/3 

+ 

- 

17.6 

N-4 

8/4 

+ 

- 

31.6 

N-5 

8/5 

+ 

+ 

- 

N-6 

8/6 

+ 

+ 

- 

N-7 

8/7 

+ 

+ 

- 

N-8 

8/8 

+ 

+ 

- 

N-9 

8/9 

+ 

- 

- 

N-10 

8/10 

+ 

-f 

- 

N-l  1 

8/11 

+ 

- . 

- 

N-12 

8/12 

+ 

+ 

- 

N-13 

8/13 

+ 

+ 

- 

N-14 

8/14 

+ 

- 

- 

S-l 

8/15 

+ 

- 

- 

S-2 

8/16 

+ 

- 

- 

S-3 

8/17 

+ 

- 

- 

SA 

8/18 

+ 

- 

- 

S-5 

8/19 

- 

- 

- 

S-6 

8/20 

+ 

- 

- 

S-7 

8/21 

- 

- 

S-8 

8/22 

• 

- 

- 

S-9 

statary  shift 


1981]  Topoff,  Rothstein,  Pujdak,  & Dahlstrom — Army  Ants  155 


colony  to  begin  raiding  earlier  than  usual.  Thus,  although  we 
arrived  at  the  site  by  1800  hr,  a long  (60  m)  emigration  was  already 
in  progress.  Given  the  large  size  of  the  colony,  we  decided  to 
terminate  food-augmentation.  The  colony  remained  nomadic  for 
the  next  nine  days,  during  which  time  it  emigrated  on  six  nights. 

In  order  to  determine  whether  we  had  been  successful  in  delaying 
the  onset  of  the  nomadic  phase,  three  independent  types  of  evidence 
were  analyzed:  (1)  phase  length:  (2)  callow  pigmentation;  and  (3) 
larval  size.  Collectively,  our  data  indicate  that  the  nomadic  phase 
was  indeed  delayed  for  4-8  days. 

Phase  Length:  Because  the  colony  was  temporally  anchored,  July 
13  can  be  considered  the  first  statary  day,  August  5 the  first  nomadic 
day.  Thus,  the  statary  interval  becomes  23  days  (Table  1).  Accord- 
ing to  Mirenda  and  Topoff  (1980),  the  range  of  statary-phase 
duration  for  N.  nigrescens  in  the  same  study  area  is  15-19  days,  with 
a modal  length  of  16  days.  This  suggests  that  the  minimum  delay  in 
nomadic  onset  for  our  colony  was  4 days.  If  we  use  instead  Mirenda 
and  Topoffs  modal  duration,  the  delay  is  calculated  as  7 days. 

Callow  pigmentation:  Newly  eclosed  callows  of  N.  nigrescens  are 
yellow  and  acquire  adult-like  pigmentation  between  7-12  days. 
Several  hundred  callows  were  collected  from  the  colony  during  its 
first  emigration  on  August  5,  and  compared  with  preserved  samples 
collected  daily  from  nomadic  colonies  in  previous  years.  Although 
this  form  of  visual  comparison  can  not  always  pinpoint  the  exact 
post-eclosion  day,  callows  from  the  artificially-fed  colony  were 
substantially  more  pigmented  than  those  typically  collected  from 
other  colonies  on  the  first  nomadic  night.  Our  comparison  between 
these  callows  and  previously  preserved  specimens  indicated  a post- 
eclosion  age  of  between  5-8  days. 

Larval  size:  Several  hundred  larvae  were  collected  by  aspiration 
from  the  first  emigration.  By  visual  inspection,  we  separated  the  10 
largest  and  10  smallest  larvae  and  measured  them  with  the  aid  of  a 
dissecting  microscope  fitted  with  an  ocular  micrometer.  The  mean 
length  of  the  large  group  was  4.0  mm  (range  = 3. 8-4. 2 mm),  as 
compared  with  a mean  of  1.5  mm  (range  = 1.3- 1.7  mm)  for  the 
small  group.  When  these  data  are  compared  with  Mirenda  and 
Topoffs  (1980)  graph  of  larval  growth  versus  nomadic  day,  they 
correspond  to  a range  of  nomadic  days  between  4-6. 

The  nightly  patterns  of  activity  for  colony  no.  2 and  for  the 
control  colony  are  summarized  in  Table  2.  For  this  small  colony,  we 


156 


Psyche 


[Vol.  88 


<u 

£ £ 

X 

t— 

fN 

m 

"T 

>n 

NO 

r- 

oo 

On 

o 

_ 

fN 

rn 

Tf 

>n 

£ ^ 

C/0 

c/o 

z 

Z 

Z 

z 

z 

Z 

Z 

z 

Z 

Z 

z 

Z 

z 

z 

z 

o 

c 

>> 

<d  o 

c 

o ■- 

o 

o 

>> 

c 

S 2 

t ao 

1 

1 

+ 

+ 

1 

+ 

+ 

1 

+ 

+ 

1 

+ 

+ 

1 

.+ 

+ 

+ 

jo 

a £ 

o 

o 

O <D 

d 

_> 

O 

c 

o 

o 

o 

< 

o 

o 

o 

o 

«n 

>n 

o 

o 

m 

o 

o 

o 

o 

>n 

in 

Tt 

Tf 

o 

o 

m 

m 

fN 

o 

m 

o 

o 

fN 

T3 

m 

X 

<n 

tt 

m 

Tt 

Tt 

Tf 

m 

Tt 

rn 

Tt 

Tt 

NO 

NO 

*->  3 

o 

O 

o 

o 

o 

o 

o 

o 

o 

o 

o 

o 

o 

o 

o 

o 

3 

U 

03  i- 

1 

m 

O 

o 

1 

m 

o 

i 

o 

o 

l 

o 

1 

o 

o 

1 

in 

o 

o 

1 

o 

o 

1 

>n 

CN 

Tf 

o 

m 

Tt 

in 

o 

fN 

Tt 

m 

Tt 

fN 

Q ° 

ON 

oo 

o 

oo 

On 

o 

ON 

oo 

oo 

oo 

r- 

oo 

OO 

ON 

o 

o 

fN 

fN 

C 

>, 

c 

<u 

l/~> 

NO 

r~~ 

OO 

ON 

o 

— 

fN 

ro 

Tf 

in 

NO 

r-~ 

oo 

ON 

o 

— - 

o 

fN 

fN 

fN 

fN 

fN 

fN 

fN 

fN 

fN 

fN 

<n 

m 

o 

Q 

r-~ 

r- 

r-~ 

t" 

t" 

r- 

r- 

r-~ 

r- 

r- 

I-" 

r- 

r- 

K 

Cj 

^3 

C/ 

c/i 

S*. 

,&c 

-o 

aj  _2 

c5 

5 

x 

Cj 

c/3  -o 

o 

Cl,  <U 

O 

S-15 

S-16 

N-l 

N-2 

N-3 

N-4 

N-5 

N-6 

N-7 

N-8 

N-9 

o 

Z 

Z 

fN 

z 

m 

Z 

Tt 

z 

S-l 

S-2 

S-3 

aj 

■a 

^ -C 

o 

£ 

0-  cL 

S' 



X 

CJ 

Q 

ao 

■ C 

' ” 

aj 

Cj 

-a  -a 

O OJ 

q 

in 

On 

fN 

rn 

in 

q 

q 

in 

aj 
C A 

o 

O -a 

ON 

1 

1 

r-" 

1 

1 

NO 

1 

d 

ON 

oo 

1 

1 

oo 

d 

On 

1 

1 

1 

cS 

O 

Pm  > 
O 

ab 

c 

3 

O 

> 

— 

cu 

•a 

<u 

X 

3 

-a 

<u 

X 

CJ 

03 

C\| 

d 

c 

C 

<D  o 
cj  •- 

^aj 

>3 

X 

o 

3 

o 

c q 

* 

* 

T3 

>3 

'> 

j-  i 

c 

_o 

o 

>» 

I)  2 

t.  ao 

3 6 

1 

1 

1 

1 

1 

1 

1 

1 

1 

1 

1 

1 

+ 

1 

1 

1 

1 

1 

<x> 

<U 

C/3 

V 

> 

3 

> 

3 

CJ 

CJ 

< 

U 

’> 

cj  <l> 

o 

<D 

X 

<u 

a 

3 

X 

fN 

< 

o 

o 

o 

m 

in 

o 

o 

in 

o 

<n 

o 

O 

X 

o 

fN 

ro 

Tf 

m 

o 

rn 

o 

m 

T3 

jy 

o ^2 

m 

fN 

fN 

fN 

m 

m 

m 

rn 

m 

"O 

CJ 

X 

w ‘5 

03  i- 

CN 

| 

| 

O 

1 

| 

| 

fN 

O 

1 

o 

fN 

fN 

l 

o 

i 

| 

o 

1 

fN 

1 

fN 

1 

fN 

1 

| 

1 

1 

OJ 

u 

CJ 

l_ 

/M 

03 

*-•  , 

in 

o 

o 

o 

o 

O 

<n 

o 

o 

<n 

o 

O 

a 

H 

5 o 

m 

fN 

o 

3- 

Tt 

Tt 

o 

m 

Q ° 

fN 

ON 

fN 

m 

fN 

o 

fN 

m 

o 

C/3 

fN 

fN 

o 

fN 

fN 

fN 

o 

fN 

fN 

fN 

fN 

c 

r3 

_o 

C/3 

Ct 

>3 

ab 

3 

<D 

O 

fN 

m 

Tj- 

m 

NO 

r~- 

oo 

ON 

O 

fN 

<n 

Tt 

l/~> 

£ 

3 

r- 

oo 

ON 

fN 

fN 

fN 

fN 

fN 

(N 

UJ 

Q 

oo 

oo 

OO 

oo 

OO 

OO 

oo 

oo 

oo 

OO 

oo 

oo 

OO 

OO 

OO 

OO 

oo 

OO 

OO 

* 

1981]  Topoff,  Rothstein,  Pujdak,  & Dahlstrom — Army  Ants  157 

were  able  to  monitor  the  time  of  onset  and  the  duration  of  each 
night’s  raid,  in  addition  to  the  emigration  frequency.  This  colony 
was  released  from  a laboratory  nest  at  1900  hr  on  August  7.  Because 
this  was  statary  day  15,  most  of  the  pupae  were  fully  pigmented.  The 
colony  promptly  moved  into  a subterranean  nest  beneath  a small 
hole  in  the  desert  floor.  The  first  raiding  column  appeared  shortly 
after  2200  hr,  at  which  time  9.0  g of  Novomessor  brood  and  termites 
were  placed  near  the  raiding  front.  The  army  ants  removed  the 
booty  in  less  than  1 hr,  after  which  all  surface  activity  ceased.  For 
the  next  seven  nights,  the  colony  was  either  not  active  on  the  surface 
or,  at  best,  conducted  brief  raids  (each  of  which  was  immediately 
followed  by  artificial  feeding)  but  no  emigrations.  On  August  15  we 
arrived  at  the  study  site  after  2200  hr,  and  found  the  colony 
emigrating  25  m to  the  NW.  Because  previously-collected  food  was 
being  transported  to  the  new  nest,  but  no  larvae  had  yet  appeared, 
we  considered  the  emigration  to  be  in  an  early  stage.  Accordingly, 
10.2  grams  of  booty  were  placed  near  the  emigration  column,  1 m 
from  the  old  bivouac.  This  resulted  in  recruitment  of  ants  both  from 
the  short  column  leading  to  the  old  nest,  and  from  the  longer 
emigration  column.  All  of  the  artificially-placed  food  was  taken 
back  to  the  old  nest,  and  the  emigration  was  aborted. 

On  August  19  (nomadic  day  1 1),  after  2 days  of  not  having  been 
fed,  the  colony  conducted  its  only  successful  emigration.  The  move 
took  the  colony  19  m to  the  N,  beneath  an  Ephedra  bush.  On 
August  25,  we  excavated  the  colony  and  forced  it  to  shift  its  statary 
bivouac.  This  procedure  verified  that  the  colony’s  larvae  had 
pupated.  Thus,  throughout  a nomadic  phase  lasting  14  days,  the 
colony  conducted  only  one  completed  emigration.  On  4 nomadic 
nights  no  raiding  occurred.  During  the  10  nights  in  which  raiding 
took  place,  the  median  time  for  raid  onset  was  2200  hr,  and  the 
median  duration  of  each  raid  was  1.5  hr. 

The  control  colony,  which  was  also  released  from  the  laboratory 
at  the  end  of  a statary  phase,  exhibited  more  typical  patterns  of 
nomadic  behavior  (Table  2).  During  a 15-day  nomadic  phase,  the 
colony  emigrated  on  1 1 nights.  Some  degree  of  raiding  took  place 
on  every  nomadic  night.  The  median  time  of  raid  onset  for  the 
control  colony  was  1850  hr,  and  the  median  duration  of  raiding  was 
9.7  hr. 


158 


Psyche 


[Vol.  88 


Discussion 

Much  of  the  discussion  generated  by  Schneirla’s  brood-stimula- 
tion theory  concerns  the  relative  degree  to  which  raiding  and 
emigrations  are  influenced  by  interactions  between  brood  and 
adults  (internal  processes),  and  by  external  environmental  factors. 
Theoretical  support  for  emphasizing  brood-related  processes  stems 
not  only  from  Schneirla’s  own  research  with  army  ants  (Schneirla, 
1957,  1958,  1971),  but  from  studies  of  other  social  insects  as  well. 
For  example,  honeybee  workers  can  collect  protein-rich  pollen  or 
carbohydrate-rich  nectar.  Louveaux  (1950)  found  that  the  amount 
of  pollen  collected  by  an  incipient  colony  is  small,  but  increases  as 
the  brood  population  increases.  In  another  experiment  (Louveaux, 
1958),  he  removed  the  colony  queen  from  a mature  colony  and 
found  that  pollen  collection  was  unaffected  until  many  of  the  larvae 
had  pupated.  Further  evidence  of  larval  stimulation  of  adult 
foraging  came  from  Fukuda,  1960  (in  Free,  1967),  who  showed  that 
foraging  workers  from  a recently-divided  colony  collected  very  little 
pollen  until  the  eggs  laid  by  the  new  queen  hatched  into  larvae. 
Finally,  Free  (1967)  demonstrated  that  adult  worker  foraging  was 
influenced  more  by  direct  access  to  the  brood  than  by  brood  odor 
alone.  Perhaps  most  significant  was  the  additional  finding  that 
artificially  feeding  a colony  with  pollen  resulted  in  a decrease  in 
pollen  collection  and  a corresponding  increase  in  nectar  collection. 

Although  Schneirla  was  primarily  concerned  with  the  role  of 
callow  and  larval  excitation,  he  did  recognize  the  role  of  food  as  an 
ecological  parameter.  Thus,  at  an  early  stage  of  his  field  research 
with  the  neotropical  genus  Eciton,  he  reported  (Schneirla,  1938) 
that  colonies  frequently  emigrate  along  the  heaviest  raiding  route  of 
that  day.  Nevertheless,  it  was  Rettenmeyer  (1963)  who  first  sug- 
gested that  the  location  and  amount  of  captured  food  might 
influence  not  only  the  path  of  colony  movements,  but  the  very 
tendency  to  emigrate  in  the  first  place.  The  idea  that  colony 
excitation  could  be  related  to  brood  satiation  has  received  empirical 
support  from  Free’s  (1967)  study  of  honeybees  and  from  related 
research  with  the  myrmicine  ant  genus  Myrmica  (Brian,  1957,  1962; 
Brian  and  Abbott,  1977;  Brian  and  Hibble,  1963).  It  was  therefore 
significant  that  by  the  time  of  Schneirla’s  last  field  study,  concerning 
emigration  behavior  in  the  paleotropical  army  ant  genus  Aenictus, 
he  conceded  that  short-term  variations  in  colony  excitation  may 


1981]  Topoff,  Rothstein,  Pujdak,  & Dahlstrom — Army  Ants  159 


depend  upon  the  “alimentary  condition  prevalent  in  the  brood” 
(Schneirla  and  Reyes,  1969),  and  that  emigrations  are  likely  to  begin 
soon  after  food  has  run  low. 

In  a recent  series  of  field  and  laboratory  studies  of  nomadic 
behavior  in  nearctic  colonies  of  N.  nigrescens  (Topoff  and  Mirenda, 
1980  a,b;  Mirenda  et  al.,  in  press),  we  demonstrated:  (1)  that  the 
location  of  booty  clearly  influences  the  direction  of  raiding  and 
therefore  of  emigrations;  and  (2)  that  artificially-fed  colonies  exhibit 
a lower  frequency  of  emigrations.  The  present  study  differs  from 
these  in  that  food  augmentation  began  late  in  the  statary  phase, 
before  most  of  the  callow  population  had  eclosed.  In  addition  to 
delaying  the  onset  of  the  nomadic  phase  by  reducing  excitation  from 
newly-eclosed  callows  (Topoff  et  ah,  1980a),  this  was  our  first 
attempt  to  eliminate  emigrations  through  a complete  nomadic  phase 
in  the  field. 

During  the  six  days  of  food  augmentation  for  colony  no.  1,  we 
provided  a total  of  144  g of  booty.  Since  the  colony  generated  few 
additional  raiding  columns,  the  artificially-administered  booty 
represents  over  90%  of  the  colony’s  total  food  intake  for  that  period. 
According  to  Mirenda  et  ah  (1980),  colonies  of  N.  nigrescens  gather 
approximately  0.4  mg  of  booty/ larva/ nomadic  night.  Thus,  on  the 
average,  we  provided  colony  no.  1 each  night  with  an  amount  of 
food  that  would  be  collected  by  a colony  containing  about  60,000 
larvae.  Although  our  estimate  of  colony  size  contains  an  error  of  ± 
20%,  we  can  be  reasonably  certain  of  having  provided  this  colony 
with  about  1.2  times  the  amount  of  food  it  would  normally  gather. 
Although  the  large  size  of  this  colony  dictated  that  we  could  no 
longer  supplement  its  food  to  the  same  degree  throughout  the 
remainder  of  the  nomadic  phase,  the  evidence  from  phase  length, 
callow  pigmentation,  and  larval  size  supports  the  conclusion  that 
the  onset  of  the  nomadic  phase  was  delayed  for  4-8  days. 

For  colony  no.  2,  which  was  considerably  smaller  and  more 
precisely  counted,  intensive  overfeeding  was  more  feasible.  On  the 
average,  8.8  g of  booty  were  provided  on  food-supplemented  nights. 
This  is  more  than  5 times  the  amount  of  food  that  a colony  of  this 
size  would  collect  in  the  field.  In  view  of  this  feeding  regime,  it  is  not 
surprising  that  the  colony  conducted  only  one  completed  emigration 
throughout  its  14-day  nomadic  phase.  We  must  emphasize,  how- 
ever, that  a reduction  of  the  frequency  of  nomadic  emigrations  is  by 
itself  not  sufficient  to  infer  a relationship  between  food  supply  and 


160 


Psyche 


[Vol.  88 


colony  arousal.  In  all  of  our  laboratory  and  field  studies,  we  always 
placed  food  near  raiding  fronts,  within  a few  meters  of  the  bivouac. 
In  most  cases,  the  army  ants  established  few  or  even  no  additional 
raid  columns  beyond  the  artificial  feeding  site.  Thus,  our  feeding 
procedure  reduces  the  ability  of  the  ants  to  locate  a suitable  nesting 
site,  which  is  a prerequisite  for  a successful  emigration  (Mirenda  et 
al.,  in  press).  The  case  for  a relationship  between  food  supply  and 
colony  arousal  is  made  considerably  stronger  by  considering,  in 
addition  to  emigration  frequency,  the  temporal  aspects  of  the  ants’ 
raiding  behavior.  Colony  no.  2 conducted  no  raids  on  4 nomadic 
nights.  By  comparison,  the  complete  absence  of  raiding  (on 
stormless  nights)  for  a nomadic  colony  of  N.  nigrescens  has  never 
been  reported,  although  it  is  a common  occurrence  for  statary 
colonies.  Finally,  when  we  include  the  data  on  raid  onset  and 
duration  for  colony  no.  2,  we  conclude  that  overfeeding  can 
effectively  shift  the  level  of  overall  colony  activity  from  a nomadic 
to  a statary  condition. 


Acknowledgments 

The  base  of  operations  for  this  study  was  the  Southwestern 
Research  Station  of  The  American  Museum  of  Natural  History. 
This  research  was  supported  by  NSF  Grant  BNS-8004565,  and  by 
PSC-CUNY  Grant  13492. 

Literature  Cited 

Brian,  M.  V. 

1957.  Food  distribution  and  larval  size  in  cultures  of  the  ant  Myrmica  rubra  L. 
Physiol.  Comp.  Oecol.  4:  329-345. 

Brian,  M.  V.  and  A.  Abbott 

1977.  The  control  of  food  flow  in  a society  of  the  ant  Myrmica  rubra  L.  Anim. 
Behav.  25:  1047-1055. 

Brian,  M.  V.  and  J.  Hibble 

1963.  Larval  size  and  influence  of  the  queen  on  growth  in  Myrmica.  Insectes 
Soc.  10:  71-81. 

Free,  J.  B. 

1967.  Factors  determining  the  collection  of  pollen  by  honeybee  foragers. 
Anim.  Behav.  15:  134  144. 

Louveaux,  J. 

1950.  Observations  sur  le  determinisme  de  la  recolte  du  pollen  par  les  colonies 
d’abeilles.  C.R.  Acad.  Sci.,  Paris  231:  921-922. 

1958.  Recherches  sur  la  recolte  du  pollen  par  les  abeilles  ( Apis  mellifica  L.) 
Ann.  Abeille  1:  1 13-188,  197-221. 


1981]  Topoff,  Rothstein,  Pujdak,  & Dahlstrom — Army  Ants  161 


Mirenda,  J.  and  H.  Topoff 

1980.  Nomadic  hehavior  by  ants  in  a desert-grassland  habitat.  Behav.  Ecol. 
Sociobiol.  7:  129-135. 

Mirenda,  J.,  D.  Eakins,  K.  Gravelle,  and  H.  Topoff 

1980.  Predatory  behavior  and  prey  selection  by  army  ants  in  a desert-grassland 
habitat.  Behav.  Ecol.  Sociobiol.  7:  119-127. 

Mirenda,  J.,  D.  Eakins,  and  H.  Topoff 
in  press.  Relationship  between  raiding  and  emigration  in  the  nearctic  army  ant 
Neivamyrmex  nigrescens.  Insectes  Soc. 

Rettenmeyer,  C.  W. 

1963.  Behavioral  studies  of  army  ants.  Univ.  Kansas.  Sci.  Bull.  44:  281-465. 

Schneirla,  T.  C. 

1938.  A theory  of  army  ant  behavior  based  on  the  analysis  of  activities  in  a 
representative  species.  J.  Comp.  Psychol.  25:  51-90. 

1957.  Theoretical  considerations  of  cyclic  processes  in  doryline  ants.  Proc. 
Amer.  Phil.  Soc.  101:  106-133. 

1958.  The  behavior  and  biology  of  certain  nearctic  army  ants:  last  part  of  the 
functional  season,  southeastern  Arizona.  Insectes  Soc.  5:  215-255. 

1971.  Army  ants:  a study  in  social  organization.  Freeman,  Calif.  349  p. 

Schneirla,  T.  C.  and  R.  Z.  Brown 

1950.  Army-ant  life  and  behavior  under  dry-season  conditions.  4.  Further 
investigations  of  the  cyclic  processes  in  behavioral  and  reporductive 
functions.  Bull.  Amer.  Mus.  Nat.  Hist.  95:  265-353. 

Schneirla,  T.  C.  and  A.  Y.  Reyes 

1969.  Emigrations  and  related  behavior  in  two  surface-adapted  species  of  the 
Old  World  doryline  ant,  Aenictus.  Anim.  Behav.  17:  87-103. 

Topoff,  H.  and  J.  Mirenda 

1980a.  Army  ants  on  the  move:  relationship  between  food  supply  and  emigra- 
tion frequency.  Science  207:  1099-1100. 

1980b.  Army  ants  do  not  eat  and  run:  relationship  food  supply  and  emigration 
behaviour  in  Neivamyrmex  nigrescens.  Anim.  Behav.  28:  1040  1045. 

Topoff,  H.,  J.  Mirenda,  R.  Droual,  and  S.  Herrick 

1980a.  Onset  of  the  nomadic  phase  in  the  army  ant  Neivamyrmex  nigrescens: 
distinguishing  between  callow  and  larval  excitation  by  brood  substitu- 
tion. Insectes  Soc.  27:  175-179. 

1980b.  Behavioural  ecology  of  mass  recruitment  in  the  army  ant  Neivamyrmex 
nigrescens.  Anim.  Behav.  28:  779  789. 


LIFE  HISTORY  OF  ANTAEOTR1CHA  SP. 

(LEPIDOPTERA:  OECOPHORIDAE:  STENOMATINAE) 

IN  PANAMA* 

By  Annette  Aiello 
Smithsonian  Tropical  Research  Institute 
P.  O.  Box  2072,  Balboa,  Panama 

The  subfamily  Stenomatinae  (Oecophoridae)  is  a New  World 
microlepidopteran  group  of  approximately  35  genera  and  more 
than  1200  species.  Its  range  is  from  the  United  States  through 
Argentina;  South  America  is  especially  rich  in  species.  Little  is 
known  of  the  biology  of  these  moths,  but  those  that  have  been 
studied  include  leaf  miners,  stem  borers,  and  seed  eaters.  The  genus 
Antaeotricha  Zeller,  of  similar  range,  comprises  more  than  400 
species,  many  of  which  are  leaf  tiers. 

Three  individuals  of  Antaeotricha  sp.  near  fractilinea  (Walsing- 
ham)  (Figure  1)  were  reared  from  larvae  collected  29  March  through 
4 April  1980  on  Barro  Colorado  Island  (BCI),  Panama.  Two 
additional  individuals  were  preserved,  one  in  its  final  instar,  the 
other  as  a pupa. 

The  larvae  had  constructed  tubes  (Figure  2)  of  silk,  frass,  and  cast 
head  capsules,  on  the  undersides  of  the  leaves  of  Mascagnia  nervosa 
(Malpighiaceae). 

At  the  time  of  collection  two  of  the  three  larvae  were  in  their  final 
instar  and  these  pupated  five  days  later.  The  third  individual, 
probably  a first  instar  judging  from  its  small  size  and  tiny  tube, 
molted  the  day  after  collection.  Due  to  the  uncertainty  regarding 
instar  number,  letters  instead  of  numbers  are  used  to  refer  to  instars. 

Mascagnia  nervosa  is  a liana  which  grows  into  the  canopy  of  the 
BCI  forest.  Seedlings  are  found  frequently  around  the  edges  of 
clearings  and  in  tree  falls.  Antaeotricha  larvae  were  common  on  the 
older  leaves  of  plants  10-30  cm  tall  and  bearing  three  to  eight  leaves 
each;  some  leaves  supported  as  many  as  four  larval  tubes,  although 
one  or  two  were  most  common.  Possibly  Antaeotricha  attacks 
leaves  of  this  plant  in  the  forest  canopy  as  well. 

Head  capsule  widths  (Table  1)  ranged  from  0.18  mm  (instar  A)  to 
1.38  mm  (final  instar).  Instar  durations  for  the  larva  collected  as 
instar  A were:  5(B),  4(C),  4(D),  3(E),  5(F),  and  8(G)  days. 

♦Manuscript  received  by  the  editor  September  14,  1981. 


163 


164 


Psyche 


[Vol.  88 


Figures  13.  Antaeotricha  sp.  (Rearing  lot  80-30).  1.  Adult  (individual  no.  4). 

Scale  = 3 mm.  2.  Silk  and  frass  tubes  on  underside  of  leaf  of  Mascagnia  nervosa. 
Scale  = 7 mm.  3.  Final  instar  larva  (individual  no.  1)  reaching  out  of  its  tube. 
Scale  = 2 mm. 


Early  instars  (A,  B)  had  pale-colored  heads  and  green  bodies. 
They  constructed  hard  tubes  of  silk  and  frass,  0.5  mm  in  diameter, 
on  the  undersides  of  leaves.  Most  tubes  were  initiated  along  major 
leaf  veins,  a position  which  may  offer  protection  to  young  larvae. 
The  initial  tube  was  dense  and  the  larva  inside  could  not  be  seen 
through  its  wall. 

Subsequent  instars  extended  the  tube,  each  tube  addition  being 
wider  and  longer  than  previous  portions. 

Larvae  did  not  leave  their  tubes  to  feed;  they  reached  out  (Figure 
3)  and  scraped  cells  from  the  leaf  surface. 

Commencing  with  instar  C,  larvae  had  dark  heads  and  a pink 
mesothorax.  The  remainder  of  the  body  was  green  as  before. 
Intermediate  instars  (C,  D)  continued  construction  of  the  hard  tube, 
and  also  fed  by  scraping  the  leaf  surface  immediately  in  front  of  the 
tube  opening. 

Later  instars  (E-G)  ate  whole  leaf.  Portions  of  the  tube, 


1981] 


Aiello — Life  History  of  Antaeotricha 


165 


Table  1.  Summary  of  variables  for  seven  instars  of  individual  no.  4 of  Rearing 
lot  80-30.  (P  = pale,  pink;  D = dark;  G = green;  S = skeletonizer,  soft;  H = hard;  W = 
eats  whole  leaf). 


A 

B 

C 

IN  STAR 
D 

E 

F 

G 

Instar  duration  (days) 

? 

5 

4 

4 

3 

5 

8 

Head  capsule  width  (mm) 

0.18 

0.25 

0.35 

0.48 

0.63 

0.83 

1.38 

Head  capsule  color 

P 

P 

D 

D 

D 

D 

D 

Mesothorax  color 

G 

G 

P 

P 

P 

P 

P 

Feeding  habit 

S 

S 

S 

S 

W 

W 

W 

Tube  consistency 

H 

H 

H 

H 

S 

S 

S 

constructed  by  these  instars,  were  soft  and  less  dense  than  previous 
sections.  The  apical  2-3  cm  of  these  tubes  were  extremely  diffuse 
and  the  larvae  inside  could  be  seen  clearly.  Having  consumed  all 
nearby  leaf,  hungry  larvae  dismantled  this  diffuse  portion  and 
shifted  it  laterally  until  additional  leaf  surface  was  located,  some- 
times on  other  leaves  of  the  same  plant. 

The  final  instar  larva  (G)  (Figure  4)  was  about  7 mm  long,  lacked 
secondary  setae;  had  prolegs  on  abdominal  segments  3-6  and  10; 
crochets  uniordinal,  arranged  in  a circle;  prespiracular  wart  of 
prothorax  long,  curving  part  way  around  spiracle,  and  bearing  three 
setae;  mesothorax  with  a single  seta  on  tubercle  pi;  abdomen  with 
setae  alpha  and  beta  widely  separated,  setae  eta  and  kappa  adjacent, 
seta  beta  on  ninth  segment  placed  higher  up  than  alpha,  and  setae 
beta  on  ninth  segment  the  same  distance  apart  from  one  another 
across  the  dorsum  as  on  previous  abdominal  segments;  head  with 
front  extending  about  one-half  the  way  to  the  vertex.  Because  the 
two  sides  of  the  mesothorax  bore  slightly  different  setation,  on  the 
specimen  studied,  both  are  illustrated  in  the  setal  maps  (Figure  4). 
For  ease  of  comparison,  the  map  of  the  right  mesothorax  is  shown 
in  mirror  image. 

The  day  before  pupation,  final  instar  larvae  abandoned  their 
tubes  and  dropped  to  the  floor  of  the  cage.  No  cocoon  was 
constructed. 

The  pupa  (Figure  5)  was  ovate  in  outline,  5 mm  long,  2 mm  wide; 
with  wings  extending  slightly  beyond  caudal  margin  of  fourth 
abdominal  segment;  apical  1.3  mm  of  antennae  adjacent  on  the 
meson  except  at  their  extreme  tips,  and  bearing  a distinctive  raised 


166 


Psyche 


[Vol.  88 


Figure  4.  Antaeotricha  sp.  Final  instar  head  capsule  (front  and  lateral  views), 
and  setal  maps.  P = prothorax,  MS  = mesothorax,  L = left,  R = right,  MT  = 
metathorax,  A = abdomen,  S = suranal  plate.  Map  of  right  mesothorax  is  shown  in 
mirror  image.  Scale  for  head  capsule  = 0.5  mm. 


1981] 


Aiello — Life  History  of  Antaeotricha 


167 


Figure  5.  Antaeotricha  sp.  Pupa  (ventral,  lateral,  and  dorsal  views).  Scale  = 1 
mm. 


structure  on  the  scapes;  mesothoracic  legs  and  maxillae  ending  in 
the  “V”  formed  by  the  meeting  of  the  antennae;  prothoracic  legs 
slightly  shorter  than  mesothoracic  legs;  labial  palpi  evident  as  tiny 
triangle  between  maxillae  bases;  fronto-clypeal  and  epicranial 
sutures  distinct;  abdominal  segments  1-4  longer  than  others; 
spiracles  evident  on  abdominal  segments  2-8;  cremaster  of  8 weak 
setae. 

Pupation  lasted  10,  11,  and  12  days  for  the  three  individuals 
reared.  Total  development  time,  for  the  individual  collected  as 
instar  A,  was  40  days.  Allowing  an  additional  three  days  for  instar 
A,  and  four  days  for  egg  maturation,  actual  development  time  was 
probably  close  to  47  days. 

Spread  adults,  pointed  head  capsules  and  pupal  skins,  and  a larva 
and  pupa  in  alcohol  are  in  the  collection  of  the  author,  deposited  in 
the  National  Museum  of  Natural  History,  and  are  labelled  as 
Rearing  lot  80-30. 

I would  like  to  thank  the  Smithsonian  Tropical  Research 


168 


Psyche 


[Vol.  88 


Institute,  Panama,  for  use  of  their  facilities,  J.  B.  Heppner  (SI)  for 
identification  of  the  moth,  and  R.  Silberglied  for  reading  the 
manuscript. 


Reference 

Watson,  A.  and  P.  E.  S.  Whalley 

1975.  The  Dictionary  of  Butterflies  and  Moths  in  Color,  xiv  + 296  pp.,  405  figs. 


POLISTES  GALUCUS  IN  MASSACHUSETTS 
(HYMENOPTERA:  VESPIDAE)* 

By  Mary  A.  Hathaway 
Museum  of  Comparative  Zoology, 

Harvard  University 

Cambridge,  Massachusetts  02138  USA 
Introduction 

Polistes  gallicus  (Linnaeus),  a common  and  widespread  paper 
wasp  in  the  palearctic  region,  has  been  introduced  into  the  United 
States  in  the  Boston,  Massachusets,  area.  During  1981  specimens 
were  collected  in  Cambridge,  Somerville,  Belmont,  and  Newton, 
Massachusetts.  P.  gallicus  was  also  collected  in  Cambridge  in  1980, 
but  was  not  seen  in  Belmont  that  year  (R.  J.  McGinley,  personal 
communication).  Species  identification  was  verified  by  Dr.  Arnold 
S.  Menke  of  the  Systematic  Entomology  Laboratory  of  the  United 
States  Department  of  Agriculture.  Presumably  P.  gallicus  was  only 
recently  introduced;  otherwise  it  would  surely  have  been  reported 
before  this.  It  is  a brightly  colored  and  conspicuous  wasp. 

The  purpose  of  this  paper  is  to  report  the  introduction  of  Polistes 
gallicus.  The  biology  of  the  species  in  the  Old  World  is  reviewed 
briefly,  and  some  observations  of  the  wasp  in  Massachusetts  are 
reported.  Information  on  how  to  recognize  P.  gallicus  is  also 
included. 


Biology  of  Polistes  gallicus  (Linnaeus) 

Polistes  gallicus  is  ubiquitous  in  the  palearctic  region,  especially 
in  the  south.  It  is  the  most  common  Polistes  in  Spain  (Giner  Mari, 
1945).  The  species’  range  extends  north  to  Paris,  but  gallicus 
becomes  rare  in  far  northern  France.  It  exists  in  warmer  parts  of 
Belgium  and  Germany,  but  does  not  occur  in  England,  Denmark,  or 
Scandinavia  (Guiglia,  1972).  Spradbury  (1973)  states  that  occasion- 
ally Polistes  are  introduced  into  the  British  Isles,  but  for  some 
reason  the  genus  is  not  able  to  sustain  itself  there.  To  the  south,  the 
range  of  P.  gallicus  includes  northern  Africa,  where  the  species  is 
known  from  desert  oases  (Richards,  1953),  and  extends  east  through 
Israel  and  Iran.  In  Asia  P.  gallicus  has  been  collected  in  southern 

♦Manuscript  received  by  the  editor  September  29,  1981. 


169 


170 


Psyche 


[Vol.  88 


U.S.S.R.  and  throughout  China,  east  to  the  Pacific  coast  (Guiglia, 
1972;  Yoshikawa,  1962).  Generally,  the  species  is  found  in  warmer 
and  dryer  localities  within  its  range.  It  is  not  common  above  1000  m 
elevation,  although  in  southern  Spain  specimens  have  been  collected 
above  2000  m (Guiglia,  1972). 

The  biology  of  P.  gallicus  varies  considerably  between  the 
climatic  extremes  of  the  area  it  inhabits.  North  of  the  Alps,  nests  are 
built  in  enclosed  places,  such  as  metal  containers  and  gutter  pipes. 
This  type  of  nest  has  also  been  reported  by  Pardi  from  the  coast  of 
Tuscany  in  Italy  (Guiglia,  1972,  and  references  therein).  Throughout 
most  of  Italy,  however,  the  nests  of  P.  gallicus  are  built  in  the  open, 
and  typically  hang  from  eaves,  branches,  or  other  protective 
horizontal  structures.  The  nest  hangs  from  a slender  peduncle  with 
its  disc  oriented  horizontally,  and  the  cells  opening  downward. 
Often  there  are  several  accessory  peduncles.  Infrequently,  nests  are 
found  whose  discs  are  oriented  vertically. 

In  Italy,  where  the  species  has  been  studied  extensively,  nests  of 
close  to  500  cells  have  been  reported  on  several  occasions  (Guiglia, 
1972).  It  is  apparently  common  for  colonies  of  this  species  to 
become  quite  large  in  the  south. 

Polistes  gallicus  colonies  are  haplometrotic  (with  a single  found- 
ress) in  northern  Germany  and  presumably  throughout  the  northern 
extent  of  the  species’  range.  Further  south,  for  instance  in  southern 
Germany,  pleiometrotic  colonies  (with  several  foundresses)  are 
occasionally  reported  (Richards,  1953).  In  Italy  the  species  is 
typically  pleiometrotic,  although  as  with  the  pleiometrotic  colonies 
reported  from  Germany,  one  queen  is  clearly  dominant  and  lays 
most  of  the  eggs.  The  subordinate,  or  accessory  females  function  as 
workers  in  the  nest.  According  to  Pardi  (1948),  after  the  first 
workers  emerge  the  accessory  females  are  chased  off  the  nest  or 
killed  by  the  queen.  This  situation  resembles  nest  founding  in  P. 
fuscatus  (Fabricius),  the  common  paper  wasp  in  northeastern 
United  States,  except  that  in  P.  fuscatus  subordinate  females  are 
usually  allowed  to  remain  on  the  nest  after  workers  have  emerged 
(West,  1967).  In  Africa  P.  gallicus  colonies  reproduce  by  swarming, 
with  a reproductive  female  leaving  her  nest  in  the  company  of 
several  workers  (Richards,  1953). 

Local  Observations 

I report  here  on  2 nests  of  P.  gallicus  in  Cambridge,  Massachu- 


1981] 


Hathaway — Polistes  gallicus 


171 


setts.  Both  were  in  enclosed  situations,  similar  to  those  described  as 
typical  in  the  northern  parts  of  the  species’  range  in  Europe.  The 
first  nest  was  located  inside  a metal  pole  supporting  a stop  sign.  The 
pole  was  IV2  cm  in  diameter,  and  open  at  the  top.  The  single 
peduncle  of  this  nest  was  located  28  cm  from  the  top  of  the  pole.  The 
nest  was  suspended  from  the  pole’s  side  and  faced  north.  This  nest 
contained  134  cells,  and  measured  8 cm  high  and  5 cm  across. 

A second  nest  was  located  inside  an  open  vertical  pipe,  35  cm  tall 
and  8 cm  in  diameter.  The  nest  was  suspended  from  the  side  of  the 
pipe  and  faced  west-north-west.  Its  peduncle  was  located  6 cm  from 
the  top  of  the  pipe.  This  nest  contained  153  cells  and  also  measured 
8 cm  X 5 cm. 

P.  gallicus  does  not  seem  to  be  an  aggressive  species.  I have  been 
able  to  observe  a nest  from  as  close  as  15  cm,  apparently  without 
disturbing  the  wasps. 

The  prognosis  for  permanent  establishment  of  P.  gallicus  in  the 
western  hemisphere  appears  good.  The  species  seems  quite  able  to 
withstand  the  climate  in  the  northeast.  1980-1981  was  an  unusually 
cold  year  in  Boston,  with  5,819  degree  days  accumulated  between 
June  1 and  May  30,  as  opposed  to  the  30-year  normal  of  5,597 
(United  States  National  Weather  Service  statistics,  telephone  in- 
formation). 


Recognition  of  Polistes  gallicus 

In  northeastern  United  States  P.  gallicus  would,  more  likely  be 
confused  with  a yellow  jacket  ( Vespula  spp.)  than  with  another 
paper  wasp.  Although  its  shape  and  flight  are  similar  to  the  native 
Polistes,  it  is  relatively  small  and  its  markings  and  coloring  are 
strikingly  different.  P.  gallicus  is  black  with  bright  yellow  macula- 
tions  (see  figure  1). 

Specimens  collected  in  Massachusetts  have  varied  considerably  in 
their  markings,  with  some  showing  more  yellow  than  others, 
especially  on  the  clypeus.  This  has  also  been  true  of  specimens 
collected  from  the  same  nest.  P.  gallicus  is  known  to  be  quite 
variable  in  Europe  (Guiglia,  1972). 

Males  have  completely  yellow  faces  and  their  antennae  are  curled 
at  the  tips,  a characteristic  common  in  the  genus.  Their  antennae  are 
quite  short,  however,  compared  to  males  of  other  species.  In  other 
superficial  respects,  males  of  P.  gallicus  resemble  the  females. 


172 


Psyche 


[Vol.  88 


Figure  1.  Polistes  gallicus  worker  (left)  and  male  (right)  collected  in  Cambridge, 
Massachusetts.  Length  of  worker,  1.4  cm. 


Acknowledgements 

I wish  to  thank  Dr.  Howard  E.  Evans  and  Dr.  Ronald  J. 
McGinley  for  their  helpful  comments  on  the  manuscript,  Dr.  Frank 
M.  Carpenter  for  photographing  the  specimens,  and  Dr.  Arnold  S. 
Menke  for  the  species  identification.  I also  wish  to  thank  Mr.  Loren 
Preston  of  the  Cambridge  Department  of  Traffic  and  Parking,  for 
the  loan  of  a stop  sign  pole,  and  Mrs.  Rita  Kelley  for  permission  to 
work  on  Her  property. 


References 


Giner  Mari,  J. 

1945.  Himenopteros  de  Espana,  fams.  Vespidae,  Eumenidae,  Masaridae, 
Sapygidae,  Scoliidae,  y Thynnidae.  Madrid:  Instituto  Espanol  de 
Entomologia.  142  pp. 

Gijiglia,  Delfa 

1972.  Les  Guepes  Sociales  d’Europe  Occidentale  et  Septentrionale.  Faune  de 
l’Europe  et  du  Bassin  Mediterranean,  VI.  Paris:  Masson  et  Cie.  181  pp. 

Pardi,  L. 

1948.  Dominance  order  in  Polistes  wasps.  Physiol.  Zool.  21(1):  1-13. 


1981] 


Hathaway — Polistes  gallicus 


173 


Richards,  O.  W. 

1953.  The  social  insects.  London:  Macdonald  and  Co.  219  pp. 

Spradbury,  J.  Philip 

1973.  Wasps:  an  account  of  the  biology  and  natural  history  of  solitary  and 
social  wasps.  Seattle:  University  of  Washington  Press.  408  pp. 

West,  Mary  Jane 

1967.  Foundress  associations  in  polistine  wasps:  dominance  hierarchies  and 
the  evolution  of  social  behavior.  Science  157:  1584-1585. 

Yoshikawa,  Kimio 

1962.  Introductory  studies  on  the  life  economy  of  polistine  wasps,  VI: 
geographical  distribution  and  its  ecological  significances.  J.  Biol.  Osaka 
City  University  14:  19-43. 


NOTES  ON  THE  POPULATION  ECOLOGY  OF  CICADAS 
(HOMOPTERA:  CICADIDAE)  IN  THE  CUESTA  ANGEL 
FOREST  RAVINE  OF  NORTHEASTERN  COSTA  RICA* 


By  Allen  M.  Young 
Department  of  Invertebrate  Zoology, 

Milwaukee  Public  Museum 
Milwaukee,  Wisconsin  53233 

Introduction 

Several  previous  field  studies  of  cicadas  (Homoptera:  Cicadidae) 
in  Costa  Rica  have  revealed  that  different  sympatric  genera  and 
species  often  exhibit  allochronic  (seasonal)  annual  adult  emergence 
patterns  and  habitat  associations  (Young  1972;  1974;  1975a;  1976; 
1980a, b,c;  1981a,b,c).  Most  of  these  studies  concerned  cicadas 
associated  with  lowland  tropical  forest  and  the  Central  Valley 
regions  of  Costa  Rica,  although  one  study  in  particular  (Young 
1975)  examined  some  aspects  of  the  population  ecology  of  cicadas 
in  a mountain  forest.  Because  different  species,  and  sometimes 
genera,  of  cicadas  are  found  in  different  climatic  and  geographical 
regions  of  Costa  Rica  (Young  1976),  it  is  necessary  to  examine  the 
population  ecology  of  these  insects  in  as  many  of  these  ecological 
zones  as  possible.  This  paper  summarizes  an  ecological  survey  of  the 
cicadas  thriving  in  the  steep  and  very  rugged  forest  ravine  known  as 
“Cuesta  Angel”  in  the  Central  Cordillera  of  northeastern  Costa 
Rica.  The  information  reported  here  complements  the  studies  of 
cicadas  in  other  ecological  zones  of  Costa  Rica,  although  by  no 
means  does  as  extensively  owing  to  the  difficulties  working  on  the 
very  steep  slopes  of  the  ravine.  It  is  shown  tentatively  that  (a)  the 
cicada  fauna  of  this  region  includes  at  least  two  species  not 
discussed  or  found  in  the  other  regions  studied,  (b)  the  resident 
species  exhibit  different  annual  emergence  patterns,  and  (c)  nymphal 
skins  of  several  species  are  distributed  at  very  low  densities  and  in 
association  with  various  genera  and  species  of  leguminous  canopy- 
size  trees  in  the  ravine  habitat. 


* Manuscript  received  by  the  editor  June  12,  1981 


175 


176 


Psyche 


[Vol.  88 


Fig.  1.  The  ravine  forest  at  Cuesta  Angel,  near  Cariblanco,  Heredia  Province, 
Costa  Rica. 


Methods 

The  Cuesta  Angel  ravine  is  an  extensive  strip  of  very  steep 
primary  and  river-bottom  forest  (Fig.  1)  filtering  down  from  the 
highest  mountains  of  Costa  Rica’s  Cordillera  Central  and  tapering 
into  the  northeastern  lowlands  known  as  Sarapiqui.  Because  of  its 
rugged  profile  much  of  the  ravine  remains  blanketed  in  forest  even 
though  surrounding  level  areas  have  been  largely  converted  to 
pastures.  This  ravine  is  within  the  recently  extended  Carillo  Nation- 
al Park.  There  have  been  relatively  few  field  studies  of  plants  and 
animals  in  the  ravine,  even  though  both  its  invertebrate  and 
vertebrate  faunas  contain  many  forms  not  found  in  other  parts  of 
Costa  Rica.  “Cuesta  Angel”  is  located  about  10  km  south  of  the 
village  of  Cariblanco  (10°  16'N,  84°  10'W),  Heredia  Province,  and  is 
classified  as  montane  tropical  wet  forest  (elev.  about  1200  m) 
(Holdridge  1967).  The  vertical  drop  in  the  ravine  is  about  300  m. 


1980] 


Young — Ecology  of  Cicadas 


177 


As  shown  by  1972  and  1973  rainfall  data,  the  region  is  very  wet 
and  with  a short  and  erratic  dry  season  during  January  and 
February  (Fig.  2).  For  either  collections  of  nymphal  skins  or 
determination  of  species  active  by  calling  songs  or  collection  of 
specimens,  the  locality  was  visited  the  following  dates:  27-30  June 
1972,  14  August  1972,  15-17  February  1973,  20-24  March  1973, 
18-20  April  1973  (beginning  of  nymphal  skin  regular  censuses), 
22-25  May  1973,  6 10  June  1973,  4-7  July  1973,  7-9  May  1975,  3 
April  1976,  1 and  5 November  1980.  Dates  of  visit  included  both  wet 
and  dry  periods  for  this  region.  During  the  April  1976  visit.  Dr. 
Thomas  E.  Moore  recorded  calling  songs  of  the  species  active  at 
that  time. 

The  1973  visits  were  concerned  primarily  with  attempting  to 
census  the  nymphal  skins  of  various  species  active  at  different  times 
of  the  year  while  other  dates  were  devoted  to  listening  and  collecting 
adult  specimens.  The  nymphal  skins  of  recently  emerged  cicadas  are 
relatively  easy  to  distinguish  from  those  of  a previous  years’ 
emergence  owing  to  discoloration  and  disintegration  of  some  parts 
(Young  1980a)  and  therefore  provide  an  accurate  record  of  a recent 
or  current  emergence  within  the  year.  The  locations  of  nymphal 
skins  in  the  habitat  also  provide  information  on  the  possible  feeding 
associations  of  the  nymphs  in  the  ground  and  other  aspects  of 
microhabitat.  I censused  nymphal  skins,  with  the  assistance  of  at 
least  one,  and  usually  two,  trained  student  assistants  by  marking  off 
rectangular  or  square  plots  (usually  5X5  meters)  immediately  be 
neath  a tree  or  other  spot  where  at  least  one  nymphal  skin  was 
found.  Initially  we  crawled  through  the  forest  along  transects  to 
determine  where  nymphal  skins  were  found  and  then  marked  off  the 
trees  and  places  having  them.  The  transect  approach  was  used  in  the 
survey  of  the  very  rocky  terrain  comprising  the  river-edge  forest  on 
relatively  flat  ground,  but  working  on  the  steep  slopes  entailed  spot- 
checking various  places  owing  to  the  difficulty  of  the  terrain  and 
often  very  misty  conditions.  Thus  the  nymphal  skin  census  program 
involved  repeated  censuses  at  twelve  marked  canopy-size  trees  on 
the  slopes,  and  four  large  river-edge  plots  of  forest,  each  plot 
containing  many  trees.  The  four  river-edge  plots,  each  one  widely 
separated  from  the  other  by  at  least  100  meters  of  forest,  ranged  in 
size  from  462m2  to  300m2,  the  differences  being  due  to  rivulet 
channels  and  other  interruptions  in  the  forest.  With  the  exceptions 
of  marked  trees  2,  6,  and  7 (each  of  which  was  a plot  of  about  90m2), 


178  Psyche  [Vol.  88 


SUCCESSIVE  MONTHS 

Fig.  2.  A sample  of  three  separate  years  of  monthly  rainfall  patterns  at 
Cariblanco.  In  all  three  years  portrayed,  a short  dry  spell  occurs  between  January 
and  March,  although  conditions  are  not  completely  dry  as  in  other  regions  of  Costa 
Rica  with  distinct  dry  seasons. 


1980] 


Young — Ecology  of  Cicadas 


Fig.  3.  The  forest  habitat  at  the  top  of  the  ravine,  and  above  the  Sarapiqui 
roadcut.  The  cicada  Fidicina  n.sp.  is  abundant  here. 


most  tree  plots  on  the  slopes  were  25m2.  The  twelve  tree  plots  gave  a 
total  habitat  area  of  about  484m2  sampled  for  nymphal  skins  several 
times  and  a total  of  6,957m2)  of  river-edge  forest  sampled  as  well 
(total  sample  area  of  7,441m2).  The  tree  plots  were  widely  scattered 
with  the  closest  being  no  less  than  30  meters  apart.  The  sample 
included  the  hill-top  forest  above  the  Sarapiqui  roadcut  (Fig.  3)  as 
well  as  the  forest  habitat  to  either  side  of  the  secondary  road  down 
into  the  ravine  (Fig.  4).  A census  consisted  of  making  an  exhaustive 
collection  of  all  cicada  nymphal  skins  found  within  each  plot, 
including  those  attached  to  plants  and  tree  trunks  and  those  lying  in 
the  ground  litter.  The  contents  were  placed  into  a plastic  bag  and 
labeled  appropriately.  Later  the  skins  were  determined  to  species 
and  sex.  The  nymphal  skins  of  the  cicadas  studied  were  readily 
separated  to  species  in  my  field  samples  on  the  basis  of  marked 
differences  in  size,  color,  and  body  profile.  Skins  were  matched  with 


180 


Psyche 


[Vol.  88 


others  obtained  from  collecting  skins  when  adults  were  emerging.  In 
previous  studies  (Young  1972;  1975a;  1980a, b;  1981a, c)  I have 
illustrated  and  discussed  distinguishing  features  of  cicada  nymphal 
skins.  Based  upon  these  materials,  a key  to  the  Costa  Rican  cicada 
fauna,  using  both  adults  and  nymphal  skins,  is  being  formulated 
(T.E.  Moore  and  A.M.  Young,  in  preparation).  In  the  present  study, 
it  was  very  easy  to  distinguish  nymphal  skins  of  Fidicina  species 
(three  species)  on  clear-cut  differences  in  color  pattern  and  size;  the 
Zammara  species  studied  has  nymphal  skins  very  different  in  color 
and  body  profile  from  the  others  (see  also  Young  1972),  while  the 
two  species  of  Carineta  species  had  nymphal  skins  differing  in  color, 
even  though  of  very  similar  size.  One  species  has  a very  dark  brown 
nymphal  skin,  and  the  other,  light  brown.  Based  upon  matching  of 
skins  with  adults  done  by  myself  and  T.E.  Moore,  I am  reasonably 
certain  that  matches  of  field  collections  of  skins  with  adults  is  very 
reliable.  Voucher  specimens  of  fruits  and  leaves  of  the  trees  having 
nymphal  skins  beneath  them  were  collected  and  sent  to  specialists 
for  determination. 

Other  observations  included  determining  the  places  on  the  ravine 
where  adult  cicadas  were  heard  chorusing  as  a means  of  estimating 
preferences  among  species  for  the  river-edge  area  and  top  of  the 
ravine.  In  some  instances,  diurnal  patterns  of  calling  were  also  noted 
and  the  trees  used  for  calling.  Once  the  species  were  determined, 
records  of  captures  of  cicadas  in  other  regions  of  Costa  Rica  were 
checked  by  examining  the  University  of  Michigan  collections  and 
data  bank  on  Neotropical  species  in  other  museums,  as  a means  of 
determining  if  the  Cuesta  Angel  species  were  found  elsewhere  in 
Costa  Rica.  Because  virtually  nothing  is  known  about  the  geo 
graphical  distributions  and  habits  of  Neotropical  cicadas  in  general, 
vouchers  of  both  adults  and  nymphal  skins  were  saved  and  placed  in 
collections  at  the  University  of  Michigan  and  the  Milwaukee  Public 
Museum. 

Owing  to  the  steep  terrain  and  heavy  rains  of  the  region,  a small 
experiment  was  conducted  on  estimating  the  rate  of  disintegration 
of  cicada  nymphal  skins  on  both  forested  slope  and  river-edge 
forest.  Such  a test  would  tell  me  how  many  skins  were  being  missed 
between  census  intervals  because  they  were  possibly  disintegrated, 
particularly  on  the  slopes,  before  the  next  census  was  taken.  Thus  in 
the  May  1973  census,  two  groups  of  fresh  skins  of  one  of  the  larger 
species,  each  group  containing  ten  skins,  were  established,  one 


1980] 


Young — Ecology  of  Cicadas 


181 


Fig.  4.  The  forest  habitat  along  the  secondary  road  going  to  the  bottom  of  the 
ravine.  Cicadas  such  as  Fidicina  sericans,  F.  mannifera,  and  two  species  of  Carineta 
are  heard  in  the  trees  along  this  road. 

group  of  a patch  of  forest  slope  where  this  species  emerges,  the  other 
on  a level  area  adjacent  to  the  Sarapiqui  River.  The  skins  in  each 
group  were  randomly  distributed  (by  throwing)  within  a one-meter 
square  area  of  ground.  The  numbers  of  skins  remaining  in  each  plot 
were  then  checked  in  June  and  July  1973. 

Results 

The  six  species  of  cicadas  found  and  studied  at  Cuesta  Angel  are 
shown  in  Fig.  5,  and  they  are:  Zammara  tympanum  Distant, 
Fidicina  sericans  Stal,  Fidicina  “new  species”  (n.sp.),  a new  species, 
Fidicina  mannifera  Fabricius,  Carineta  postica  Walker,  and  Cari- 
neta sp.  Three  of  these,  Z.  tympanum,  F.  sericans,  and  F.  manni- 
fera, are  large-bodied  cicadas  with  very  loud  calls,  while  F.  n.sp.  is 
medium-sized,  and  the  two  species  of  Carineta  are  considered  small- 
sized (or  at  the  low  end  of  the  medium-size  range),  the  latter  two 


182 


Psyche 


[Vol.  88 


Fig.  5.  Cicadas  found  in  the  Cuesta  Angel  ravine  forest,  top,  from  left  to  right: 
Zammara  typanum,  Fidicina  mannifera,  F.  sericans;  bottom,  left  to  right:  F.  n.sp., 
Carineta  sp.,  and  C.  postica.  The  vertical  black  line  to  the  left  of  each  cicada  gives  the 
scale  of  one  cm.  relative  to  the  body  shown  in  each  photograph. 


1980] 


Young — Ecology  of  Cicadas 


183 


cicadas  having  very  soft  calls.  Zammara  tympanum  adults  are  heard 
throughout  most  of  the  year  and  sometimes  during  the  dry  season 
and  they  call  from  the  moss  and  other  epiphyte-covered  trunks  of 
forest  trees  primarily  along  the  river-edge.  This  cicada  is  mottled 
green  and  brown  and  has  brown  spots  on  the  wings,  immediately 
distinguishing  it  from  the  others.  The  call  is  a “winding  up-like 
pulsating  buzz.  Adults  when  calling  occur  at  one  per  tree,  and  there 
are  usually  no  more  than  two  or  three  calling  males  present  within 
approximately  800m2  parcels  of  river-edge  forest  during  an  optimal 
calling  period.  Males  call  throughout  the  day,  including  overcast 
and  light  drizzle  conditions.  Males  are  bright  green  with  brown 
markings  while  females  are  drab  olive  green  and  brown. 

Fidicina  sericans,  both  sexes,  are  black  with  green  markings  on 
the  thorax  and  smoky  wings.  The  call  is  a steady  rather  high-pitched 
buzz  most  frequently  heard  during  sunny  weather  and  during  the 
dry  season.  Sometimes  several  males  congregate  in  the  same  tree, 
particularly  if  it  is  along  an  edge  of  forest,  and  sometimes,  under 
these  conditions,  several  adjacent  exposed  trees  may  have  males 
calling  at  the  same  time.  The  calling  males  are  seen  perched  on  the 
upper  portions  of  the  trunk  and  on  branches,  and  they  are  easily 
spotted  on  light-colored  bark  species  such  as  Pourouma  and 
Cecropia.  Adult  densities,  as  indicated  by  calling  males,  probably 
are  about  1-20  cicadas  per  800m2  of  forest  during  a period  of  peak 
calling,  although  this  may  be  an  underestimate  since  only  a fraction 
of  males  may  chorus  at  any  one  time.  Calling  males  are  heard 
primarily  on  the  forest  slope  and  less  so  at  the  bottom  of  the  ravine 
and  at  the  very  top. 

Fidicina  n.sp.,  both  sexes,  possesses  a green  head  and  thorax  and 
black  and  orange-banded  abdomen,  sometimes  with  patches  of 
silvery  hairs  laterally.  Of  all  of  the  cicadas  in  Costa  Rica,  this  species 
is  the  most  difficult  one  to  catch  because  of  their  habit  to  perch  very 
high  in  trees  and  to  change  trees  after  one  call.  Based  on  compari- 
sons with  type  materials  and  other  specimens,  this  is  most  likely  a 
new  species.  It  has  a very  distinctive  two-part  call:  the  first  part  is  a 
series  of  pulsating  chirps  followed  by  a longer  period  of  siren-like 
and  pulsating  calls.  Unlike  this  species,  both  Z.  tympanum  and  F. 
sericans,  as  well  as  the  other  species  to  be  discussed,  often  make 
repeated  complete  calls  from  the  same  perch,  even  if  interspersed 
with  periods  of  silence  lasting  several  minutes  or  an  hour  or  two.  F. 
n.sp.  is  heard  during  the  dry  season  and  it  occurs  in  the  ravine  and 


184 


Psyche 


[Vol.  88 


above  the  Sarapiqui  roadcut.  Adult  densities  appear  to  be  very  low, 
similar  to  that  of  Z.  tympanum,  but  difficult  to  determine  due  to  the 
highly  mobile  habits  of  males. 

Fidicina  mannifera,  both  sexes,  is  dark  brown  with  some  dark 
green  markings  on  the  head  and  thorax  and  with  tinges  of  brown 
along  the  veins  of  the  wings.  The  body  is  very  pubescent.  Males 
generally  call  at  dusk  and  dawn  and  usually  for  about  15  20  minutes 
during  each  period.  The  call  is  a very  intense  pulsating  shrill  buzz. 
Based  upon  observing  a total  of  close  to  20  individuals  at  this 
locality,  there  is  about  a 50:50  chance  that  a male  just  completing  a 
call  will  stay  in  the  same  tree.  Males  are  heard  primarily  inside  the 
forest  and  on  the  lower  slope  and  along  the  river.  Densities  are  very 
low  with  probably  only  one  or  two  males  per  1000m2  of  forest 
habitat. 

Carineta  postica,  also  illustrated  in  Young  (1975),  is  black  with 
green  markings  on  the  head  and  thorax  and  with  the  entire  body 
blanketed  in  setae.  The  wings  are  smoky  and  calling  males  have  the 
habit  of  perchng  head-downward  on  the  trunks  of  forest  trees,  a 
behavioral  trait  separating  the  larger-sized  members  of  the  genus 
from  all  other  Neotropical  cicadas.  Males  sing  from  moss-covered 
tree  trunks  and  branches  inside  the  river-edge  forest  and  along  the 
river  itself.  Densities  are  low,  with  1-5  calling  males  per  500m2  of 
forest  and  with  calling  limited  to  the  late  afternoon  or  overcast 
conditions  during  the  dry  season.  The  call  consists  of  repeated 
coarse  “zip-zip”  sounds,  and  is  reminiscent  of  a muted  version  of  the 
call  of  the  familiar  cone-headed  grasshopper  of  North  America. 
This  species  may  also  be  C.  trivitatta  Walker  as  specimens  of  both 
species  are  very  similar  in  size  and  coloration.  Clarification  awaits 
further  study. 

Carineta  sp.  is  pea-green  with  clear  wings  and  calls  from  forest 
edge  trees  such  as  Ceeropia  during  the  wet  season.  It  is  of  same  size 
and  profile  as  C.  postica  but  is  most  abundant  near  the  top  of  the 
ravine.  The  call  is  also  similar  to  that  of  C.  postica  but  somewhat 
louder  and  calling  is  generally  a dusk  phenomenon.  Sometimes  as 
many  as  eight  males  have  been  seen  perched  at  different  heights  on 
the  trunk  of  the  same  Ceeropia  tree. 

The  data  on  temporal  emergence  patterns  annually  from  the 
censuses  of  nymphal  skins  present  a more  diffuse  picture  of 
seasonality  in  the  cicadas  at  Cuesta  Angel  (Fig.  6).  Caution  is  given 
here  in  that  these  data  are  very  fragmentary  and  discontinuous, 


10 

o 

20 

10 

0 

20 

lO 

0 

to 

O 

20 

10 

O 

70 

60 

50 

40 

30 

20 

10 

O 

ontl 


Young — Ecology  of  Cicadas 


185 


Carinefa  postica 

>g 


81 


Carineta  sp. 

iiii  1 

«l 

Zarorora  tympamam 

- 

" 

ON 

Fictfcina  mannifera 

Fadicina  n.  sp. 

NO 

DATA 

— : 

Fidicina  sericarts 

MMj 

- 

mu 

NO 

DATA 

Itfl 

L. 

April 

1973 


May  June  July  August 


Months  of  census 


collections  of  cicada  nymphal  skins  from  tree  plots  and  river- 
x Angel. 


186 


Psyche 


[Vol.  88 


although  the  best  there  are  at  this  time.  With  the  exception  of  C. 
postica,  there  appears  to  be  a trend  for  most  species  to  emerge 
during  both  wet  and  dry  seasons,  when  considering  both  the 
nymphal  skin  and  call  records  together.  Thus  although  F.  sericans  is 
heard  in  abundance  during  the  short  dry  season,  there  is  some 
evidence  of  emergence  well  into  the  wet  season  (Fig.  6).  But 
examining  the  1973  rainfall  data  shows  a marked  dip  in  rainfall 
during  July  (Fig.  2),  giving  a brief  dry  spell  that  month.  If  it  is 
assumed  that  the  data  are  actually  representative  of  emergence 
patterns  of  cicadas  at  Cuesta  Angel,  it  then  appears  that  another  dry 
season  species,  C.  postica,  did  not  respond  to  the  July  1973  dry  spell 
as  there  was  no  emergence  (Fig.  6).  At  the  same  time,  the  dry  spell 
was  apparently  insufficient  in  intensity  to  block  the  emergence  of 
wet  season  species  such  as  Z.  tympanum.  Perhaps  even  more 
interesting  is  the  wet  season  emergence  of  another  supposedly  dry 
season  species,  F.  n.sp.  (Fig.  6).  Adults  of  such  species  were  not 
heard  at  these  times  although  my  sample  sizes  are  very  small. 
Different  patterns  of  emergence  may  be  associated  with  different 
years  in  whch  monthly  rainfall  regimes  are  very  different.  For 
example,  during  7-9  May  1975,  there  was  an  abundance  of  F.  n.sp 
calling  in  the  ravine  as  was  the  case  for  4-7  July  1973.  Both  of  these 
months,  in  different  years,  were  drier  than  in  other  years,  and  the 
rainfall  data  for  1972  and  1973  clearly  show  the  year-to-year 
variation  in  monthly  rainfall  patterns  at  this  locality  (Fig.  2). 
Furthermore,  when  F.  n.sp.  emerged  during  the  wet  season,  calling 
was  restricted  to  the  dry  periods  of  the  day.  All  of  the  cicadas 
studied  exhibit  bursts  of  calling  near  dusk  (see  also  Young  1981b). 

The  distribution  of  nymphal  skins  for  each  species  studied  by 
marked  trees  is  given  in  Table  1.  Even  though  approximately  70 
species  of  canopy-size  trees  were  included  along  the  initial  transects 
to  determine  the  locations  of  cicada  emergence  patches  in  the 
ravine,  patches  were  found  to  be  confined  to  the  species  of 
Leguminosae  listed  in  Table  1.  Note  that  the  estimation  of  relative 
abundance  of  adults  among  the  species  discussed  above  is  confirmed 
here  in  terms  of  nymphal  skins:  by  far  the  most  abundant  species  is 
F.  sericans,  whose  nymphal  skins  comprised  almost  64%  of  the  total 
241  skins  collected  in  the  1973  survey  of  tree  plots  alone  (Table  1). 
F.n.sp.,  Z.  tympanum,  and  Carineta  sp.  are  about  evenly  distributed 
in  terms  of  abundance  of  nymphal  skins  in  the  tree  plots.  As  in 
previous  studies  of  cicadas  in  Costa  Rica,  sex  ratios  are  close  to 


Table  1.  Census  history  of  cicada  nymphal  skins  in  legume  tree  plots*  in  the  forest  ravine,  “Cuesta  Angel”,  near  Cariblanco, 
Heredia  Province,  Costa  Rica. 


1980] 


Young — Ecology  of  Cicadas 


187 


w <3 

<D  ^ 

c *c: 
a o 


a to 

c c 


O O O O 


O O <N  O O O O 


Tf  — O O — O 


O rM  O O 


O r’b  m O O O — r-  — Tf  — m 


0 — 00 


O O O O <N  O O 


or-  — ornniTj-oomo 


— — Tt0I^-0^v~>0 


§3  ~g 

£ s; 


a,  *5  a. 


a <3  53 

be  be  be 

c c c 


(NmTt^ot^oooNO 


O O 

+1 


O'-  Qv 
— CN 
+1 


+1 


m in 
rn  \o 
O O 

+1 


— nj 

+1 


m r- 

r-  p 
<n  — 

— <N 

+ 1 


*Each  plot  ranged  in  size  from  5X5  meters  to  10X9  meters  around  the  base  of  individual  legume  trees. 

**A11  trees  and  cicadas  were  censused  18-20  April,  22-25  May,  4-7  July,  15  August  1973  (total  of  14  days),  except  for  trees9,  10, 
1 1,  and  12,  which  were  added  to  the  census  program  on  4 July  1973. 


188 


Psyche 


[Vol.  88 


unity.  Taking  the  most  abundant  species,  F.  sericans,  there  is 
considerable  range  in  numbers  of  skins  found  in  the  different  tree 
plots,  although  close  to  40%  of  all  skins  of  this  species  were  found 
beneath  one  individual  of  Pithecollobium  latifolium  (Table  1).  Yet  a 
second  individual  of  this  tree  species  yielded  only  four  skins  of 
cicadas  overall  and  none  of  F.  sericans.  Such  data,  although  limited, 
indicate  the  considerable  variation  encountered  over  different 
patches  of  the  same  resource  for  a cicada  species  in  tropical  forests. 
Two  different  individuals  of  Inga  and  one  P.  latifolium  together 
account  for  almost  65%  of  all  skins  found.  That  such  data  may  be 
underestimates  of  true  values,  even  for  an  abundant  species  such  as 
F.  sericans,  is  suggested  by  the  results  of  the  estimate  of  rate  of 
disintegration  of  nymphal  skins:  at  the  end  of  a five-week  period, 
between  50%  (level  ground)  and  80%  (slope)  of  the  F.  sericans 
nymphal  skins  studied  had  disappeared.  These  samples  are  pitifully 
small,  but  it  is  the  best  we  have  at  this  time.  The  intervals  between 
censuses  in  my  study  are  of  this  magnitude  and  greater,  thereby 
indicating  the  likelihood  that  some  skins  were  missed  owing  to  their 
rapid  disintegration  under  very  wet  conditions.  The  examination  of 
nymphal  skin  distributions  by  tree  plots  and  river-edge  plots 
separately  provides  further  confirmation  of  the  data  shown  in  Table 
1 (Table  2).  Although  high  percentages,  if  not  all,  of  plots  are 
occupied  by  skins  of  Z.  tympanum,  the  emergence  is  one  of  very  low 
density  since  only  a small  number  of  skins  occur  in  the  plots  studied 
(Table  2).  The  tree  plots,  although  only  representing  an  area  of 
about  6.5%  of  the  combined  area  of  tree  plots  and  river-edge  plots, 
account  for  almost  80%  of  all  skins  recovered  (Table  2).  The  larger 
river-edge  plots  include  a wide  variety  of  tree  species  whereas  the 
tree  plots  each  include  one  individual  of  a leguminous  tree  species 
and  understory  plants.  Most  striking  is  the  relatively  high  density  of 
the  nymphal  skins  of  F.  sericans  in  the  tree  plots,  almost  0.4 
skins/ m2  (Table  2).  Yet  the  same  cicada,  in  a much  larger  and 
representative  tract  of  forest,  representing  an  area  about  five  times 
that  of  the  tree  plots,  has  the  very  low  density  of  about  0.010 
skins/  m2  (Table  2).  Other  patterns  of  nymphal  skin  density  between 
tree  plots  and  river-edge  plots  are  self-evident  and  support  the 
pattern  discussed  for  F.  sericans  (Table  2).  From  such  results,  one 
can  readily  appreciate  the  distortion  of  density  estimates  when 
different  size  patches  of  the  environment,  with  different  biological 
attributes,  are  combined  to  give  a summary  figure  (Table  2).  And 


Table  2.  Some  popluation  parameters  of  cicada  species  in  the  ravine  forest,  “Cuesta  Angel”,  Costa  Rica 


1980] 


189 


Young — Ecology  of  Cicadas 


SP-  18  mm  9 3 0.002/ m2  23  8 0.060/ m2  0.005/ m2 

*There  are  four  river-edge  plots  of  these  sizes:  2,145m2,  1,350m2,  462m2,  and  3,000m2  for  total  area  of  6,957m2. 

* All  estimates  of  density  based  on  occupied  plots  only:  no  empty  plots  included. 

*Most  of  these  plots  are  25m2  for  a total  area  of  484m2. 


190 


Psyche 


[Vol.  88 


the  data  also  show,  that  for  larger  areas  of  environment  as  typified 
here  by  the  river-edge  plots,  there  are  not  necessarily  going  to  be 
increases  in  densities  of  insects  recovered. 

Discussion 

Of  the  six  species  studied  at  Cuesta  Angel,  none  are  exclusive  to 
the  locality,  but  other  locality  records  from  Costa  Rica  indicate 
similar  elevations  and  habitat.  Fidicina  n.sp.,  Zammara  tympanum, 
and  both  species  of  Carineta  have  been  collected  at  Turrialba, 
Cartago  Province  as  shown  by  specimens  in  the  collections  at  The 
University  of  Michigan  Museum  of  Zoology.  The  species  I term  C. 
postica  may  also  be  C.  trivitatta  Walker,  which  has  also  been 
collected  from  the  San  Jose  area,  Guapiles  (Limon  Province)  and 
Bajo  La  Hondura  (San  Jose  Province).  Two  other  cicadas,  F. 
sericans  and  F.  mannifera,  have  much  more  extensive  ranges  in 
Costa  Rica  as  both  have  been  collected  and  studied  in  premontane 
and  lowland  tropical  wet  forest  regions  of  the  Atlantic  coastal 
watershed  (Young  1972;  1980b),  and  mannifera  also  occurs  in  the 
semi-dry  to  dry  forest  region  of  the  western  provinces  of  Puntarenas 
and  Guanacaste  (Young  198 la, c).  Given  the  topography  of  the 
Cuesta  Angel  region  relative  to  the  adjacent  lowlands,  it  is  not 
surprising  to  find  species  such  as  sericans  and  mannifera  along  a 
more  or  less  continuous  elevational  gradient  within  the  wet  forest 
region  and  over  a range  of  about  90  1 100  meters.  Yet  this  is  not  true 
for  the  genus  Zammara  or  Carineta  since  entirely  different  species 
occur  in  the  adjacent  premontane  and  lowland  wet  forest  regions  of 
northeastern  Costa  Rica  (Young  1972;  1976;  1980b).  From  both 
records  of  adults  calling  and  nymphal  skins,  both  sericans  and 
mannifera  occur  at  much  lower  densities  in  the  Cuesta  Angel 
montane  wet  forest  than  they  do  in  adjacent  premontane  and 
lowland  wet  forests.  Given  these  records,  it  is  concuded  tentatively 
that  cicadas  such  as  F.  n.sp.,  Z.  tympanum,  and  the  two  species  of 
Carineta  studied  are  montane  species  associated  with  wet  forests 
while  F.  sericans  and  F.  mannifera  are  lower  elevation  forms  also 
associated  with  generally  wet  forests  and  semi-dry  forests.  Thus  the 
Cuesta  Angel  cicada  fauna  is  a mixture  of  montane  and  lower 
elevation  tropical  wet  forest  cicadas. 

Both  generic  and  specific  richness  of  cicadas  at  Cuesta  Angel  are 
not  as  high  as  they  are  in  the  adjacent  lower  elevations.  There  are  six 


1980] 


Young — Ecology  of  Cicadas 


191 


genera  and  about  ten  species  of  cicadas  found  in  the  adjacent 
premontane  tropical  wet  forest  (Young  1980b)  as  studied  about  25 
km  from  the  Cuesta  Angel  locality.  Young  (1975)  found  only  two 
genera,  each  monospecific,  at  another  montane  wet  forest  locality, 
Bajo  La  Hondura.  There  are  also  greater  numbers  of  genera  and 
species  found  in  mid-elevation  moist  forest  (Young  1980a)  and 
lowland  tropical  dry  forest  (Young  1981a)  in  Costa  Rica.  Cicadas 
such  as  F.  sericans  and  F.  mannifera  are  tentatively  interpreted  as 
being  ecological  “leaks”  into  the  forested  ravine  at  Cuesta  Angel. 
Given  the  continuous  accessibility  to  lower  elevation  wet  forest 
habitats  moving  along  the  ravine  into  the  lowlands,  it  is  un- 
reasonable to  expect  some  highly  mobile  insects  to  colonize  at  either 
end  (Young  1975b). 

Elsewhere  in  Costa  Rica,  cicadas  have  been  found  to  have  distinct 
seasonal  patterns  of  adult  emergences  each  year  (Young  1972; 
1975a;  1980a, b;  198 la, c)  with  the  recognition  of  usually  three  kinds 
of  cicadas:  dry  season,  wet  season,  and  transitional  forms  between 
dry  and  wet  seasons.  From  the  studies  of  cicadas  in  premontane 
tropical  wet  forest  in  particular  (Young  1980b),  however,  it  became 
apparent  that  brief  spells  of  wetness  in  a dry  period  and  of  dryness 
in  the  wet  season  may  trigger  emergence  of  wet  season  and  dry 
season  species  respectively.  In  the  premontane  tropical  wet  forest 
zone  of  northeastern  Costa  Rica,  typically  wet  season  cicadas  such 
as  Z.  smaragdina  Walker  will  emerge  in  low  numbers  during  a rainy 
spell  of  about  five  days  or  longer  within  the  dry  season  (Young 
1980b:  pers.  obs.).  During  a dry  spell  within  the  long  rainy  season  at 
the  same  locality,  F.  sericans,  a typical  dry  season  cicada,  can  also 
be  heard  and  fresh  nymphal  sLins  found  (Young  1980b;  pers.  obs.). 
Such  observations  indicate  that  “seasonality”  in  tropical  cicadas  is  a 
very  flexible  sort  of  emergence  strategy,  perhaps  determined  by 
critical  periods  of  wetness  or  dryness,  depending  upon  the  species 
and  locality.  Such  an  effect  may  explain  the  anolamous  emergence 
of  F.  sericans  in  the  wet  season  at  Cuesta  Angel.  The  data  from 
Cuesta  Angel  very  tentatively  provide  additional  support  for  this 
phenomenon,  as  shown  for  species  such  as  F.  n.sp.  The  proximal 
cues  triggering  emergence  in  tropical  cicadas  have  not  been  studied 
to  my  knowledge,  although  some  ideas  have  been  suggested  for 
study  (Young  1975;  1980a, b;  1981a).  What  are  also  needed  are 
detailed  studies  of  the  effects  of  small  changes  in  air  temperature 
and  humidity,  and  light  intensity  on  the  behavior  of  adult  cicadas  in 


192 


Psyche 


[Vol.  88 


the  tropics  over  a typical  diurnal  cycle.  Different  species  may 
possess  different  levels  of  physiological  capacity  to  cope  with 
stressful  environmental  conditions  imposed  by  either  too  much 
wetness  or  too  little  wetness.  From  my  work  on  cicadas  in  Costa 
Rica  over  the  past  eleven  years,  and  particularly  from  data  on 
densities  of  nymphal  skins  of  co-occurring  species  in  the  same 
patches  of  habitat,  it  seems  doubtful  that  seasonal  emergence 
patterns  in  cicadas  is  related  to  interspecific  competition  in  develop- 
ing cicadas.  From  what  little  information  I have,  there  is  little 
reason  to  suspect  competition  for  oviposition  sites.  But  the  great 
diversity  in  the  properties  of  the  calling  songs  among  co-occurring 
species  in  tropical  forests,  and  the  tendency  for  several  species  to 
form  single  species  aggregates  of  chorusing  males  (Young  1980c) 
suggest  that  there  might  be  competition  for  optimal  calling  condi- 
tions. In  cicadas,  the  calling  song  is  a major  component  of  fitness 
since  it  presumably  functions  in  mating,  and  there  might  be  strong 
selection  to  evolve  allochronic  emergence  patterns  when  the  calling 
songs  of  species  conflict  and  reduce  mating  success.  Certain  types  of 
seasonal  changes  in  the  environment,  yet  to  be  determined,  may 
provide  the  most  ready  cues  for  these  insects  to  exploit  in  evolving 
allochronic  emergence  patterns  to  reduce  losses  in  mating  success. 
The  whole  system  warrants  considerable  detailed  study  as  it  involves 
different  stages  in  the  life  cycle.  Seasonal  emergence  may  or  may  not 
have  anything  to  do  with  conditions  being  intrinsically  optimal  for  a 
certain  species  in  a certain  region  at  a certain  time  of  the  year.  If  the 
latter,  the  cicada  is  merely  locking  in  to  a convenient  cue  since, 
under  this  hypothesis,  both  wet  and  dry  periods  provide  suitable 
resources  for  adults,  including  those  associated  with  mating  needs. 

Under  the  mating  conflict  hypothesis,  it  is  implied  that  cicadas 
with  very  low  densities  and  with  unusual  calling  habits  may  forego 
entering  into  such  a selection  arena,  thereby  circumventing  this 
adaptive  pathway  and  emerging  throughout  most  of  the  year,  other 
things  being  equal.  A typical  case  in  point  is  the  almost  catholic 
habit  of  F.  mannifera  to  sing  for  a brief  period  at  dusk  and  under 
conditions  of  low  population  densities  in  Costa  Rica  (Young  1972; 
1980b;  1981b;  this  paper).  The  intensity  of  the  presumed  mating 
conflict  is  considered  to  increase  as  population  densities  of  co- 
occurring species  increase  individually. 

In  virtually  all  other  regions  studied,  the  greatest  numbers  of 
cicada  nymphal  skins  occur  beneath  adult  legume  trees  (Young 


1980] 


Young — Ecology  of  Cicadas 


193 


1972;  1980a, b;  198 la, c)  although  precise  data  on  the  abundances  of 
skins  in  legume  plots  versus  non-legume  plots  is  still  lacking.  At 
Cuesta  Angel  cicada  nymphal  skin  patches  too  were  found  beneath 
legume  trees.  If  legume  trees  provide  some  form  of  highly  suitable 
environment  for  developing  cicadas  in  tropical  forests,  emerging 
populations  each  year  will  be  spatially  disjunct  according  to  the 
spatial  distribution  of  the  legume  trees  whose  root  crowns  provide  a 
primary  resource  for  developing  nymphs.  The  suitability  of 
Leguminosae  for  developing  cicadas  may  involve  both  physical  and 
chemical  properties  of  the  classes  of  root  sizes  exploited  by  various 
age-classes  of  nymphs.  The  observed  low  densities  of  nymphal  skins 
in  all  of  the  plots  at  Cuesta  Angel,  relative  to  previously  obtained 
densities  of  the  same  or  similar  species  in  other  regions  (e.g.,  Young 
1980a, b;  1981a),  may  therefore  be  a function  of  the  very  dispersed 
condition  of  the  legume  trees  at  this  locality.  A striking  contrast  is 
made  with  the  association  of  nymphal  skins  of  species  such  as  Z. 
smaragdina  and  F.  sericans  in  relatively  large  patches  of  adult 
Pentaclethra  macroloba  in  nearby  premontane  tropical  wet  forest 
(Young  1980b).  Densities  of  these  cicadas  range  from  5.4  to  9.3m2  in 
patches  of  two  or  more  P.  macroloba,  estimates  considerably 
greater  than  for  the  same  species  at  Cuesta  Angel.  I interpret  such 
observations  to  be  the  result  of  P.  macroloba  occurring  as  clumps  of 
several  trees  thereby  increasing  the  size  of  a single  resource  patch  for 
cicadas,  which  results  in  either  greater  oviposition  in  the  patch  or 
greater  survival  of  nymphs,  or  both.  The  river-edge  plots  in  the 
Cuesta  Angel  study  illustrate  quite  well  such  an  effect.  Such  plots, 
although  quite  large,  only  contained  one  or  two  widely  scattered 
legume  trees  and  not  clumps  of  such  trees,  and  some  did  not  contain 
legumes  at  all  but  were  situated  near  such  trees.  The  observed  very 
low  densities  of  cicada  nymphal  skins  in  these  large  segments  of 
forest  is  due  to  an  absence  or  scarcity  of  suitable  root  crowns  for 
cicadas.  The  tree  plots,  on  the  other  hand,  although  very  small,  are 
highly  suitable  for  cicadas  and  therefore  densities  are  high. 

The  pattern  of  cicada  nymphal  skins  being  associated  with  legume 
trees  in  tropical  forests  can  have  other  explanations  as  well,  ones  not 
involving  a presumed  coevolved  interaction  of  the  sort  suggested 
above.  For  example,  selective  logging  of  tropical  forests  may  leave 
behind  the  relatively  soft-wood  legumes  thereby  increasing  their 
relative  abundance  as  a resource  for  insects  such  as  cicadas.  Thus 
the  likelihood  for  an  ovipositing  cicada  to  discover  a legume  tree 


194 


Psyche 


[Vol.  88 


increases  greatly  over  a period  of  years,  even  though  the  root  crowns 
and  other  cicada-related  characteristics  of  other  trees  are  equally 
suitable  for  cicada  development. 

Since  all  plots  were  located  at  or  withn  the  lower  one-fourth  of 
the  ravine,  the  instances  in  which  some  species,  such  as  F.  n.sp.  and 
Carineta  sp.  call  primarily  from  the  top  of  the  ravine  and  not  at  the 
bottom  suggests  a behavioral  response  associated  with  mating 
requirements.  Such  species  presumably  emerge  near  the  bottom  of 
the  ravine  and  fly  up  to  the  top  for  courtship.  Such  species  may  also 
emerge  near  the  top  as  well  although  this  was  not  determined  in  this 
study.  The  observed  patterns  of  adult  calling  sites  within  the  ravine 
are  presumably  related  to  the  acoustical  and  thermoregulatory 
needs  of  each  species. 


Summary 

The  genera  and  species  of  cicadas,  their  seasonal  distributions, 
habits,  and  emergence  sites  were  studied  discontinuously  over 
several  years  at  the  Cuesta  Angel  ravine,  a rugged  mountain  tropical 
wet  forest  locality  in  the  northern  portion  of  the  Central  Cordillera 
of  Costa  Rica.  Emphasis  was  placed  on  determining  the  distribution 
of  cicadas  down  one  steep  forested  side  of  this  approximately  300- 
meter  deep  ravine  and  along  a representative  portion  of  its  bottom 
(Rio  Sarapiqui).  Some  evidence  of  seasonal  fluctuations  in  abun 
dance  was  obtained  for  the  six  species  found  here,  and  the  greatest 
densities  of  nymphal  skins  of  all  species  were  found  in  small  plots 
around  individual  legume  trees.  Densities  in  the  large  river-edge 
plots,  containing  many  different  kinds  of  trees,  were  relatively  very 
low.  The  data  are  compared  to  similar  data  on  cicadas  from  other 
regions  of  Costa  Rica.  Tropical  cicada  seasonality,  interactions  with 
Leguminosae,  and  possible  mechanisms  underlying  population  den- 
sities, are  discussed. 


Acknowledgements 

This  research  was  originally  supported  by  National  Science 
Foundation  Grant  BG-33060  of  the  United  States  of  America,  and 
subsequently  by  the  Milwaukee  Public  Museum.  Field  assistance 
was  given  (1972-73)  by  students  from  Lawrence  University.  Dr.  J. 
Robert  Hunter,  then  Director,  Costa  Rican  Field  Studies  Program 
of  The  Associated  Colleges  of  the  Midwest,  provided  logistical 


1980] 


Young — Ecology  of  Cicadas 


195 


support.  Dr.  Dieter  C.  Wasshausen  of  the  Smithsonian  Institution 
provided  identifications  of  the  tree  species  discussed.  Dr.  Thomas  E. 
Moore  of  The  Museum  of  Zoology  at  The  University  of  Michigan 
assisted  with  the  identification  of  the  cicadas,  gave  the  author  access 
to  the  UMMZ  cicada  collections  and  other  information  concerning 
cicadas,  accompanied  the  author  on  one  of  the  visits  to  the  locality 
to  record  the  calls  of  cicadas,  and  discussed  cicada  biology  with  the 
author.  I thank  Dr.  Henk  Wolda,  who,  in  the  capacity  of  a referee 
for  this  journal,  made  many  helpful  suggestions  on  the  manuscript. 

Literature  Cited 


Holdridge,  L.  R. 

1967.  Life  zone  ecology.  Tropical  Science  Center,  San  Jose,  Costa  Rica. 

Young,  A.  M. 

1972.  Cicada  ecology  in  a Costa  Rican  tropical  rain  forest.  Biotropica 
4:152-159. 

1974.  The  population  biology  of  Neotropical  cicadas.  III.  Notes  on  the 
behavioral  natural  history  of  Pacarina  cicadas  in  some  Costa  Rican 
grasslands.  Entomol.  News  85:239-256. 

1975a.  The  population  biology  of  neotropical  cicadas.  I.  Emergences  of  Procol- 
lina  and  Carineta  in  a mountain  forest.  Biotropica  7:248-258. 

1975b.  “Leakage”  of  Morpho  theseus  (Lepidoptera:  Nymphalidae)  into  north- 
eastern Costa  Rica?  Brenesia  6:59-67. 

1976.  Notes  on  the  faunistic  complexity  of  cicadas  (Homoptera:  Cicadidae)  in 
northern  Costa  Rica.  Revista  de  Biologia  Tropical  24:267-279. 

1980a.  Habitat  and  seasonal  relationships  of  some  cicadas  (Homoptera:  Cicadi- 
dae) in  central  Costa  Rica.  American  Midi.  Naturalist  103:155-166. 

1980b.  Environmental  partitioning  in  lowland  tropical  rain  forest  cicadas.  J. 
New  York  Entomol.  Soc.  88:86-101. 

1980c.  Observations  on  the  aggregation  of  adult  cicadas  (Homoptera:  Cicadi- 
dae) in  tropical  forests.  Canadian  J.  Zoology  58:711-722. 

1981a.  Seasonal  adult  emergences  of  cicadas  (Homoptera:  Cicadidae)  in  north- 
western Costa  Rica.  Milwaukee  Public  Museum  Contributions  in 
Biology  and  Geology,  No.  40. 

1981b.  Temporal  selection  for  communicatory  optimization:  The  dawn-dusk 
chorus  as  an  adaptation  in  tropical  cicadas.  American  Naturalist:  117: 
826-829. 

1981c.  Notes  on  seasonality  and  habitat  associations  of  cicadas  (Homoptera: 
Cicadidae)  in  premontane  and  montane  tropical  moist  forest  in  Costa 
Rica.  J.  New  York  Entomol.  Soc.,  89:  123-142. 


CAMBRIDGE  ENTOMOLOGICAL  CLUB 


A regular  meeting  of  the  Club  is  held  on  the  second  Tuesday 
of  each  month  October  through  May  at  7:30  p.m.  in  Room  154, 
Biological  Laboratories,  Divinity  Avenue,  Cambridge.  Entomolo- 
gists visiting  the  vicinity  are  cordially  invited  to  attend. 


BACK  VOLUMES  OF  PSYCHE 

Requests  for  information  about  back  volumes  of  Psyche  should 
be  sent  directly  to  the  editor. 


F.  M.  Carpenter 

Editorial  Office,  Psyche 
16  Divinity  Avenue 
Cambridge,  Mass.  02138 


FOR  SALE 


Reprints  of  articles  by  W.  M.  Wheeler 

The  Cambridge  Entomological  Club  has  for  sale  numerous  reprints 
of  Dr.  Wheeler’s  articles  that  were  filed  in  his  office  at  Harvard 
University  at  the  time  of  his  death  in  1937.  Included  are  about 
12,700  individual  reprints  of  250  publications.  The  cost  of  the 
reprints  has  been  set  at  5c  a page,  including  postage;  for  orders 
under  $5  there  will  be  an  additional  handling  charge  of  50c.  A list  of 
the  reprints  is  available  for  $1.00  from  the  W.  M.  Wheeler  Reprint 
Committee,  Cambridge  Entomological  Club,  16  Divinity  Avenue, 
Cambridge,  Mass.  02138.  Checks  should  be  made  payable  to  the 
Cambridge  Entomological  Club. 


ISSN  0033-2615 


PSYCHE 

A JOURNAL  OF  ENTOMOLOGY 

founded  in  1874  by  the  Cambridge  Entomological  Club 

Vol.  88  1981  No.  3-4 

CONTENTS 

Dedication:  Robert  E.  Silberglied.  Frank  M.  Carpenter 197 

Sound  Production  by  Courting  Males  of  Phidippus  mystaceus  (Araneae: 

Salticidae).  G.  B.  Edwards  199 

Maternal  Behavior  and  Alarm  Response  in  the  Eggplant  Lace  Bug,  Gargaphia 
solani  Heidemann  (Tingidae:  Heteroptera).  R.  S.  Kearns  and 

R.  T.  Yamamoto 215 

Polymorphism  and  Division  of  Labor  in  the  Dacetine  Ant  Orectognathus 

versicolor  (Hymenoptera:  Formicidae).  Norman  F.  Carlin 231 

Trail  Communication  of  the  Dacetine  Ant  Orectognathus  versicolor 

(Hymenoptera:  Formicidae).  Bert  Holldobler 245 

Francis  Walker  Types  of,  and  New  Synonymies  for,  North  American 
Hydropsyche  species  (Trichoptera,  Hydropsychidae).  Andrew  P.  Nimmo  259 
Territoriality,  Nest  Dispersion,  and  Community  Structure  in  Ants. 

Sally  C.  Levings  and  James  F.A.  Traniello 265 

The  Effect  of  Flower  Occupancy  on  the  Foraging  of  Flower-Visiting  Insects. 

V.  J.  Tepedino  and  F.  D.  Parker 321 

Abdominal  Trophallaxis  in  the  Slave-Making  Ant,  Harpagoxenus  americanus 

(Hymenoptera:  Formicidae).  Robin  J.  Stuart 331 

New  Name  for  the  Extinct  Genus  Mesagyrtes  Ponomarenko  (Coleoptera: 

Silphidae  S.L.).  Alfred  F.  Newton,  Jr 335 

Historical  Development  of  Bee  Foraging  Patterns  in  Central  New  York  State. 

Howard  S.  Ginsberg 337 

Myrmecophilic  Relationship  of  Pella  (Coleoptera: Staphylinidae)  to  Lasius 
fuliginosus  (Hymenoptera:  Formicidae)  B.  Holldobler,  M.  Moglich,  and 

U.  Maschwitz 347 

Behavioral  Origin  of  Tremulation,  and  Possible  Stridulation,  in  Green 
Lacewings  (Neuroptera:  Chrysopidae).  Peter  Duelli  and  James  B.  Johnson  375 

Arthropods  Attracted  to  Luminous  Fungi.  John  Sivinski 383 

Index  to  Volume  88  391 


CAMBRIDGE  ENTOMOLOGICAL  CLUB 
Officers  for  1981-1982 

President Barbara  L.  Thorne 

Vice-President  Frances  Chew 

Secretary Heather  Hermann 

Treasurer  Frank  M.  Carpenter 

Executive  Committee  John  Shetterly 

Mary  Hathaway 


EDITORIAL  BOARD  OF  PSYCHE 

F.  M.  CARPENTER  (Editor),  Fisher  Professor  of  Natural  History, 
Emeritus,  Harvard  University 

W.  L.  BROWN,  Jr.,  Professor  of  Entomology,  Cornell  University  and 
Associate  in  Entomology,  Museum  of  Comparative  Zoology 
P.  J.  DARLINGTON,  Jr.,  Professor  of  Zoology,  Emeritus,  Harvard 
University 

B.  K.  HOLLDOBLER,  Professor  of  Biology,  Harvard  University 
H.  W.  LEVI,  Alexander  Agassiz  Professor  of  Zoology,  Harvard  University 
R.  J.  McGlNLEY,  Assistant  Professor  of  Biology,  Harvard  University 
ALFRED  F.  Newton,  Jr.,  Curatorial  Associate  in  Entomology,  Harvard 
University 

R.  E.  SlLBERGLlED,  Smithsonian  Tropical  Research  Institute,  Panama 
E.  O.  WILSON,  Baird  Professor  of  Science,  Harvard  University 


PSYCHE  is  published  quarterly  by  the  Cambridge  Entomological  Club,  the  issues 
appearing  in  March,  June,  September  and  December.  Subscription  price,  per  year, 
payable  in  advance:  $11.00,  domestic  and  foreign.  Single  copies,  $3.50. 

Checks  and  remittances  should  be  addressed  to  Treasurer,  Cambridge 
Entomological  Club,  16  Divinity  Avenue,  Cambridge,  Mass.  02138. 

Orders  for  missing  numbers,  notices  of  change  of  address,  etc.,  should  be  sent  to  the 
Editorial  Office  of  Psyche,  16  Divinity  Avenue,  Cambridge,  Mass.  02138.  For  previous 
volumes,  see  notice  on  inside  back  cover. 

IMPORTANT  NOTICE  TO  CONTRIBUTORS 
Manuscripts  intended  for  publication  should  be  addressed  to  Professor  F.  M. 
Carpenter,  Biological  Laboratories,  Harvard  University,  Cambridge,  Mass.  02138. 

Authors  are  expected  to  bear  part  of  the  printing  costs,  at  the  rate  of  $27.50  per 
printed  page.  The  actual  cost  of  preparing  cuts  for  all  illustrations  must  be  borne  by 
contributors:  the  cost  for  full  page  plates  from  line  drawings  is  ordinarily  $18.00  each, 
and  for  full  page  half-tones,  $30.00  each;  smaller  sizes  in  proportion. 


Psyche,  vol.  88,  no.  1-2,  for  1981,  was  mailed  Deember  28,  1981 


The  Lexington  Press,  Inc.,  Lexington,  Massachusetts 


Robert  Elliot  Silberglied 

This  issue  of  Psyche  is  dedicated  to  the  memory  of  Robert  E. 
Silberglied,  a victim  of  the  Air  Florida  accident  in  Washington, 
D.C.,  on  January  13,  1982. 

Born  in  Brooklyn,  N.Y.,  in  1946,  Bob  was  already  an  enthusiastic 
naturalist  and  entomologist  even  in  his  school  days.  He  graduated 
from  Cornell  University  in  1967  and  received  his  PhD  from  Harvard 
in  1973.  He  remained  at  Harvard  until  July  of  1981,  as  Assistant 
Professor  and  later  as  Associate  Professor  in  the  Department  of 
Biology,  teaching  mainly  the  courses  in  entomology  that  I had  given 
for  many  years  as  his  predecessor.  He  was  also  Assistant  Curator 
and  later  Associate  Curator  in  the  Entomology  Department  of  the 
Museum  of  Comparative  Zoology.  During  the  same  period  he  was 
associated  with  the  Smithsonian  Tropical  Research  Institute,  spend- 
ing about  half  of  each  year  in  Panama  or  other  parts  of  the 
American  tropics.  At  the  time  of  his  death  he  was  Staff  Scientist 
(Research  Entomologist)  at  the  Institute. 

Bob  joined  our  society  on  his  arrival  in  Cambridge  in  1968  and 
for  the  next  14  years  he  was  one  of  our  most  active  and  enthusiastic 
members.  He  served  as  vice-president  and  president,  and  was  a 
member  of  the  editorial  board  of  Psyche  for  the  past  decade.  At  our 
fall  and  winter  meetings,  he  could  always  be  depended  upon  to 
relate  some  unusual  collecting  experience  or  to  demonstrate  with 
superb  photographs  and  specimens  some  of  the  remarkable  insects 
that  he  had  collected  in  the  tropics.  He  combined  a warm  and 
sympathetic  personality  with  a brilliant  and  imaginative  mind.  In 
both  respects  he  has  left  a lasting  impression  on  our  society  and  its 
members. 

The  Smithsonian  Institution  has  established  the  Robert  E.  Silber- 
glied Memorial  Fund  to  support  student  research  and  training  in 
tropical  entomology.  Those  who  wish  to  contribute  a gift  of  any  size 
may  send  it  to:  Robert  E.  Silberglied  Memorial  Fund,  Accounting 
Office,  Smithsonian  Institution,  L’Enfant  3500,  Washington  D.C. 
20560. 


Frank  M.  Carpenter,  editor 


197 


Robert  Elliot  Silberglied 

Photograph  taken  in  1981 


PSYCHE 


Vol.  88 


1981 


No.  3-4 


SOUND  PRODUCTION  BY  COURTING  MALES  OF 

PHID1PPUS  MYSTACEUS  (ARANEAE:  SALTICIDAE)i 

By  G.  B.  Edwards* 2 

The  courtship  rituals  of  male  salticids  generally  are  considered  to 
be  visually-oriented,  despite  the  fact  that  a primarily  tactile  type  of 
courtship  has  been  demonstrated  for  2 species  of  Phidippus  (Ed- 
wards, 1975;  Jackson,  1977).  In  addition,  chemotactic  cues  probably 
assist  a male  in  locating  a female  in  most  species  of  jumping  spiders 
(Crane,  1949;  Richman,  1977).  I report  here  that  males  of  Phidippus 
mystaceus  (Hentz)  produce  sound  by  means  of  a palpal  stridulatory 
mechanism  as  an  integral  part  of  their  courtship.  This  is  the  first 
known  case  of  a salticid  producing  sound  with  this  type  of 
mechanism;  a similar  stridulatory  organ  has  been  reported  for 
lycosid  spiders  (Rovner,  1975). 

Petrunkevitch  (1926)  reported  that  the  salticid  Stridulattus  stridu- 
lans  Petrunkevitch  has  a stridulatory  organ  (of  type  “d”,  chelicera- 
palpus;  Legendre,  1963).  However,  he  did  not  detect  sound  produc- 
tion. The  only  other  records  of  a salticid  producing  sound  were  by 
Bristowe  (1958),  who  reported  that  Euophrys  frontalis  (Walckenaer) 
made  a “distinct  sound  as  the  tarsal  claws  (of  the  legs  I of  the  male) 
hit  the  ground  . . . ,”  and  by  Bristowe  and  Locket  (1926),  who  had 
reported  earlier  on  the  same  species,  but  had  implicated  the  legs  II 
as  the  sound  producers.  In  either  case,  it  was  not  clear  if  the  sound 
produced  by  E.  frontalis  was  an  integral  part  of  the  courtship  or 
incidentally  produced  by  the  movement  of  the  legs. 


'Contribution  No.  514,  Bureau  of  Entomology,  Division  of  Plant  Industry,  Florida 
Department  of  Agriculture  and  Consumer  Services,  Gainesville,  FL  32602. 

2Florida  State  Collection  of  Arthropods,  Division  of  Plant  Industry,  P.  O.  Box  1269, 
Gainesville,  FL  32602. 

Manuscript  received  by  the  editor  June  7,  1981. 


199 


200 


Psyche 


[Vol.  88 


Experimental  Procedure 

Four  females  and  2 males  of  P.  mystaceus  were  reared  to  maturity 
from  an  eggsac  containing  12  eggs.3  The  spiders  were  housed 
separately  in  9 x 1 cm  plastic  petri  dishes;  twice  a week  they  were 
provided  with  water  by  moistening  a wad  of  cotton  within  the  dish 
and  were  fed  larvae  of  the  cabbage  looper,  Trichoplusia  ni  (Hiibner). 

Two  different  techniques  were  used  for  observing  courtship.  In 
one  method,  the  male  was  placed  directly  into  a female’s  petri  dish, 
on  the  side  opposite  the  female.  In  the  other  method,  the  male  and 
female  were  placed  5-15  cm  apart  on  a 30  x 10  cm  section  of  a live- 
oak  branch,  in  order  to  simulate  natural  conditions.  Temperature 
ranged  from  24-26°  C for  all  sessions. 

Six  separate  filming  and/or  recording  sessions  lasted  10-90  min. 
each.  Films  were  made  using  a Beaulieu  Super-8-mm  movie  camera 
and  an  Auricon  Pro  600  16-mm  movie  camera.  Sound  recordings 
were  made  with  a Sony  TC-756-2  reel-to-reel  tape  recorder  and  a 
Turner  S22D  microphone.  The  audiospectrogram  was  produced  on 
a Kay  7029A  Sound  Spectrograph. 

Results 

Courtships  were  observed  for  one  of  the  males  (the  second  male 
was  killed  by  the  first  female  with  which  he  was  placed).  Typically  a 
male  placed  into  the  petri  dish  housing  a female  almost  immediately 
begins  palpating  the  female’s  draglines  and  her  abandoned  nests, 
continuing  this  palpal  exploration  until  he  detects  the  female 
visually. 

If  the  female  is  not  inside  a nest  when  first  seen  by  the  male 
(usually  from  3-6  cm),  the  male  begins  producing  a soft,  audible  trill 
that  is  systematically  repeated.  By  apparently  engaging  the  substrate 
with  enlarged  setae  (macrosetae,  Fig.  1;  similar  to  those  observed  on 
lycosids  by  Rovner,  1975),  leverage  is  produced  enabling  a stridu- 
latory  mechanism  on  the  palpus  to  be  operated.  This  mechanism 
consists  of  a plectrum-like  projection  of  the  tibial  apophysis  which 
fits  into  a bowl-shaped  area  on  the  cymbium  containing  a compli- 
cated file  system.  The  entire  mechanism  is  located  laterally  (ectally); 
in  lycosids  it  is  located  dorsally.  Also,  lycosids  have  the  file  on  the 

3Gravid  female  P.  mystaceus  collected  by  Robert  Dye,  26  October  1975, 4 miles  north 
of  Texas  state  line  at  a rest  stop  on  1-35  in  Oklahoma,  under  a rock.  Eggs  were  laid 
November,  1975. 


1981] 


Edwards  — Phidippus  mystaceus 


201 


tibia  facing  a cymbial  plectrum,  the  reverse  of  the  condition  in  P. 
mystaceus.  The  file  system  of  P.  mystaceus  appears  to  consist  of  2 
types  of  adjacent  file  fields  which  blend  into  one  another.  Within  the 
concavity  is  a fan-shaped  file,  while  along  the  distal  edge  of  the 
concavity  is  a linearly-arranged  file  similar  to  lycosid  files.  Neither 
file  is  as  well-defined  as  the  lycosid  files.  The  individual  ridges  of  P. 
mystaceus ’ files  are  rounded,  whereas  those  of  lycosids  have  distinct 
edges;  however,  in  P.  mystaceus,  both  types  of  file  are  overlaid  with 
numerous  short  ridges  of  variable  length  (Fig.  2). 


Fig.  1.  Distal  tip  of  palpus  of  male  P.  mystaceus  showing  ring  of  macrosetae  (M) 
encircling  whorled  chemotactile  setae  (W).  On  extreme  left  are  scale-like  setae  (S) 
which  form  part  of  a white  and/or  yellow  spot  which  probably  contributes  to  the 
overall  visual  stimulus  of  a courting  male  (100X).  Note  the  greater  number  of 
macrosetae  on  the  ectal  edge  (E);  see  text  for  explanation.  The  curved  macrosetae 
(extreme  right)  at  the  tip  of  the  cymbium  first  contact  the  substrate  and  may  facilitate 
the  backwards  sliding  motion  of  the  palpi  by  reducing  friction  with  the  substrate. 


202 


Psyche 


[Vol.  88 


Fig.  2.  Left:  stridulatory  area  of  male  P.  mystaceus  on  ectal  edge  of  cymbium  of  left  palpus,  showing  fan-shaped  file  (F),  linear 
file  (L),  and  plectrum  (P),  which  is  a branch  of  the  tibial  apophysis  (T)  (250X);  Right:  closeup  of  portion  of  file  indicated  by 
arrow  (1300X). 


1981] 


Edwards  — Phidippus  my  s face  us 


203 


For  each  sound  sequence,  both  palpi  become  engaged  nearly 
simultaneously  by  a backward  movement  in  which  the  palpi  appear 
to  be  dragged  along  the  surface  of  the  substrate  for  a distance  of 
about  1 mm.  Halfway  through  the  backward  movement,  the  cymbia 
are  bent  backward  at  an  angle  to  the  palpal  tibiae  (Fig.  3).  At  the 
end  of  the  backward  movement,  the  palpi  remain  stationary  for  a 
fraction  of  a second  while  the  cymbia  are  rotated  outward  (left 
palpus  clockwise,  right  palpus  counter  clockwise).  The  palpi  are 
then  returned  to  their  most  anterior  position,  apparently  by  lifting 
the  palpi  from  the  substrate  and  moving  them  forward.  When  the 
palpi  are  in  their  most  anterior  position,  they  are  clearly  off  the 
substrate.  A single  cycle  of  palpal  movement  is  approximately  0.8 
second  (5  frames  at  6 frames  per  second). 

Audiospectrograms  indicate  that  13-20  (x  = 17,  n = 8)  paired 
stridulations  are  made  consecutively,  separated  by  pauses  subequal 
in  timing  to  the  sound  sequences  (Fig.  4).  Alternation  of  stridula- 
tions and  pauses  occurs  at  the  rate  of  1.5  sound  sequences  per 
second  (at  approximately  25°  C). 


Fig.  3.  Diagrammatic  illustration  of  movement  of  left  palpus  (ectal  view)  by  male 
P.  mystaceus  during  stridulation.  A.  Anterior  position.  B.  Backward  movement, 
during  which  cymbium  is  bent  backward,  moving  fan-shaped  file  across  plectrum.  C. 
Rotary  movement,  during  which  macrosetae  are  engaged  in  substrate  and  cymbium 
is  rotated  outward,  moving  linear  file  across  plectrum.  F = File  cavity,  T = Tibial 
apophysis. 


204 


Psyche 


[Vol.  88 


Simultaneous  with  the  initiation  of  sound  production,  the  male 
extends  his  legs  I forward,  positioning  them  just  above  and  parallel 
to  the  substrate,  and  spread  approximately  40°  apart.  The  tarsi  and 
metatarsi  are  turned  upward  about  15°  and  occasionally  flicked 
upward  together.  On  1 occasion,  at  a distance  of  about  1 cm  from 
the  female,  the  tarsi  and  metatarsi  were  flicked  continuously  for 
several  seconds  at  approximately  2 flicks  per  second. 

The  male’s  approach  usually  is  direct,  without  the  zigzag  move- 
ment (lateral  stepping  movement)  characteristic  of  some  other 
Phidippus  species  and  many  other  salticids.  Forward  movement  is 
slow  and  halting,  the  male  often  remaining  in  one  spot  for  several 
minutes.  Total  courtship  time  is  long  compared  to  the  rapid  advance 
of  the  males  of  some  Phidippus  species,  on  3 occasions  lasting 
approximately  8 minutes  before  the  female  terminated  the  courtship 
by  leaving  the  vicinity.  These  3 longest  courtships  reached  an 
advanced  stage,  wherein  the  male  brought  his  legs  I closer  together, 
touched  the  female,  and  attempted  to  mount  her;  however,  none  of 
the  4 females  allowed  their  male  sibling  to  mate  with  them.  Instead, 
each  raised  her  legs  I to  repel  him,  and,  if  the  male  was  persistent, 
lunged  sharply  forward  with  open  fangs,  struck  downward  with  the 
legs  I,  and  forced  him  backward;  the  female  then  left  the  vicinity. 

On  2 occasions,  the  male  performed  a zigzag  display;  once  prior 
to  assuming  his  stridulatory  stance,  and  once  in  the  middle  of 
courtship  after  several  sequences  of  stridulation.  In  the  first  in- 
stance, the  zigzag  display  was  brief,  lasting  less  than  30  seconds  and 
consisting  of  4 changes  of  direction,  with  a pause  between  each 
lateral  move.  In  the  second  instance,  during  mid-courtship,  7 
multiple  zigzags  occurred  which  included  1-3  changes  of  direction 
during  each  lateral  stepping  sequence;  total  elapsed  time  was  about 
3 minutes. 

If  the  female  is  initially  in  and  remains  in  a nest  when  the  male  is 
introduced  into  the  petri  dish,  the  male  alternates  palpation  of  the 
substrate  with  sequences  of  stridulation.  Upon  finding  the  nest 
sheltering  the  female,  the  male  attempts  to  gain  entrance  by  probing 
and  pulling  at  the  silk  with  his  legs  I,  interspersing  sequences  of 
palpal  vibration  on  the  silk.  (Note:  other  species  of  Phidippus 
known  to  use  a tactile  courtship  vibrate  their  entire  body).  I could 
not  determine  the  movement  pattern  of  this  palpal  vibration  (it 


1981] 


Edwards 


Phidippus  mystaceus 


205 


previously  reported  for  any  other  spider. 


206 


Psyche 


[Vol.  88 


appeared  to  be  similar  in  timing  to  stridulation),  but  the  palpi  were 
not  in  contact  with  the  petri  dish,  and  no  audible  sound  was 
produced. 


Discussion 

Known  reproductive  behavior  of  the  males  of  species  of  Phidip- 
pus  involves  a male  locating  a female  by  visual  or  chemotactic 
means  (Richman,  in  press),  a visually-oriented  courtship  by  the 
male  consisting  of  a series  of  movements  with  the  legs  I and  palpi 
(usually  while  advancing  in  a zigzag  path),  mounting  of  the  female 
by  the  male,  and  mating.  Typically  the  male  is  conspicuously 
marked  with  bright  and/or  contrasting  colors  both  anteriorly  and 
dorsally;  the  anterior  patterns  are  displayed  during  courtship. 
Unlike  most  other  species,  both  males  and  females  of  P.  mystaceus 
are  cryptically-colored  gray  spiders  that  live  in  trees  (Specht  and 
Dondale,  1960;  Warren  et  al.,  1967,  as  P.  incertus;  see  Edwards, 
1977,  for  nomenclatorial  comments);  males  have  mostly  anteriorly- 
oriented  modifications  (Fig.  5).  While  anterior  modifications  are 
probably  used  by  each  sex  to  identify  the  other  (especially  the 
female  recognizing  the  male  as  a conspecific  and  potential  mate) 
from  distances  of  a few  centimeters,  visual  identification  at  longer 
distances  might  be  severely  handicapped  by  cryptic  coloration.  A 
mechanism  which  increases  the  chance  of  one  sex  locating  the  other 
could  be  selected  for  under  these  circumstances. 

The  role  of  acoustic  or  vibratory  signals  in  the  courtship  of  P. 
mystaceus  may  have  co-evolved  with  cryptic  coloration.  As  selec- 
tion for  cryptic  coloration  increased  in  association  with  exploitation 
of  a new  microhabitat  (most  Phidippus  species  live  in  the  herb- 
shrub  zone),  the  role  of  visual  communication  might  have  been  in 
part  supplanted  by  sound  during  courtship.  The  use  of  sound, 
whether  airborne  or  substrate-borne,  would  have  several  advantages 
over  conventional  visual  courtship,  if  the  sound  extended  the  male’s 
communicatory  distance  from  a few  centimeters  to  over  a meter  (as 
it  appears  to  do  based  on  the  audible  component  available  to  the 
human  ear).  Sound  is  transmitted  well  through  solids,  and  consider- 
ing that  in  this  case  sound  is  produced  on  the  substrate,  vibrations 
through  this  medium  may  be  most  important  for  female-to-male 
orientation  (as  Rovner,  1967,  showed  to  be  the  case  for  wolf 
spiders).  By  orienting  toward  the  male  upon  perception  of  the 


1981] 


Edwards  — Phidippus  mystaceus 


207 


Fig.  5.  Anterior  views  of  male  and  female  P.  mystaceus,  offspring  of  female 
collected  in  Oklahoma,  which  were  used  in  this  experiment. 


208 


Psyche 


[Vol.  88 


sound,  the  female  might  sooner  visually  detect  and  be  able  to 
evaluate  the  male  as  a prospective  mate,  and  thus  sooner  choose  to 
wait  for  or  flee  from  him.  The  advantages  gained  by  the  male  by 
increasing  his  communicatory  distance  might  be:  1)  alerting  a 
receptive  female  to  his  presence  at  a greater  distance,  possibly 
causing  her  to  remain  in  the  vicinity  for  a longer  period  of  time  (and 
perhaps  inhibiting  her  predatory  instincts),  so  that  the  male  has  a 
greater  chance  of  finding  and  courting  her;  2)  based  on  many 
observations  of  P.  mystaceus  and  of  other  Phidippus  species,  non- 
receptive  females  usually  avoid  advancing  males;  thus,  by  alerting  a 
female  to  his  presence  at  a greater  distance,  a male  would  reduce  the 
chance  of  stimulating  an  aggressive  response  by  a non-receptive 
female. 

As  evidence  for  these  probable  advantages,  analysis  of  courtships 
showed  that  the  male  began  courting  a female  from  2-4  times 
further  away  when  unconfined  (on  the  live-oak  branch)  than  did 
other  species  of  Phidippus  when  observed  under  unconfined  experi- 
mental conditions  (Edwards,  1975).  At  the  greater  distances,  only 
sound  was  used  initially  by  a P.  mystaceus  male  upon  sighting  a 
female,  indicating  that  this  form  of  communication  was  important 
in  alerting  a female  to  his  presence.  Sound  was  also  used  alternately 
with  palpal  exploration  of  the  silk  when  the  male  was  in  contact 
with  the  female’s  draglines  in  the  petri  dish,  even  though  she  was  not 
visible.  Under  natural  conditions,  a male  likely  would  often  en- 
counter a female’s  dragline  prior  to  locating  her;  he  could  maximize 
his  chances  of  mating  by  beginning  to  signal  immediately,  regardless 
of  whether  or  not  the  female  was  visible  to  him. 

Components  of  Behavior  and  Morphology 

The  male’s  initial  palpal  exploration  of  the  female’s  silk  draglines 
and  nests  has  been  noted  for  other  salticids  (Richman,  1977).  The 
presence  of  a contact  pheromone  on  the  silk  could  indicate  to  a male 
that  a female  was,  or  had  been,  in  the  vicinity.  Contact  pheromones 
(Hegdekar  and  Dondale,  1969)  and  dragline  following  by  males 
(Tietjen  and  Rovner,  1980)  have  been  demonstrated  for  some 
lycosids,  but  have  not  yet  been  conclusively  demonstrated  for  any 
salticid.  Foelix  (1970)  demonstrated  the  presence  of  chemosensitive 
setae  in  certain  araneid  spiders  and  hypothesized  that  those  setae 
were  contact  chemoreceptors.  He  showed  that  the  suspected  chemo- 


1981] 


Edwards  — Phidippus  mystaceus 


209 


sensitive  setae  in  araneids  were  innervated  and  structured  in  essen- 
tially the  same  manner  as  pheromone  receptors  of  insects.  Hill 
(1977a,  b)  noted  that  the  whorled  setae  on  the  tarsi  and  palpal 
cymbia  of  several  species  of  Phidippus  also  resembled  insect 
pheromone  receptors;  male  P.  mystaceus  have  the  same  type  of 
setae  on  their  palpal  cymbia  (Fig.  1). 

The  behavior  in  P.  mystaceus  of  engaging  the  palpi  against  the 
substrate  is  probably  derived  from  similar  behavior  among  its 
relatives.  Males  of  other  species  of  Phidippus  move  their  palpi  up 
and  down  or  back  and  forth  during  courtship.  This  behavior 
appears  to  pre-adapt  them  for  engaging  the  substrate,  since  only  a 
slight  change  in  the  amplitude  and / or  attitude  of  these  movements 
would  bring  the  palpi  into  contact  with  the  substrate.  The  same 
movement  occurs  more  intensely  and  rapidly  when  a male  encoun- 
ters silk  made  by  a female,  in  association  with  presumed  chemotac- 
tile  exploration;  it  is  likely  that  this  is  the  evolutionary  pathway  of 
the  development  of  the  use  of  sound  in  P.  mystaceus. 

The  shape  and  arrangement  of  the  macrosetae  at  the  tip  of  the 
cymbium  are  such  that  a downward,  forward  pressure  would  engage 
them  with  the  substrate.  By  dragging  the  palpus  backward,  enough 
leverage  apparently  is  produced  to  move  the  fan-shaped  file  across 
the  relatively  stationary  plectrum;  however,  the  backward  move- 
ment and  bend  of  the  palpus  also  may  be  a prerequisite  to 
positioning  the  macrosetae  onto  the  substrate.  Once  the  palpus  is 
anchored  onto  the  substrate,  the  cymbium  is  rotated  laterally 
outward,  then  the  palpi  are  returned  to  their  starting  position.  The 
macrosetae  are  arranged  in  a circle  around  the  tip  of  the  cymbia  in 
P.  mystaceus,  with  more  macrosetae  on  the  ectal  edge  than  on  the 
ental  edge,  which  enables  the  palpus  to  remain  engaged  with  the 
substrate  as  it  rotates  outward.  The  structure  involved  in  sound 
production  by  rotating  is  the  linear  file;  the  cymbium  must  be 
rotated  sideways  due  to  the  lateral  position  of  the  file. 

By  simulating  the  direction  of  palpal  movement  with  a model,  it  is 
apparent  that  the  backward  movement  would  cause  the  fan-shaped 
file  to  be  drawn  across  the  plectrum,  while  the  rotary  movement 
would  bring  the  linear  file  into  contact  with  the  plectrum.  The  fan 
shape  of  the  proximal  file  would  accommodate  the  arc-shaped 
movement  as  the  palpus  is  bent  on  the  backward  stroke;  however, 
sound  does  not  seem  to  be  produced  by  the  fan-shaped  file.  Only 


210 


Psyche 


[Vol.  88 


one  type  of  stridulation  is  produced,  evidently  from  the  linear  file 
(see  figure  4);  the  function  of  the  fan-shaped  file  remains  unclear. 
Although  the  files  appear  to  be  oriented  so  that  they  could  be 
stroked  from  either  direction,  the  timing  of  a complete  palpal 
movement  indicates  that  sound  is  produced  only  on  the  backstroke 
and  not  on  the  return  stroke.  The  mechanics  of  stridulation  by  P. 
mystaceus  are  still  incompletely  understood,  and  need  further  study 
with  more  sophisticated  filming  techniques. 

Comparisons  to  Other  Stridulatory  Mechanisms 

The  behavioral  application  of  the  palpi  to  the  substrate  by  P. 
mystaceus  differs  from  lycosids  in  that  P.  mystaceus  moves  the  tips 
of  the  palpi  while  stridulating  during  each  brief  sound  sequence, 
whereas  the  lycosids  apparently  remain  attached  in  one  place  to  the 
substrate  for  a prolonged  sequence  of  sound  production.  The 
mechanics  of  sound  production  with  the  linear  file  are  similar  to 
those  of  the  lycosids  with  respect  to  the  palpus  anchored  by 
macrosetae  and  the  similar  file  structure,  but  P.  mystaceus  differs 
from  the  lycosids  in  the  location  of  the  stridulatory  organ,  the  type 
of  movement  needed  to  engage  the  file,  and  the  reversed  positions  of 
the  file  and  plectrum.  Rovner  (1975)  proposed  a new  category  of 
stridulatory  organ  (as  an  extension  of  the  classification  of  Legendre, 
1963),  type  “h”  to  accommodate  those  types  of  mechanisms  in  which 
the  file  and  plectrum  (scraper)  were  on  adjacent  surfaces  of  a joint 
within  the  same  appendage.  I propose  a subdivision  of  Rovner’s 
category,  following  Legendre’s  method  of  subdividing  categories: 
type  “h  I”  in  which  the  file  is  on  the  more  proximal  segment  (as  in 
lycosids),  and  type  “h  II”  in  which  the  file  is  on  the  more  distal 
segment  (as  in  P.  mystaceus). 

The  stridulatory  mechanisms  known  in  other  spiders  incorporate 
plectrum  and  file  systems  on  opposing  faces  of  the  chelicerae  and 
palpi,  legs  I and  II,  carapace  and  legs  I,  carapace  and  abdomen 
(Gertsch,  1979),  or  between  palpal  tibia  and  tarsus  (Rovner,  1975). 
In  each  of  these  cases,  either  the  plectrum  is  moved  across  a 
stationary  file  or  both  plectrum  and  file  are  moved  together.  The 
stridulatory  mechanism  of  P.  mystaceus  differs  from  all  of  these  in 
that  the  primary  moving  part  is  the  file.  Although  the  plectrum  is 
passively  moved  in  space  during  the  movement  of  the  palpus  to 


1981] 


Edwards  — Phidippus  mystaceus 


211 


engage  the  substrate,  the  cymbium  containing  the  files  is  actively 
moved  against  the  plectrum  on  the  tibia.  When  the  palpus  is  fixed 
on  the  substrate  with  the  macrosetae,  again  it  is  the  cymbium  that  is 
moved  against  the  stationary  plectrum. 

Other  Types  of  Vibratory  Signaling 

A third  method  of  sound  production  in  spiders,  vibration  (pro- 
ducing a “buzz”  similar  to  that  of  a fly),  has  been  demonstrated  for 
the  sparassid  spider,  Heteropoda  venatoria  (L.)  (Rovner,  1980).  In 
the  same  paper,  low  amplitude  appendage  oscillations  resulting  in  a 
faint  whirring  sound  were  reported  for  Lycosa  rabida  Walckenaer. 
Phidippus  whitmani  Peckham  and  Peckham  employs  entire-body 
(?)  vibration  (lacking  an  audible  component,  but  with  a widely- 
spaced  stance  similar  to  H.  venatoria ) during  its  Type  I visual 
courtship  (Edwards,  1980).  This  is  probably  an  adaptation  to  its 
microhabitat  (mesophytic  leaf  litter),  the  same  substrate  used  for 
vibratory  signaling  by  many  lycosids.  I have  noted  another  vibra- 
tory behavior  that  also  seems  similar  to  that  of  H.  venatoria  during 
the  Type  II  tactile  courtship  of  Phidippus  regius  C.  L.  Koch,  while 
the  male  is  contacting  the  nest  of  the  female  (Edwards,  1975). 
Subsequent  laboratory  observation  showed  a similar  behavior  for 
P.  whitmani,  although  the  timing  of  vibratory  sequences  was 
different  from  those  of  P.  regius,  probably  a species-specific  differ- 
ence. Jackson  (1977)  reported  a similar  behavior  for  P.  johnsoni 
(Peckham  and  Peckham)  and  suggested  a similarity  in  some  respects 
to  the  vibratory  courtships  of  web-building  spiders.  I suspect  that 
the  vibratory  courtships  of  Phidippus  species,  although  not  pro- 
ducing an  audible  component  that  I could  detect,  may  be  more  like 
the  courtship  of  H.  venatoria  than  like  web-builders,  or  perhaps  all 
3 groups  produce  vibrations  in  essentially  the  same  way  (i.e., 
“juddering”  as  in  araneid  males;  Robinson  and  Robinson,  1980).  It 
is  curious  that  all  known  forms  of  non-tactile  direct  inter-individual 
communication  not  involving  vision  in  salticids  are  acoustic  or 
vibratory  signals  (despite  the  contention  of  Crane,  1949,  and  other 
authors,  the  use  of  airborne  pheromones  by  salticids  has  never  been 
proven).  In  the  case  of  P.  mystaceus  and  P.  whitmani,  both  visual 
and  vibratory  signals  are  used  simultaneously,  although  the  2 
species  produce  vibrations  in  different  ways. 


212 


Psyche 


[Vol.  88 


Conclusion 

The  use  of  stridulation  to  produce  sound  by  P.  mystaceus  appears 
to  represent  a third  method  of  communication  for  salticids  (a  fourth 
method,  if  the  tarsal  percussion  of  Euophrys  frontalis  is  a valid 
communicatory  process).  Despite  the  fact  that  females  used  for  the 
present  research  failed  to  respond  favorably  to  courtship  by  their 
sibling  male,  the  behavioral  and  morphological  evidence  in  the  male 
of  a functional  role  for  sound  production  during  courtship  is 
substantial.4 


Summary 

Males  of  Phidippus  mystaceus  have  a stridulatory  organ  located 
on  the  tarsal  and  tibial  segments  of  the  palpi.  This  organ  is 
employed  by  males  in  the  potential  or  actual  presence  of  adult 
females,  and  forms  the  most  significant  part  of  courtship  by  males. 
The  mechanics  of  stridulation  are  somewhat  similar  to  those  of 
lycosids,  and  as  with  the  lycosids,  substrate  vibrations  may  be  the 
most  important  component  of  stridulation.  Evolution  of  sound 
production  by  P.  mystaceus  is  hypothesized  to  have  occurred  in 
conjunction  with  the  evolution  of  cryptic  coloration.  Sound  produc- 
tion is  thought  to  extend  the  males’  communicatory  distance, 
compensating  for  fewer  visual  identification  opportunities  due  to 
the  spiders’  cryptic  coloration. 

Acknowledgments 

I would  like  to  give  special  thanks  to  the  following:  Dr.  Robert 
Paul  for  his  help  in  sound  recording  and  producing  the  audiospec- 
trogram; Dr.  Jonathan  Reiskind  for  the  S.  E.  M.  photomicrographs. 
Thanks  are  also  due;  Mr.  John  Thorne  and  Mr.  Stan  Blomely  for 

4Two  antepenultimate  P.  mystaceus  were  collected  by  G.  B.  Edwards,  28  July  1979, 
Ocala  National  Forest,  Marion  Co.,  Florida,  beating  young  scrub  live  oaks,  and 
reared  to  maturity  (October,  1979).  Although  these  specimens  were  collected  and 
reared  after  the  research  on  the  Oklahoma  specimens  was  completed,  and  the 
courtship  was  neither  filmed  nor  recorded,  a courtship  and  mating  was  observed  for 
this  pair.  Courtship  appeared  in  all  respects  to  be  identical  to  that  of  the  Oklahoma 
male,  including  type  of  sound,  stance,  and  the  rapid  upward  flicking  of  the  tarsi  and 
metatarsi  at  less  than  1 cm  distance  from  the  female.  Mating  occurred  in  the  female’s 
nest  and  lasted  87  minutes  until  the  female  left  the  nest.  Upon  separating,  the  male 
renewed  courtship,  initially  showing  a single  lateral  stepping  sequence  as  in  the 
Oklahoma  male.  The  female  avoided  the  male,  and  the  pair  was  separated. 


1981] 


Edwards  — Phidippus  mystaceus 


213 


aid  in  filming  (16-mm);  Mr.  Lloyd  R.  Davis,  Jr.,  for  obtaining  the 
gravid  P.  mystaceus  female  for  me;  and  Drs.  Jerome  S.  Rovner, 
Jonathan  Reiskind,  Thomas  J.  Walker,  and  Robert  L.  Crocker  for 
reviewing  the  manuscript. 

Literature  Cited 


Bristowe,  W.  S. 

1958.  The  World  of  Spiders.  Collins,  London.  305  p. 

Bristowe,  W.  S.,  and  G.  H.  Locket. 

1926.  The  courtship  of  British  lycosid  spiders,  and  its  probable  significance. 
Proc.  Zool.  Soc.  London  1926(2):  317-347. 

Crane,  J. 

1949.  Comparative  biology  of  salticid  spiders  at  Rancho  Grande,  Venezuela. 
Part  IV.  An  analysis  of  display.  Zoologica  34(4):  159-215. 

Edwards,  G.  B. 

1975.  Biological  studies  on  the  jumping  spider,  Phidippus  regius  C.  L.  Koch. 

M.S.  Thesis,  University  of  Florida.  64  p. 

1977.  Comments  on  some  genus  and  species  problems  in  the  Salticidae, 
including  Walckenaerian  names.  Peckhamia  1(2):  21-23. 

1980.  Taxonomy,  ethology,  and  ecology  of  Phidippus  (Araneae:  Salticidae)  in 
eastern  North  America.  Ph.D.  Dissertation,  University  of  Florida.  354  p. 
Foelix,  R.  F. 

1970.  Chemosensitive  hairs  in  spiders.  J.  Morphol.  132:  314-334. 

Gertsch,  W.  J. 

1979.  American  Spiders.  Van  Nostrand  Reinhold  Co.,  Second  Edition,  New 
York.  274  p. 

Hegdekar,  B.  M.,  and  C.  D.  Dondale. 

1969.  A contact  sex  pheromone  and  some  response  parameters  in  lycosid 
spiders.  Canad.  J.  Zool.  47(1):  1-4. 

Hill,  D.  E. 

1977a.  The  pretarsus  of  salticid  spiders.  Zool.  J.  Linnean  Soc.  London  60(4): 
319-338. 

1977b.  Modified  setae  of  the  salticid  pedipalp.  Peckhamia  1(1):  7-8. 
Jackson,  R.  R. 

1977.  Courtship  versatility  in  the  jumping  spider,  Phidippus  johnsoni  (Ara- 
neae: Salticidae).  Anim.  Behav.  25(4):  953-957. 

Legendre,  R. 

1963.  L’audition  et  remission de  sons  chez  les  araneides.  Ann.  Biol.  2:  371-390. 
Petrunkevitch,  A. 

1926.  Spiders  of  the  Virgin  Islands.  Trans.  Connecticut  Acad.  Arts  Sci.  28: 
21-78. 


Richman,  D.  B. 

1977.  The  relationship  of  epigamic  display  to  the  systematics  of  jumping 
spiders  (Araneae:  Salticidae).  Ph.D.  Dissertation,  University  of  Florida. 

162  p. 


214 


Psyche 


[Vol.  88 


1982.  Epigamic  display  in  jumping  spiders  (Araneae:  Salticidae)  and  its  use  in 
systematics.  J.  Arachnol.  10  (in  press). 

Robinson,  M.  H.,  and  Barbara  Robinson. 

1980.  Comparative  studies  of  the  courtship  and  mating  behavior  of  tropical 
araneid  spiders.  Pacific  Insects  Monograph  36:  1-218. 

Rovner,  J.  S. 

1967.  Acoustic  communication  in  a lycosid  spider  ( Lycosa  rabida  Walck- 
enaer).  Anim.  Behav.  15:  273-281. 

1975.  Sound  production  by  Nearctic  wolf  spiders:  A substratum-coupled 
stridulatory  mechanism.  Science  190:  1309-1310. 

1980.  Vibration  in  Heteropoda  venatoria  (Sparassidae):  A third  method  of 
sound  production  in  spiders.  J.  Arachnol.  8:  193-200. 

Specht,  H.  B.,  and  C.  D.  Dondale. 

1960.  Spider  populations  in  New  Jersey  apple  orchards.  J.  Econ.  Ent.  53: 
810-814. 

Tietjen,  W.  J.,  and  J.  S.  Rovner. 

1980.  Trail-following  behaviour  in  two  species  of  wolf  spiders:  sensory  and 
etho-ecological  concomitants.  Anim.  Behav.  28:  735-741. 

Warren,  L.  O.,  W.  B.  Peck,  and  M.  Tadic. 

1967.  Spiders  associated  with  the  fall  webworm,  Hyphantria  cunea  (Lepidop- 
tera:  Arctiidae).  J.  Kansas  Ent.  Soc.  40:  382-395. 


MATERNAL  BEHAVIOR  AND  ALARM  RESPONSE  IN 
THE  EGGPLANT  LACE  BUG,  GARGAPHIA  SOLANI 
HEIDEMANN  (TINGIDAE:  HETEROPTERA)1 

By  R.  S.  Kearns2  and  R.  T.  Yamamoto 

Entomology  Department 
North  Carolina  State  University 
Raleigh,  NC  27650 


INTRODUCTION 

Maternal  behavior  in  the  eggplant  lace  bug,  Gargaphia  solani 
Heidemann  (Tingidae:  Heteroptera)  was  first  reported  by  Fink 
(1915).  He  described  the  female’s  guarding  of  the  eggs  and  shepherd- 
ing of  the  nymphs  from  leaf  to  leaf.  G.  solani  is  found  on  the  native 
horse  nettle  ( Solanum  carolinense)  and  on  the  introduced  eggplant 
(Solanum  melongena).  Overwintering  adults  appear  in  late  spring, 
and  females  lay  eggs  in  circular  masses  on  the  underside  of  leaves. 
Fink  reported  that  the  number  of  eggs  is  greater  than  100,  oviposi- 
tion  lasts  4 to  5 days,  and  the  incubation  period  is  about  6 days. 
Maternal  care  persists  through  the  development  of  the  nymphs,  and 
the  life  cycle  is  approximately  20  days.  Females  observed  in  this 
study  usually  laid  less  than  100  eggs  over  a period  of  3 to  4 days 
(Kearns  1980). 

Maternal  behavior  has  been  reported  for  a number  of  heterop- 
terans  (Melber  and  Schmidt  1977)  and  for  two  other  species  of  the 
genus  Gargaphia:  Gargaphia  tiliae  (Weiss  1919,  Torre-Bueno  1935, 
Sheeley  and  Yonke  1977)  and  Gargaphia  irridescens  (Torre-Bueno 
1942).  These  accounts  give  few  details.  The  maternal  behavior  of  G. 
solani  has  much  in  common  with  that  exhibited  by  treehoppers 
(Membracidae:  Homoptera)  (Wood  1974,  1976a,  1976b,  1977  and 
Hinton  1977).  The  complex  behavior  patterns  of  membracids  and  of 
G.  solani  suggest  that  aggregations  of  these  insects  depend  upon  a 


•Paper  number  6950  of  the  Journal  Series  of  the  North  Carolina  Agricultural 
Research  Service,  Raleigh,  North  Carolina. 

2This  work  was  completed  in  partial  fulfillment  of  the  requirements  for  the  degree  of 
Master  of  Science  in  Entomology. 

Manuscript  received  by  the  editor  August  30,  1981. 


215 


216 


Psyche 


[Vol.  88 


pheromonal  communication  system  which  facilitates  group  move- 
ments. In  the  membracids,  there  is  indirect  evidence  for  aggregation 
pheromones  (Hinton  1976,  1977).  Alarm  pheromones  of  membra- 
cids are  released  only  when  the  body  wall  is  ruptured,  and  this  type 
of  release  has  not  been  reported  for  any  other  insects.  Pheromones 
causing  alarm  responses  are  known  to  be  present  in  at  least  3 
membracid  species.  They  are  interspecific  in  action  but  have  not 
been  identified  (Nault  et  al.  1974). 

This  study  reports  an  examination  of  the  movements  of  G.  solani 
aggregations  on  host  plants  with  particular  emphasis  on  the  female’s 
behavior. 

MATERIALS  AND  METHODS 

MAINTENANCE  OF  INSECTS 

G.  solani  was  collected  from  horse  nettle  growing  in  or  near 
Raleigh,  N.C.,  and  aggregations  were  maintained  for  more  than  a 
year  on  horse  nettle  or  eggplant  either  in  the  laboratory  on  a 16:8 
light:  dark  cycle  or  in  a greenhouse.  The  movements  of  nymphs  were 
studied  after  an  aggregation  consisting  of  nymphs  and  a female  had 
been  transferred  to  a small  sprig  of  horse  nettle  having  an  un- 
branched stem  with  5 or  more  leaves.  A piece  of  leaf  containing  a 
group  was  pinned  to  the  upper  surface  of  the  second  or  third  leaf 
from  the  bottom  of  a horse  nettle  stem.  In  time,  the  aggregation 
moved  off  the  leaf  fragment  and  onto  the  fresh  leaf. 

FEEDING  MOVEMENTS 

Feeding  movements  were  measured  in  light,  darkness,  and  with  a 
light  placed  below  the  aggregation.  Groups  chosen  for  these  studies 
were  nymphs  in  the  third  or  fourth  instars,  with  smaller  numbers  of 
the  other  instars  present.  For  dark  conditions,  an  aggregation  on  a 
horse  nettle  sprig  was  placed  in  a tightly  covered  metal  can  which 
had  been  sprayed  inside  with  a dull-finish  black  paint.  Directional 
lighting  was  provided  by  placing  the  sprig  or  plant  within  a 
darkened  enclosure  and  positioning  a light  about  0.6  m from  the 
bottom  of  the  plant. 

ALARM  RESPONSE  AND  ALARM  PHEROMONE 

G.  solani  nymphs  exhibited  an  alarm  response  after  they  were 
presented  with  a nymph,  freshly  squeezed  and  held  by  fine  forceps, 


1981] 


Kearns  & Yamamoto  — Gargaphia 


217 


Table  1 . Elements  of  behavior  of  G.  solani  females  when  their  broods  were  responding 


to  alarm  pheromone 

% of  responses 

Behavior 

observed  (1) 

1. 

Female  positioned  slightly  below  exit  axil 

17 

2. 

Female  positioned  at  exit  axil 

50 

3. 

Female  positioned  between  axils 

39 

4. 

Female  positioned  at  entrance  axil 

61 

5. 

Female  moved  off  leaf  shortly  after  nymphs  left 

78* 

6. 

Female  followed  nymphs  onto  new  leaf 

78* 

7. 

Female  returned  to  old  leaf  after  group  had  moved 

33 

8. 

Female  used  at  least  one  of  the  first  four  elements  listed 

89* 

(1)  18  different  trials 

* Significant  at  the  95%  confidence  level  (Binomial  distribution.  Table  2,  partial  sums, 
Eisenhart  1952.  Confidence  intervals,  Table  A-22,  Natrella  1963.) 


or  a nymph  pierced  with  a pin  against  a small  disk  of  filter  paper 
(after  Nault  et  al.  1974).  Blank  disks  of  filter  paper  placed  near  an 
aggregation  did  not  elicit  an  alarm  response.  Nymphs  of  G.  solani 
were  also  collected  and  stored  in  a small  container  of  chloroform. 
Several  hundred  were  sufficient  to  provide  a crude  extract.  Before 
the  extract  was  tested  for  activity,  the  chloroform  was  allowed  to 
evaporate  from  a small  sample  contained  in  a Pasteur  pipette.  The 
tip  was  then  brought  close  to  an  aggregation,  and  the  bulb  was 
pressed  gently  to  force  the  remaining  volatile  pheromone  from  the 
tip.  As  a control,  a pipette  with  evaporating  chloroform  was  also 
used.  Chloroform  vapors  disturbed  resting  aggregations,  causing 
individuals  to  move  away  from  the  pipette;  but  chloroform  alone 
did  not  elicit  a full  alarm  response  with  characteristic  group 
movement  to  a new  leaf.  Alarm  responses  were  studied  under  well- 
lighted  conditions  with  and  without  the  female  present. 

RECORDING  OF  DATA 

Observations  were  written,  tape  recorded,  or  photographed. 
Direction  of  movement  of  aggregations  on  the  host  plant  was 
recorded  as  up,  down,  or  “up  and  down”  (part  of  the  group  moved 
up,  part  moved  down).  Movement  up  or  down  was  considered 
“directed”;  movement  both  up  and  down,  “undirected”.  The  new 


218 


Psyche 


[Vol.  88 


position  of  the  group  was  recorded  as  1,  2,  3,  or  4 or  more  leaves 
above  starting  position.  The  designation  “4  or  more”  included  leaf 
no.  4 and  several  small  leaves  at  the  growing  tip  of  the  plant. 

RESULTS 

Egg  masses  of  G.  solani  are  deposited  on  the  underside  of  leaves, 
and  first  instar  larvae  feed  from  the  leaf  surface  between  the  eggs 
and  then  from  the  areas  adjacent  to  the  eggs.  The  larvae  are  usually 
in  a compact  circular  formation  while  feeding,  and  the  aggregation 
moves  away  from  the  oviposition  site  as  leaf  tissue  is  destroyed. 
Feeding  sites  become  yellow  or  brown  in  color  and  also  brittle.  On  a 
large  eggplant  leaf,  nymphs  may  pass  through  several  instars  before 
consuming  most  of  the  leafs  soft  tissue.  On  the  relatively  smaller 
horse  nettle  leaf,  an  aggregation  consumes  the  edible  portion  of  a 
leaf  more  quickly  and  then  moves  to  another  leaf. 

MOVEMENTS  TO  NEW  FEEDING  SITES 

Movements  to  new  sites  on  the  same  leaf  seemed  to  proceed 
gradually  and  with  little  intervention  from  the  female.  As  individ- 
uals in  the  aggregation  withdrew  their  stylets,  they  moved  away 
from  where  they  were  feeding,  bumping  into  adjacent  nymphs. 
These  bumped  nymphs  in  turn  withdrew  their  stylets  and  moved  or 
milled  about,  bumping  into  other  nymphs  until  the  entire  aggrega- 
tion was  activated.  Movement  from  the  feeding  site  to  another 
feeding  site  on  the  same  leaf  then  ensued.  Movement  to  a new  leaf 
usually  occurred  after  75%  or  more  of  the  leaf  was  damaged  and 
often  lasted  for  about  an  hour.  The  parent  female,  also  activated  by 
the  milling  nymphs,  usually  moved  slowly  down  the  petiole  while 
keeping  close  physical  contact  with  the  nymphs  immediately  behind 
her.  If  there  was  any  break  in  contact,  the  nearest  nymphs  moved 
forward  and  touched  the  tips  of  the  female’s  wings  with  their 
antennae,  or  the  female  turned  around  and  touched  her  antennae  to 
those  of  the  nearest  advancing  nymphs.  During  one  group  move- 
ment, the  female  waited  first  at  the  axil  of  the  new  leaf  and  then  on 
the  underside  of  the  petiole,  as  nymphs  filed  by  her.  This  behavior 
was  identical  to  that  observed  during  alarm  responses.  On  two 
occasions,  the  female  seemed  to  initiate  movement  of  the  nymphs  by 
forcing  her  way  into  the  cluster  of  feeding  nymphs;  but  this  occurred 
only  when  the  nymphs  were  in  the  earlier  instars  and  were  moving 


1981] 


Kearns  & Yamamoto  — Gargaphia 


219 


Table  2.  Direction  of  movement  of  G.  solani  aggregations  on  a host  plant 


Female  Present  Female  Absent 

% of  groups  observed  % of  groups  observed 


Upward 

Directed 

Upward 

Directed 

Feeding  Movements 

movement 

movement 

movement 

movement 

In  light 

82* 

96*  (1) 

78 

100*  (2) 

In  darkness 

89* 

100*  (2) 

100* 

100*  (3) 

Light  source  below 

aggregation 

71 

78  (2) 

75 

67  (3) 

Alarm  Response 

In  light 

90* 

100*  (4) 

62 

67  (5) 

(1)  23  observations 

(2)  9 observations 

(3)  6 observations 

(4)  19  observations 

(5)  12  observations 

* Significant  at  the  95%  confidence  level 


from  one  surface  of  the  leaf  to  the  opposite  surface.  At  no  time  was 
“herding”  by  the  female  observed  as  described  by  Fink  (1915). 

Movements  to  a new  leaf  were  difficult  to  predict  and  lengthy  to 
monitor.  Because  of  time  considerations,  it  was  not  feasible  to  make 
a statistical  study  of  the  female’s  total  behavior  pattern  during  these 
group  movements.  When  a female  was  present,  she  led  the  group  to 
a new  leaf.  In  the  absence  of  a female,  the  nymphs  moved  on  their 
own.  Females  sometimes  wandered  about  on  adjacent  leaves  but 
usually  returned  to  their  aggregations. 

ALARM  RESPONSES 

When  an  aggregation  of  the  third  through  fifth  instar  nymphs  of 
G.  solani  was  alarmed  with  a squashed  fifth  instar  nymph,  the  group 
responded  quickly,  usually  within  10  seconds.  The  duration  of  the 
response  was  from  4 to  20  minutes.  If  the  nymphs  were  on  the  top 
surface  of  the  leaf  they  moved  to  the  underside,  and  conversely.  In 
either  case,  at  least  some  of  the  nymphs  moved  quickly  to  the  midrib 
and  from  there  to  the  petiole  of  the  leaf.  At  the  exit  axil,  the  nymphs 
moved  up  or  down  the  stem;  but  they  were  more  likely  to  move  up 
the  stem  (Table  2,  Female  Present).  During  this  activity,  the  female 


220 


Psyche 


[Vol.  88 


Figure  1.  Adult  female  of  G.  solani  positioned  at  the  exit  axil  during  an  alarm 
response. 


moved  quickly  to  the  axil  of  the  leaf  and  oriented  herself  a little  to 
one  side  of  the  nymphs’  path  (Fig.  1).  When  she  was  in  this  position, 
the  nymphs  moved  up  the  stem  rather  than  down. 

As  more  of  the  nymphs  left  the  leaf,  the  female  sometimes  moved 
about  the  axil  and  positioned  herself  along  the  side  of  the  stem  as 
the  nymphs  moved  past  her.  When  she  was  in  these  positions,  she 
seemed  to  have  little  physical  contact  with  the  nymphs  which  filed 
past  her  unless  they  happened  to  bump  into  her  in  passing.  Part  way 
through  an  alarm  response,  the  female  moved  quickly  up  the  stem, 
usually  to  the  axil  of  the  first  leaf  above  the  previously  occupied 
leaf.  Many  of  the  nymphs  had  already  reached  this  axil  and  had 
moved  down  the  petiole  onto  the  new  leaf.  A few  nymphs  often 
proceeded  above  this  axil  and  continued  up  the  plant  before 
returning  to  the  group.  The  female  oriented  herself  at  the  axil  of  the 
new  leaf  (Fig.  2)  and  waited  there  as  more  nymphs  arrived.  After 
most  of  the  nymphs  had  passed  along  the  petiole,  the  female  joined 
the  aggregation  on  the  new  leaf. 


1981] 


Kearns  & Yamamoto  — Gargaphia 


221 


There  were  always  a few  nymphs  that  remained  behind  or  that 
failed  to  keep  up  with  the  bulk  of  the  aggregation.  These  slower 
individuals  wandered  out  onto  other  leaves  but  did  not  settle  down 
until  they  found  the  group.  Apparently  the  nymphs  maintain 
locomotor  activity  unless  they  have  sufficient  physical  or  chemical 
contact  with  other  nymphs.  There  were  variations  in  the  female’s 
behavior  (Table  1),  but  it  was  not  obvious  what  environmental 
conditions  might  cause  the  female  to  include  or  change  a particular 
element  of  her  behavior.  When  an  aggregation  of  first  and/or 
second  instars  was  alarmed,  the  group  was  likely  to  relocate  on  the 
same  surface  of  the  leaf  rather  than  to  move  to  the  opposite  surface 
or  off  the  leaf. 

Certain  elements  of  the  female’s  behavior  were  clearly  recogniz- 
able and  repeated  more  than  once.  These  elements  are  recorded  in 
Table  1 with  the  frequency  of  their  occurrence  in  18  different 
experimental  responses  to  alarm  pheromone.  In  89%  of  the  alarm 
responses,  the  female  exhibited  at  least  one  of  the  first  4 elements 


Figure  2.  Adult  female  of  G.  solani  positioned  at  the  extrance  axil  during  an  alarm 
response. 


222 


Psyche 


[Vol.  88 


listed.  The  fourth  element,  female  positioned  at  the  entrance  axil 
(Figure  2),  was  repeated  most  frequently,  although  the  element  was 
not  itself  significant.  There  was  more  variation  in  the  female’s  posi- 
tion on  the  stem  at  the  beginning  of  the  alarm  response  than  at  the 
end. 


DIRECTION  OF  MOVEMENT 

Aggregations  with  and  without  the  female  present  were  studied  in 
order  to  determine  the  significance  of  the  female’s  role  in  feeding 
movements  or  alarm  responses.  It  was  hypothesized  that  the 
female’s  presence  would  keep  the  aggregation  together  and  inhibit 
random  movements  on  the  plant.  The  relocation  of  the  group  was 
studied  in  terms  of  the  direction  of  movement  on  the  plant  (up, 
down,  or  in  both  directions)  and  the  choice  of  a new  leaf  on  which  to 
feed.  Movement  was  considered  directed  if  a group  moved  in  one 
direction  or  the  other,  but  not  in  both.  A table  for  a binomial 
distribution  was  used  to  evaluate  significance  at  the  95%  confidence 
level  (Table  2,  partial  sums,  Eisenhart  1952;  Table  A-22,  confidence 
intervals,  Natrella  1963).  The  results  for  upward  and  directed 
movements  of  aggregations,  with  and  without  the  female  present, 
are  recorded  in  Table  2 as  percentages  of  groups  observed.  Those 
results  significant  at  the  95%  confidence  level  are  marked  with  an 
asterisk. 

Feeding  Movements 

In  light  or  in  darkness,  feeding  movements  with  or  without  the 
female  present  were  directed  rather  than  random,  and  the  group 
usually  moved  upward.  When  the  source  of  light  was  180°  away 
from  the  usual  direction,  neither  moving  upward  nor  directed 
movement  was  significant;  but  the  aggregations  did  not  reverse  their 
direction  of  movement  and  move  toward  the  light.  It  is  possible  that 
the  abnormal  position  of  the  light  source  acted  as  a conflicting 
stimulus  which  confused  some  of  the  aggregations. 

A small  field  sample  of  horse  nettle  plants  (13)  showing  damage 
from  G.  solani  was  examined  for  evidence  of  group  movements. 
Eighty-five  percent  of  the  groups  had  moved  upward  on  the  plants 
from  leaves  containing  the  remains  of  egg  masses.  Moving  up  was 
significant  at  the  95%  confidence  level  and  closely  matched  the 
results  obtained  in  the  laboratory. 


1981] 


Kearns  & Yamamoto  — Gargaphia 


223 


Alarm  Response 

The  female’s  presence  or  absence  made  a significant  difference 
when  the  aggregation  was  alarmed.  When  the  female  was  present, 
90%  of  the  groups  moved  upward,  and  10%  moved  downward;  but 
none  split  and  moved  in  both  directions.  When  the  female  was 
absent,  67%  of  the  groups  moved  either  upward  or  downward;  one- 
third  of  the  groups  split. 

POSITION  OF  AGGREGATIONS  FOLLOWING  FEEDING 
MOVEMENTS  OR  ALARM  RESPONSES 

The  results  of  experiments  for  both  feeding  and  alarm  movements 
were  combined,  and  a comparison  was  made  between  female 
present  and  female  absent. 

Choice  of  Leaf 

With  a choice  of  4 leaf  positions  above  the  one  occupied  by  the 
aggregation,  the  probability  of  an  aggregation’s  reassembling  on 
any  one  of  the  leaves  was  0.25.  Using  the  binomial  distribution 
(Eisenhart  1952),  we  compared  the  choice  of  leaf  no.  1 with  the 
choice  of  any  other  leaf  (Table  3).  Whether  females  were  present  or 
absent,  the  aggregations  were  more  likely  to  move  to  leaf  no.  1 than 
to  any  of  the  other  leaves.  If  the  aggregations  split  between  leaves, 
the  split  usually  included  leaf  no.  1.  This  behavioral  pattern  of  the 
nymphs  increased  the  likelihood  that  the  group  would  remain 
together  following  movements  on  the  host  plant.  When  the  female 
was  present,  the  group  was  more  likely  to  move  as  a unit  to  leave 
no.  1. 

Choice  of  Single  or  Multiple  Leaves 

Movement  to  a single  leaf  was  compared  to  movement  to  multiple 
leaves  (Table  4).  The  probability  associated  with  this  choice  was  0.5. 
When  females  were  present,  aggregations  usually  moved  as  a unit  to 
a single  leaf  on  the  host  plant  but  when  females  were  absent, 
aggregations  split  up  as  often  as  they  chose  a single  leaf. 

WING  FANNING  BY  THE  FEMALE 

Fink  (1915)  reported  that  on  one  occasion  he  saw  an  adult  female 
G.  solani  chase  a ladybeetle  (Hippodamia  convergens  Guer.)  away 


224 


Psyche 


[Vol.  88 


Table  3.  Choice  of  leaf  position  by  G.  solani  aggregations  following  movements 
(1)  on  a host  plant 


% of  groups  observed 
Female  Present  Female  Absent 

Leaf  no.  1 

56 

36 

Leaf  no.  1 in  combination  with  one  or 

17 

36 

more  leaves 

Total  positions  including  leaf  no.  1 

72*  (2) 

71*  (3) 

(1)  Feeding  movements  and  alarm  responses  combined 

(2)  36  observations 

(3)  14  observations 

* Significant  at  the  95%  confidence  level 


from  an  aggregation  of  feeding  nymphs:  the  female  “with  out- 
stretched, slightly  raised  wings  suddenly  darted  toward  the  intruder, 
driving  it  from  the  leaf.”  In  the  laboratory,  adult  females  of  G. 
solani  responded  similarly  (Fig.  3)  to  ladybeetles,  anthocorids,  ants, 
the  tip  of  a brush,  and  a tomato  pinworm  caterpillar  which  was 
spinning  a cocoon.  Beamer  (1930)  and  Wood  (1976a,  1976b,  1977, 
and  1978)  reported  wing  fanning  in  a total  of  4 species  of  mem- 
bracids.  In  each  of  these  species,  wing  fanning  was  used  by  the  adult 
female  as  a response  to  a predator  (Beamer  1930;  Wood  1976a, 
1976b,  1977)  or  a threatening  stimulus,  such  as  a pencil  used  to  prod 
the  female  (Wood  1978).  Sheeley  and  Yonke  (1977)  observed  wing 
fanning  by  the  tingid  Corythucha  bulbosa  when  a jumping  spider 


Table  4.  Choice  of  single  or  multiple  leaves  by  G.  solani  aggregations  following 
movements  (1)  on  a host  plant 


% of  groups  observed 
Female  Present  Female  Absent 

Choice 

Single  leaf 

81*  (2) 

50  (3) 

Multiple  leaves 

19  (2) 

50  (3) 

(1)  Feeding  movements  and  alarm  responses  combined 

(2)  36  observations 

(3)  14  observations 

* Significant  at  the  95%  confidence  level 


1981] 


Kearns  & Yamamoto  — Gargaphia 


225 


approached,  and  they  reported  that  the  spider’s  response  to  touch- 
ing the  tingid  suggested  the  presence  of  a defensive  chemical. 

Wing  fanning  in  G.  solani  occurred  not  only  in  response  to  a 
predator,  but  also  under  other  circumstances.  It  was  often  associ- 
ated with  alarm  responses  and  was  directed  toward  the  nymphs  as 
well  as  toward  a possible  predator.  For  27  brooding  females,  143 
occurrences  of  fanning  were  recorded  in  2 categories:  deterring  a 
predator  (26%)  and  controlling  the  nymphs  in  one  of  several  ways 
(74%). 

Deterring  Predators 

The  brooding  female  responded  to  predators  quickly  after  she 
detected  their  presence.  The  relatively  large  coccinellids  ( Hippo - 
damia  convergens,  Olla  abdominalis)  were  detected  more  readily 
than  smaller  predators  such  as  Pharaoh  ants  ( Monomorium  phara- 
onis)  or  the  anthocorid,  Orius  insidiosus.  Attacks  by  ants  and 
anthocorids  were  observed  with  a dissecting  microscope.  Females 


Figure  3.  Adult  female  of  G.  solani  fanning  her  wings. 


226 


Psyche 


[Vol.  88 


and  nymphs  failed  to  respond  when  a single  ant  removed  an  egg  or 
ate  a newly  hatched  first  instar  nymph.  Ants  carried  tiny  nymphs 
away  from  the  brood,  and  the  release  of  alarm  pheromone  was 
apparently  not  detected.  When  two  or  more  ants  moved  in  front  of  a 
brooding  female,  she  responded  with  wing  fanning  and  moved  her 
body  over  the  egg  mass.  Orius  insidiosus  nymphs,  which  were  about 
the  size  of  second  instar  G.  solani  nymphs,  attacked  their  victims  by 
penetrating  intersegmental  membranes.  O.  insidiosus  was  not  al- 
ways detected  by  the  brooding  female  or  nymphs,  perhaps  because 
the  site  of  penetration  was  often  in  the  coxal  area  rather  than  on  the 
abdomen.  When  an  attack  occurred  in  front  of  a brooding  female, 
she  responded  by  fanning  her  wings  and  prodding  the  anthocorid 
with  her  head.  Anthocorids  responded  by  remaining  motionless  for 
periods  of  up  to  55  minutes  in  length. 

An  attendant  female  responded  to  a coccinellid  by  rushing  at  it, 
fanning  her  wings,  and,  occasionally,  by  prodding  it  with  her  head. 
In  5 experiments  with  the  adult  coccinellid  Hippodamia  convergens, 
first  and  second  instar  nymphs  were  killed  each  time.  In  3 of  those 
encounters,  the  adult  female  lace  bug  was  successful  in  driving  away 
the  coccinellid,  preventing  further  loss  of  nymphs.  In  2 encounters 
with  starved  coccinellids,  the  female  lace  bug  was  not  able  to  drive 
the  attacker  away.  The  remaining  nymphs  survived  because  they 
fled  apparently  in  response  to  an  alarm  pheromone  released  by 
crushed  nymphs.  In  3 encounters  with  the  coccinellid  Olla  abdomi- 
nalis,  the  female  lace  bug  chased  the  approaching  beetle  successfully 
(Fig.  4);  however,  the  beetle  did  not  attack  any  nymphs  or  show 
much  interest  in  them. 

Controlling  Nymphs 

Females  used  wing  fanning  in  their  interactions  with  the  nymphs. 
On  at  least  6 occasions,  the  attendant  female  went  ahead  of  the 
moving  aggregation  and  waited  on  the  new  leaf  for  the  nymphs  to 
arrive.  While  waiting  for  the  nymphs,  the  females  fanned  their  wings 
repeatedly. 

There  were  a number  of  instances  in  which  wing  fanning  was  used 
to  quiet  a restless  aggregation  or  one  which  had  recently  dispersed 
to  a new  leaf.  The  adult  female  circled  the  group  with  rapid,  jerky 
movements  and  stopped  occasionally  to  fan  her  wings.  For  2 
different  females  and  aggregations,  the  female  backed  up  to  the 


1981] 


Kearns  & Yamamoto  — Gargaphia 


227 


Figure  4.  Adult  female  of  G.  solani  responding  to  coccinellid,  Olla  abdominalis. 


nymphs  and  pointed  the  tip  of  her  abdomen  toward  them  as  she 
fanned  her  wings.  This  behavior  suggests  that  the  fanning  may  be 
used  to  propel  a pheromone  toward  the  group.  One  female  was 
observed  to  use  wing  fanning  to  prevent  the  movement  of  an 
aggregation.  The  female  was  oriented  at  the  base  of  the  leaf,  headed 
toward  the  group  of  nymphs.  When  2 nymphs  left  the  group,  moved 
down  the  mid-vein,  and  approached  her,  she  fanned  her  wings.  The 
nymphs’  response  was  a retreat. 

AGGREGATION 

There  is  some  indirect  evidence  for  an  aggregation  pheromone  or 
for  the  nymphs’  need  for  physical  contact  with  each  other.  Upon 
hatching,  nymphs  feed  near  the  egg  mass  for  a short  period  and  then 
move  away  as  an  aggregation.  Older  nymphs  wander  away  from 
their  own  aggregations  and  join  others,  stray  fifth  instars  being 
particularly  conspicuous  when  they  join  groups  of  first  and  second 


228 


Psyche 


[Vol.  88 


instars.  Following  one  experiment  on  the  alarm  response,  5 fifth 
instar  nymphs  from  a second  aggregation  were  released,  one  at  a 
time,  slightly  above  the  leaf  just  vacated  by  the  first  aggregation. 
Four  of  the  fifth  instars  moved  directly  up  the  stem  and  onto  the 
newly  occupied  leaf,  and  the  remaining  nymph  wandered  about, 
first  on  higher  leaves  and  then  on  the  vacated  leaf  before  moving  to 
the  occupied  leaf.  Other  experiments  showed  that  nymphs  would 
reaggregate  after  they  were  separated  by  the  experimenter.  Need  for 
physical  contact  might  explain  this  adequately,  but  the  presence  of 
an  aggregation  pheromone  should  not  be  ruled  out.  Reaggregation 
is  essential  if  the  alarm  response  is  to  occur  repeatedly. 

female’s  behavioral  maturation 

Preliminary  experiments  (Kearns  1980)  indicate  that  females 
undergo  behavioral  maturation  from  the  time  of  oviposition  through 
egg  hatch  and  early  development  of  the  nymphs.  Females  at 
different  stages  of  development  were  substituted  for  females  which 
were  attending  aggregations  of  nymphs.  Only  those  substitutes 
which  had  attended  aggregations  of  their  own  behaved  normally 
during  an  alarm  response.  When  females  which  were  still  oviposit- 
ing were  used  as  substitute  mothers,  they  either  avoided  the  alarmed 
nymphs  or  failed  to  interact  with  them. 

CHLOROFORM  EXTRACT 

A chloroform  extract  of  G.  solani  nymphs  proved  to  be  as 
effective  in  eliciting  an  alarm  response  as  a fifth  instar  nymph 
squeezed  with  forceps  or  squashed  on  filter  paper.  Crushed  adults 
also  released  the  alarm  pheromone,  but  the  nymphal  response  was 
slower  by  a minute  or  less,  to  a crushed  adult  than  to  a crushed 
nymph.  Preliminary  attempts  were  made  to  test  for  alarm  pher- 
omones in  3 other  tingids  available  locally:  Corythucha  ciliata,  the 
sycamore  lace  bug;  Corythucha  cydoniae,  the  hawthorne  lace  bug; 
and  Corythucha  marmorata,  the  chrysanthemum  lace  bug.  Nymphs 
of  each  species  showed  an  alarm  response  to  a crushed  nymph  of 
the  same  species.  There  were  also  cross  responses  between  G.  solani 
and  each  of  the  three  species  of  Corythucha.  Since  not  all  three 
species  of  Corythucha  overlap  in  time,  it  will  be  necessary  to  rear  the 
insects  in  the  laboratory  or  to  make  extracts  of  each  for  testing  cross 
responses. 


1981] 


Kearns  & Yamamoto  — Gargaphia 


229 


DISCUSSION 

Gargaphia  solani  and  some  of  the  membracids  (Wood  1974, 
1976b)  are  unusual  in  having  maternal  care  extend  from  the  time  of 
oviposition  through  the  maturation  period  of  the  nymphs.  This  long 
brooding  period  appears  to  be  an  adaptation  to  environments  in 
which  predation  is  an  important  factor.  The  host  plants  of  G.  solani 
grow  close  to  the  ground,  and  ants  appear  to  be  the  most  numerous 
predators.  Maternal  care  in  this  tingid  may  have  evolved  as  a 
response  to  ants  or  to  low-flying  predators  or  to  both.  Sheeley  and 
Yonke  (1977)  were  unable  to  find  predators  for  some  of  the  7 species 
of  tingids  studied,  perhaps  because  the  host  plants  of  6 species  are 
trees  rather  than  small  annuals.  Gargaphia  tiliae,  having  maternal 
care,  might  be  expected  to  live  close  to  the  ground,  but  it  is  a tree- 
dwelling species.  Sheeley  and  Yonke  found  no  natural  enemies  of 
this  insect,  but  the  predators  could  have  included  tiny  anthocorid 
nymphs  which  escaped  detection. 

It  seems  worthwhile  to  compare  G.  solani  with  some  of  the 
membracid  species  since  there  are  striking  similarities,  including 
wing  fanning  by  the  attendant  female  and  the  release  of  an  alarm 
pheromone  when  the  body  wall  is  ruptured.  If  G.  solani  and  the 
membracids  represent  examples  of  parallel  evolution,  they  may  be 
responding  to  similar  environmental  stresses. 

Literature  Cited 


Beamer,  R.  H. 

1930.  Maternal  instinct  in  a Membracid  ( Platycotis  vittata)  (Homopt.)  Ento- 
mol.  News  41(10):  330-331. 

Eisenhart,  C. 

1952.  Tables  of  the  Binomial  Probability  Distribution.  National  Bureau  of 
Standards  Applied  Mathematics  Series  6.  U.  S.  Govt.  Printing  Office, 
Washington. 

Fink,  D.  E. 

1915.  The  eggplant  lace-bug.  Bull.,  U.  S.  Dept.  Agricult.  239:  1-7. 

Hinton,  H.  E. 

1976.  Maternal  care  in  the  Membracidae.  Proc.  Roy.  Entomol.  Soc.  London 
(C).  41:  3-4. 

1977.  Subsocial  behavior  and  biology  of  some  Mexican  membracid  bugs. 
Ecological  Entomology  2:  61-79. 

Kearns,  R.  S. 

1980.  Maternal  behavior  in  the  eggplant  lace  bug  Gargaphia  solani  Heide- 
mann  (Tingidae:  Heteroptera).  M.S.  thesis,  North  Carolina  State  Uni- 
versity, Raleigh. 


230 


Psyche 


[Vol.  88 


Melber,  A.,  and  G.  H.  Schmidt 

1977.  Sozialphanomene  bei  Heteropteren.  Sonderdruck  aus  Zoologica.  127: 
19-53. 


Natrella,  M.  G. 

1963.  Experimental  Statistics.  National  Bureau  of  Standards  Handbook  91.  U. 
S.  Govt.  Printing  Office,  Washington. 

Nault,  L.  R.,  T.  K.  Wood,  and  A.  M.  Goff 

1974.  Treehopper  (Membracidae)  alarm  pheromones.  Nature  249:  387-388. 
Sheeley,  R.  D.  and  T.  R.  Yonke 

1977.  Biological  notes  on  seven  species  of  Missouri  tingids  (Hemiptera: 
Tingidae).  J.  Kansas  Entomol.  Soc.  50:  342-356. 

Torre-Bueno,  J.  R. 

1935.  Notes  on  Gargaphia  tiliae.  Bull.  Brooklyn  Entomol.  Soc.  30:  78. 
1942.  Maternal  solicitude  in  Gargaphia  iridescens  Champion.  Bull.  Brooklyn 
Entomol.  Soc.  37:  131. 

Weiss,  H.  B. 

1919.  Notes  on  Gargaphia  tiliae  Walsh,  the  linden  lace-bug.  Proc.  Biol.  Soc. 
Wash.  32:  165-168. 


Wood,  T.  K. 

1974.  Aggregating  behavior  of  Umbonia  crassicornis  (Homoptera:  Membra- 
cidae). Can.  Ent.  106:  169-173. 

1976a.  Alarm  behavior  of  brooding  female  Umbonia  crassicornis  (Membra- 
cidae: Homoptera).  Ann.  Entomol.  Soc.  Amer.  69:  340-344. 

1976b.  Biology  and  presocial  behavior  of  Platycotis  vittata  (Homoptera:  Mem- 
bracidae). Ann.  Entomol.  Soc.  Amer.  69:  807-811. 

1977.  Role  of  parent  females  and  attendant  ants  in  the  maturation  of  the 
treehopper,  Entylia  bactriana  (Homoptera:  Membracidae).  Sociobiol- 
ogy 2:  257-272. 

1978.  Parental  care  in  Guayaquila  compressa  Walker  (Homoptera:  Membra- 
cidae). Psyche  85:  135-145. 


POLYMORPHISM  AND  DIVISION  OF  LABOR  IN  THE 
DACETINE  ANT  ORECTOGNATHUS  VERSICOLOR 
(HYMENOPTERA:  FORMICIDAE)* 

By  Norman  F.  Carlin 

Department  of  Organismic  and  Evolutionary  Biology, 
Harvard  University,  Cambridge,  Mass.  02138 

Introduction 

The  ants  of  the  myrmicine  tribe  Dacetini  exhibit  a primary 
evolutionary  trend  from  primitive  epigaeic  and  subarboreal  foragers 
to  advanced  cryptobiotic  forms;  in  association  with  this  trend  are  a 
number  of  secondary  tendencies,  including  reduction  in  body  size 
and  mandible  length,  increasing  specialization  on  collembolan  prey, 
and  loss  of  worker  caste  differentiation  (Brown  and  Wilson  1959). 
The  subarboreal  and  impressively  long-mandibulate  subtribe  Orec- 
tognathiti,  comprising  the  genera  Orectognathus  and  Arnoldidris, 
occupies  an  intermediate  position  between  the  primitive  polymor- 
phic genus  Daceton  and  the  largely  monomorphic  higher  subtribes 
Epopostrumiti  and  Strumigeniti.  All  but  one  of  the  twenty-nine 
known  species  of  Orectognathus  are  monomorphic,  the  exception 
being  O.  versicolor , which  possesses  a distinctive  major  caste 
(Taylor  1977,  1979).  Caste  differentiation  in  this  species  is  con- 
sidered to  have  evolved  secondarily,  from  the  monomorphic  generic 
stock  (Brown  and  Wilson  1959). 

The  extreme  polymorphism  of  Daceton  armigerum,  the  only 
lower  dacetine  whose  behavior  has  been  studied,  is  put  to  work  in  an 
equally  extreme  division  of  labor  (Wilson  1962).  The  minor  workers 
are  strictly  limited  to  brood  care  tasks  (in  which  they  are  aided  by 
callows  of  larger  castes),  and  to  regurgitation  with  other  adults. 
Small  medias  forage  widely  and  actively,  but  larger  medias  and 
majors  tend  to  rest  in  “way-stations”  some  distance  from  the  nest. 
These  large  workers  take  prey  away  from  returning  smaller  foragers, 
bringing  it  into  the  nest  themselves,  so  that  little  prey  is  carried  back 
by  those  that  hunt  for  it.  The  species  takes  a broad  variety  of  prey 
items;  it  has  been  suggested  that  the  dietary  specialization  on 
collembolans  seen  in  higher  dacetines  might  account  for  their 


♦Manuscript  received  by  the  editor  December  1,  1981. 


231 


232 


Psyche 


[Vol.  88 


surrendering  the  polymorphism  and  polyethism  of  Daceton  (Wilson 
1971). 

Orectognathus  versicolor , as  the  sole  polymorphic  intermediate 
dacetine,  is  of  special  interest  for  polyethism  analysis.  The  species  is 
also  an  easy  one  to  study,  its  slow-moving  habits  and  small  colony 
size  making  possible  the  recording  of  nearly  every  behavioral  act 
performed  by  each  individual  worker.  The  minor  workers  possess 
the  same  long,  slender  mandibles,  with  pointed  apical  teeth,  that 
their  congeners  bear.  Majors,  however,  have  massive,  relatively 
short  mandibles,  with  apical  teeth  thick,  blunt  and  recessed;  their 
large  occipital  lobes  contain  disproportionately  developed  mandible 
adductor  muscles  (figs.  1 and  2).  In  mandible  allometry,  at  least,  this 
species  may  be  the  most  exaggeratedly  polymorphic  of  all  dacetines. 
The  division  of  labor  by  which  such  morphologically  divergent 
forms  are  utilized,  particularly  since  the  major  caste  is  a secondary 
development,  may  shed  light  on  the  advantages  of  specialized  castes 
in  the  context  of  dacetine  evolution.  To  what  use  are  the  singular 
majors  put?  Does  the  polyethism  of  O.  versicolor  in  any  way 
resemble  that  of  Daceton , or  is  it  entirely  independent?  Has  the 
return  to  polymorphism  been  accompanied  by  a return  to  the 
polyphagy  of  Daceton , or  is  O.  versicolor  a collembolan  specialist, 
as  the  rest  of  its  genus  is  thought  to  be  (Brown  1953)?  An 
opportunity  to  address  these  questions  in  the  laboratory  arose  when 
Bert  Holldobler  brought  a live  queenright  colony  of  these  ants  from 
North  Queensland,  Australia;  the  results  of  observation  of  this 
colony  are  reported  below. 

Materials  and  Methods 

The  O.  versicolor  colony  was  settled  in  a glass  test  tube  (2  cm  in 
diameter),  with  water  trapped  at  its  end  behind  a tight  cotton  plug. 
The  tube  was  placed  in  a plaster-floored  clear  plastic  container  (18 
cm  by  12  cm  by  6 cm),  and  a dissecting  microscope  was  set  over  it  on 
a moveable  mount  to  permit  viewing  of  ants  both  inside  the  nest 
tube  and  out  on  the  container  floor.  A total  of  45  hours  of 
observation  were  made  over  a period  of  five  weeks,  during  which 
7,891  separate  behavioral  acts  were  recorded.  Estimation  of  the 
completeness  of  caste  behavior  repertories  was  made  by  fitting  the 
data  to  a lognormal  Poisson  distribution,  following  the  method  of 
Fagen  and  Goldman  (1977).  The  ants  were  offered  various  food 


1981] 


Carlin  — Polymorphism  in  Orectognathus 


233 


Figure  1:  An  Orectognathus  versicolor  colony.  The  queen  is  at  the  left;  to  her  right  are  two  major  workers — note  their 
mandibles  and  head  size  and  shape.  To  their  right  are  another  major  (top),  minor  (middle)  and  media  (bottom)  workers. 


234 


Psyche 


[Vol.  88 


items;  to  examine  their  defensive  behavior,  small  Solenopsis  invicta 
workers  were  introduced  into  their  container. 

The  two  morphological  castes  were  easily  distinguished  on  the 
basis  of  mandible  thickness.  In  order  to  record  division  of  labor 
among  individuals  of  different  sizes,  yet  similar  proportions — so 
critical  in  weakly  allometric  species  such  as  Daceton — the  minor 
workers  were  arbitrarily  divided  into  small  and  medium  size  classes, 
also  distinguishable  by  eye.  For  convenience,  these  subcastes  will  be 
referred  to  as  “minors”  and  “medias”,  as  in  Wilson  1978.  By-eye 
assignment  of  caste  to  preserved  specimens,  subsequently  measured, 
produced  the  following  definitions  of  size  classes  and  castes:  minors, 
head  width  less  than  1.12  mm;  medias,  head  width  between  1.13  and 
1.64  mm;  majors,  head  width  greater  than  1.65  mm.  After  some 
initial  die-off,  the  colony  contained  fifty-two  adults  for  the  duration 
of  the  study:  one  queen,  thirty  minors,  fifteen  medias  and  six 
majors. 


Results 

O.  versicolor  is  in  fact  polyphagous.  Live  flightless  Drosophila 
were  readily  accepted,  and  young  were  successfully  raised  on  this 
diet.  The  ants  also  accepted  Drosophila  larvae,  and,  not  surprisingly, 
collembolans.  (Alternative  foods  were  not  offered  simultaneously  to 
test  preferences;  however,  most  collembolan  specialist  species  would 
not  touch  other  prey  even  if  starving.)  The  same  colony  had  been  fed 
mealworm  and  cockroach  fragments,  various  diptera  and  honey- 
water  in  Australia  (B.  Holldobler,  pers.  comm.). 

The  ethogram  or  behavioral  catalogue  of  workers  and  queen  is 
presented  in  table  1,  which  gives  both  numbers  of  individual  acts 
performed  and  the  relative  frequencies  of  acts  in  the  total  repertory 
of  each  caste.  The  colony  repertory  consisted  of  twenty-seven 
categories  of  behavior.  (Worker  regurgitation  with  the  queen  was 
added  as  a twenty-eighth  because  it  was  seen  twice  during  prelimi- 
nary observations,  though  never  during  the  study.)  The  observed 
minor  and  media  repertories  both  contained  twenty-seven  behavior 
categories;  the  observed  major  repertory  contained  twenty-four. 
Using  the  Fagen-Goldman  statistical  method,  the  estimated  total 
repertory  size  for  minors — the  observed  repertory  plus  an  estimate 
of  the  number  of  categories  not  observed — was  calculated  to  be 
twenty-nine,  with  a 95%  confidence  interval  of  (27,32)  acts.  The 


1981] 


Carlin  — Polymorphism  in  Orectognathus 


235 


236 


Psyche 


[Vol.  88 


Table  1 : Ethogram  of  Orectognathus  versicolor.  The  values  given  are  numbers  of 
individual  acts  performed  by  members  of  each  caste.  In  parentheses  are  given  relative 
frequencies  of  performance  of  each  act  in  the  total  repertory  of  the  caste. 


Minor 

Media 

Major 

Queen 

Self-groom 

1370  ( .3481) 

845  ( .2996) 

462  ( .4306) 

45  ( .7258) 

Allogroom  minor 

437  ( .1110) 

93  ( .0330) 

18  ( .0168) 

1 ( .0161) 

Allogroom  media 

144  ( .0366) 

224  ( .0794) 

16  ( .0149) 

0 

Allogroom  major 

100  ( .0254) 

52  ( .0184) 

15  ( .0140) 

0 

Allogroom  queen 

27  ( .0069) 

20  ( .0071) 

8 ( .0075) 

/ 

Regurgitation 

with  minor 

48  ( .0122) 

15  ( .0053) 

3 ( .0028) 

2 ( .0323) 

with  media 

10  ( .0025) 

21  ( .0074) 

7 ( .0065) 

0 

with  major 

13  ( .0033) 

2 ( .0007) 

2 ( .0019) 

0 

with  queen 

0 

0 

0 

/ 

Carry  or  manipu- 

late  egg 

4 ( .0010) 

2 ( .0007) 

1 ( .0009) 

0 

Lick  egg 

18  ( .0046) 

5 ( .0018) 

1 ( .0009) 

1 ( .0161) 

Carry  or  manipu- 
late larva 

53  ( .0135) 

37  ( .0131) 

2 ( .0019) 

0 

Lick  larva 

602  ( .1529) 

520  ( .1844) 

155  ( .1445) 

9 ( .1425) 

Regurgitate  with 
larva 

4 ( .0010) 

1 1 ( .0039) 

2 ( .0019) 

0 

Feed  larva  solids 

19  ( .0048) 

20  ( .0071) 

0 

0 

Carry  or  manipu- 
late pupa 

7 ( .0018) 

9 ( .0032) 

6 ( .0056) 

0 

Lick  pupa 

54  ( .0137) 

63  ( .0223) 

18  ( .0168) 

0 

Forage 

364  ( .0925) 

356  ( .1262) 

126  ( .1174) 

0 

Capture  prey 

19  ( .0048) 

26  ( .0092) 

2 ( .0019) 

0 

Return  prey  to 
nest 

19  ( .0048) 

6 ( .0021) 

0 

0 

Process  prey 

45  ( .0114) 

25  ( .0089) 

4 ( .0037) 

0 

Eat  prey 

132  ( .0335) 

109  ( .0387) 

19  ( .0177) 

2 ( .0323) 

Guard 

313  ( .0795) 

280  ( .0993) 

186  ( .1733) 

0 

Manipulate  nest 

material 

67  ( .0170) 

13  ( .0046) 

2 ( .0019) 

1 ( .0161) 

Lick  tube  wall 

27  ( .0069) 

26  ( .0092) 

12  ( .0112) 

1 ( .0161) 

Remove  refuse 
(in  tube) 

9 ( .0023) 

1 ( .0004) 

0 

0 

Remove  refuse 
(out  of  tube) 

12  ( .0030) 

19  ( .0067) 

4 ( .0037) 

0 

Carry  dead  ant 

19  ( .0048) 

20  ( .0071) 

2 ( .0019) 

0 

Total  # acts 

3936(1.0  ) 

2820(1.0  ) 

1073(1.0  ) 

62(1.0  ) 

# categories 

27 

27 

24 

8 

# individuals 

30 

15 

6 

1 

1981] 


Carlin  — Polymorphism  in  Orectognathus 


237 


estimated  total  repertory  size  for  medias  was  twenty-eight,  the  95% 
confidence  interval  (27,33);  for  majors,  twenty-seven,  with  a confi- 
dence interval  of  (24,37). 

Minor  and  media  workers  engaged  in  the  same  tasks  with 
essentially  similar  frequencies,  while  majors,  with  a smaller  reper- 
tory, also  performed  certain  acts  with  quite  different  frequencies. 
Self-grooming  was  the  commonest  act  in  all  castes.  Allogrooming 
and  regurgitation  occurred  freely  among  all  castes,  with  a tendency 
among  minors  and  medias  to  interact  with  their  own  class.  After 
self-grooming,  brood  care  and  foraging  were  the  most  frequently 
performed  acts  in  the  minor  and  media  repertories.  An  ant  was 
scored  as  “foraging”  any  time  it  left  the  nest  tube  - an  act  that  does 
not  necessarily  signify  hunting  for  food.  Though  majors  did  “forage” 
by  this  definition,  they  captured  almost  no  prey  and  returned  none 
to  the  nest.  “Processing”,  in  which  workers  tore  at,  dismembered 
and  occasionally  stung  prey  that  had  been  brought  inside  the  tube, 
was  rarely  performed  by  majors,  despite  the  seeming  usefulness  of 
their  heavy  mandibles  for  such  a task. 

The  province  of  the  majors  was  “guarding”:  walking  to  the  tube 
mouth  and  facing  outward  without  setting  foot  on  the  container 
floor;  after  self-grooming,  it  was  their  most  frequent  act.  A guarding 
ant  might  station  itself  at  the  opening  for  less  than  a minute  or  up  to 
half  an  hour.  That  this  is  in  fact  a defensive  behavior  will  be  shown 
below.  Minors  and  medias  also  guarded  in  large  numbers,  but  less 
frequently  than  they  foraged  or  attended  brood. 

Nest  maintenance  was  undertaken  almost  exclusively  by  the  small 
size  classes.  Carrying  refuse  down  the  tube,  to  be  dropped  inside  or 
just  outside  the  entrance,  was  defined  as  “in-tube  refuse  removal”, 
while  carrying  trash  out  to  corner  refuse  piles  on  the  container  floor 
(to  which  dead  ants  were  also  brought)  was  defined  as  “out-of-tube 
refuse  removal.”  “Manipulation  of  nest  material”,  that  is,  of  the 
fibers  of  the  cotton  plug,  may  not  be  an  actual  maintenane  behavior 
used  in  natural  colony  sites  (under  stones,  in  rotting  wood); 
similarly,  ants  may  lick  the  tube  wall  only  to  drink  condensation  on 
the  glass,  and  not  exhibit  any  such  behavior  in  the  wild. 

The  division  of  labor  among  minor  and  media  size  classes,  and 
the  role  of  the  major  caste,  were  better  elucidated  by  constructing 
polyethism  curves,  depicting  the  percent  contributions  of  each  caste 
to  the  total  colony  performance  of  behaviors  (figs.  3 and  4).  For 
simplicity,  certain  behavioral  categories  from  the  ethogram  were 


238 


Psyche 


[Vol.  88 


NEST  MAINTENANCE 


ATTACK  ALIEN 


Ml  ME  MA 


1981] 


Carlin  — Polymorphism  in  Orectognathus 


239 


combined,  so  that  the  polyethism  curves  indicated  represent  groups 
of  tasks.  There  was  a tendency  to  divide  those  tasks  performed 
primarily  by  small  workers  among  the  size  classes  on  the  basis  of 
size  of  objects  handled  and  task  location  (fig.  3).  Minors  performed 
most  in-nest  maintenance;  medias  performed  somewhat  more  out- 
of-nest  maintenance  than  did  minors.  Minors  contributed  most  to 
egg  care.  While  both  size  classes  attended  larvae  and  pupae,  minors 
contributed  less  to  larva  care  than  to  egg  care,  still  less  to  pupa  care, 
medias  compensating  by  putting  more  effort  into  care  of  larger 
brood. 

On  the  introduction  of  Solenopsis  workers,  the  function  of  the 
guarding  majors  became  apparent.  As  an  alien  ant  approached,  they 
spread  their  mandibles  about  120°  apart.  When  the  tip  of  the 
invader’s  head  was  within  a major’s  gape,  the  mandibles  snapped 
shut,  pinching  the  invader’s  extremity  with  sufficient  force  to  shoot 
it  away  like  a squirted  watermelon  seed.  This  very  effective  defensive 
behavior,  which  was  termed  “bouncing”,  kept  nearly  all  alien  ants 
from  gaining  entrance  to  the  nest.  Only  majors,  with  their  large 
mandibles  and  powerful  adductor  muscles,  are  equipped  to  do  this 
properly  (fig.  3).  Major  bouncers,  guarding  the  tube  mouth,  could 
propel  invaders  backward  for  up  to  8 or  9 cm;  a single  large  media 
was  able  to  bounce  an  invader,  but  not  for  very  far.  The  blunt  apical 
teeth  of  majors  pinched  but  did  not  penetrate — invaders  were  not 
injured  at  all,  just  repelled.  Ants  of  all  castes  struck  at  invaders  that 
managed  to  get  past  the  bouncers,  majors  contributing  most  to  these 
attacks  (fig.  3).  They  did  not  attempt  to  bounce  a successful  invader, 
but  instead  grabbed  it  in  their  mandibles  and  dragged  it  out, 
unharmed,  after  which  they  resumed  the  guarding  position. 

Minor  and  media  workers  foraged  in  nearly  equal  numbers,  but 
did  not  participate  equally  in  predatory  behavior.  More  prey  was 


Figure  3:  Polyethism  curves  of  nest-centered  activities,  showing  the  percent 
contribution  of  workers  of  each  caste  to  the  total  colony  performance  of  given  tasks. 
MI  = Minor  worker;  ME  = media;  M A = Major.  Some  tasks  are  composites  of  several 
behavior  categories  in  theethogram  (table  1):  Egg  care  = carry  or  manipulate  egg  + lick 
egg;  larva  care  = carry  or  manipulate  larva  + lick  larva  + regurgitate  with  larva  + feed 
larva  solid  food;  pupa  care  = carry  or  manipulate  pupa  + lick  pupa.  In-tube  nest 
maintenance  = manipulate  nest  material  + lick  tube  wall  + remove  refuse  (in  tube); 
out-of-tube  maintenance  = remove  refuse  (out  of  tube)  + carry  dead  ant.  Attacking 
alien  ants  and  “bouncing”  described  in  the  text. 


240 


Psyche 


[Vol.  88 


captured  by  medias,  while  most  was  returned  to  the  nest  by  minors 
(fig.  4);  minors  also  contributed  most  to  processing,  an  in-nest 
activity.  Medias  brought  back  only  about  one-fifth  of  the  prey  they 
caught.  It  is  possible  that  minors  play  a role  similar  to  that  of  majors 
in  Daceton , bringing  in  food  captured  by  foragers  of  another  caste, 
not  themselves  hunting  as  actively.  However,  minors  were  never 
observed  to  take  prey  away  from  medias.  They  simply  retrieved  prey 
that  medias  had  dropped,  a rather  slipshod  method  of  transferring 


FORAGING  CAPTURE  PREY 


RETURN  PREY  PROCESS  PREY 


Figure  4:  Polyethism  curves  of  predatory  behavior.  Behavior  categories  are  the 
same  as  in  the  ethogram  (table  1). 


1981] 


Carlin  — Polymorphism  in  Oreetognathus 


241 


food.  Alternatively,  the  medias  may  have  been  killing  flies  as 
trespassers  approaching  the  nest  too  closely,  rather  than  as  prey, 
whereupon  minor  foragers  picked  up  the  remains.  Majors  could  not 
leave  their  post  at  the  entrance  to  engage  in  defense  of  the  nest 
vicinity  without  exposing  the  opening  to  invaders.  Besides,  the 
bouncing  strategy  would  be  less  effective  in  the  open;  it  requires  an 
invader  to  walk  directly  into  the  defender’s  mandibles. 

Callow  workers  being  easily  recognizable  by  their  lighter  body 
color,  the  repertories  of  age  groups  within  castes  were  examined  for 
age  polyethism.  Callows  exhibited  fewer  categories  of  behavior  than 
older  adults.  As  in  Daceton  and  many  other  ant  species  (Wilson 
1971)  they  tended  to  concentrate  on  safe,  in-nest  tasks.  Callow 
majors  were  notably  more  involved  in  brood  care  than  older  majors. 

As  Brown  (1957)  had  reported  the  genus  to  be  nocturnal, 
observations  were  taken  both  during  the  day  and,  under  red  light,  at 
night.  Most  foraging  did  indeed  occur  at  night,  but  the  ants  engaged 
in  a greater  total  number  of  acts,  in  more  behavior  categories, 
during  the  day,  due  to  a diurnal  rise  in  brood  care  and  in-nest 
maintenance  activity.  This  result  suggests  that  more  complete 
behavioral  repertories  can  be  compiled  in  the  laboratory  by  studying 
ants  during  their  periods  of  “inactivity”,  when  they  are  not  investing 
so  much  of  their  effort  in  foraging. 

Discussion 

Polymorphic  workers  of  Oreetognathus  versicolor  exhibit,  all  in 
all,  a fairly  elementary  division  of  labor:  Minor  and  media  reper- 
tories are  predictably  similar,  while  majors  constitute  a distinct  caste 
on  behavioral  as  well  as  morphological  grounds.  The  minor  size 
class  contributes  most  to  in-  and  near-nest  activity,  including  prey 
retrieval;  the  medias  have  a somewhat  greater  tendency  to  perform 
out-of-nest  tasks  and  care  for  large  brood;  and  the  majors  defend. 

Even  if  the  medias  are  capturing  prey  and  dropping  it  for  minors 
to  bring  in,  the  resemblance  to  the  polyethism  pattern  of  Daceton  is 
convergent  at  most.  Daceton  majors,  not  minors,  return  prey  to  the 
nest;  Oreetognathus  majors  are  bouncers.  Daceton  minors  are 
restricted  to  brood  care,  while  medias  perform  in-nest  processing 
and  refuse  disposal  (Wilson  1962);  Oreetognathus  minors  attend  all 
these  tasks.  The  polyethism  of  O.  versicolor  is  entirely  unrelated  to 
that  of  Daceton , having  apparently  arisen  de  novo  along  with  its 
secondary  polymorphism. 


242 


Psyche 


[Vol.  88 


O.  versicolor  has  also  returned  to  polyphagy  along  with  poly- 
morphism, consistent  with  the  general  correlation  seen  in  its  tribe — 
the  only  higher  dacetine  to  return  secondarily  to  polymorphism, 
Strumigenys  loriae , is  also  polyphagous  (Brown  and  Wilson  1959). 
The  degree  of  dietary  specialization  in  the  genus  Orectognathus  as  a 
whole  may  have  been  overestimated:  A colony  of  the  monomorphic 
species  O.  clarki , collected  by  Holldobler  in  New  South  Wales, 
Australia,  was  maintained  at  a subsistance  level  on  a diet  of 
cockroach  and  mealworm  fragments  and  honey- water  (Holldobler, 
pers.  comm.).  However,  this  colony  did  not  thrive,  while  the  O. 
versicolor  colony  on  the  same  diet  flourished,  raising  many  new 
workers  and  even  males.  Clearly  O.  versicolor  does  take  non- 
collembolan  prey  more  readily;  what  is  not  clear  is  the  causality 
behind  this  correlation.  The  polyethism  of  Daceton,  at  least,  is 
associated  with  predatory  behavior.  I had  speculated  that  the  O. 
versicolor  majors  might  serve  as  “arthropod  millers”,  analogous  to 
the  seed-miller  majors  of  Solenopsis  geminata  (Wilson  1978),  their 
heavy  mandibles  used  in  processing  a variety  of  prey  with  hard 
exoskeletons.  Instead,  they  proved  to  be  soldiers;  perhaps  in 
defending  so  efficiently,  they  somehow  free  smaller  workers  to 
forage  for  different  prey  items,  which  might  require  wandering 
further  from  the  nest  vicinity  than  would  foraging  for  abundant 
collembolans.  But  this  reasoning  is  vague  at  best  and  requires 
further  investigation. 

It  is  the  major  caste  and  its  role  that  make  this  species  noteworthy, 
among  dacetines  and  among  ants  in  general.  “Bouncing”  is  a new 
kind  of  nest  defense  strategy,  ideally  suited  for  repelling  enemies  in  a 
species  whose  modified  mandibles,  designed  for  impaling  soft- 
bodied  prey,  are  of  no  use  in  fighting.  Minors  and  medias  can  be 
seriously  injured,  in  attacking  invaders  they  are  unable  to  harm. 
Bouncing  minimizes  contact  between  defenders  and  invaders,  expel- 
ling the  latter  without  a fight.  Presumably,  large  workers  of  the 
monomorphic  species  ancestral  to  O.  versicolor , modifying  slightly 
the  prey-capturing  strike  to  pinch  an  extremity  rather  than  pierce, 
found  themselves  able  to  shoot  enemies  away  for  short  distances. 
This  defense  was  so  advantageous  that  heavier  mandibles  with 
blunt,  pinching  teeth  were  strongly  selected  for,  along  with  guarding 
behavior,  eventually  producing  the  modern  majors.  Generally, 
major  castes  in  ants  serve  as  soldiers.  In  a few  species,  they  specialize 
in  physically  blocking  the  nest  opening  with  their  large  heads 


1981] 


Carlin  — Polymorphism  in  Orectognathus 


243 


(certain  Camponotus  species,  Wilson  1971;  Zacryptocerus  texanus, 
Creighton  and  Gregg  1954).  In  Zacryptocerus  varians,  which  also 
has  modified  mandibles  useless  for  fighting,  majors  use  their  saucer- 
shaped heads  to  actively  “bulldoze”  invaders  out  (Wilson  1976). 
Major  bouncers  of  O.  versicolor  are  unique  in  using  their  mandibles 
to  expel  invaders  without  injury. 

To  produce  a caste  so  specialized  for  this  form  of  defense, 
colonies  must  be  under  considerable  pressure  from  ant  species 
approximately  the  same  size  as  Solenopsis  (it  would  be  hard  to 
shoot  a larger  ant).  When  bouncing  fails,  majors  do  attack  in  a more 
conventional  manner,  as  is  seen  in  their  response  to  successful 
invaders.  (Bouncing  might  accidentally  shoot  these  further  into  the 
nest.)  It  has  recently  been  shown  (Holldobler  1982)  that  majors  also 
respond  to  alarm-recruitment  pheromones. 

Other  dacetines,  including  O.  clarki,  the  monomorphic  species 
most  closely  related  to  O.  versicolor , often  post  “sentinels”  at  nest 
entrances  (Brown  1953;  he  also  observed  occasional  “retrosalience”, 
an  ant  striking  at  a hard  surface  and  shooting  itself  backward — the 
same  motor  act  as  bouncing,  but  apparently  accidental). The  O. 
clarki  colony,  when  subjected  to  size  class  polyethism  analysis, 
revealed  a weak  division  of  labor  very  similar  to  that  of  O. 
versicolor  minors  and  medias.  It  is  easy  to  conceive  of  these  size 
classes  as  the  “primitive  caste”  (Wilson  1980)  typifying  the  mono- 
morphic ancestor  of  both  species,  from  which  increasing  defensive 
specialization  turned  the  sentinels  still  seen  in  the  former  into  the 
bouncers  of  the  latter. 


Acknowledgments 

I am  very  grateful  to  Dr.  B.  Holldobler  and  Dr.  E.  O.  Wilson,  for 
the  use  of  materials  and  the  suggestion  of  methods,  for  helpful 
advice,  for  criticizing  the  manuscript,  and  for  allowing  me  this  entry 
into  the  insect  societies.  I would  also  like  to  thank  Mark  Moffett  for 
suggestions,  assistance,  comments  on  the  manuscript  and  moral 
support,  Dr.  R.  Taylor  for  identifying  the  ants,  David  S.  Gladstein 
for  help  with  the  repertory  size  estimations  and  polyethism  curves, 
Dr.  Holldobler  for  the  photograph  in  figure  1,  Edward  Seling  for 
the  electron  photomicrographs,  and  Kathleen  Horton  for  the  word 
“bouncer”. 

This  work  was  supported  in  part  by  grants  from  the  National 


244 


Psyche 


[Vol.  88 


Science  Foundation,  #8NS80-02613,  and  from  the  National  Geo- 
graphic Society,  both  to  B.  Holldobler. 

References 


Brown,  W.  J.,  Jr. 

1953.  A revision  of  the  dacetine  ant  genus  Orectognathus.  Mem.  Queensland 
Mus.  13:84-104. 

1957.  A supplement  to  the  revision  of  the  dacetine  ant  genera  Orectognathus 
and  Arnoldidris,  with  keys  to  the  species.  Psyche  64(1):  17-29. 

Brown,  W.  L.  Jr.  and  E.  O.  Wilson 

1959.  The  evolution  of  the  dacetine  ants.  Quart.  Rev.  Biol.  34(4):278-294. 

Creighton,  W.  S.  and  R.  E.  Gregg 

1954.  Studies  on  the  habits  and  distribution  of  Cryptocerus  texanus  Santschi 
(Hymenoptera:  Formicidae).  Psyche  61(2):4 1 —57. 

Fagen,  R.  and  R.  Goldman 

1977.  Behavioral  catalogue  analysis  methods.  Anim.  Behav.  25:261-274. 

Holldobler,  B. 

1982.  Trail  communication  in  the  dacetine  ant  Orectognathus  versicolor. 
Psyche  88:245-257. 

Taylor,  R.  W. 

1977.  New  ants  of  the  genus  Orectognathus,  with  a key  to  the  known  species. 
Austr.  J.  Zool.  25:581-612. 

1979.  New  Australian  ants  of  the  genus  Orectognathus,  with  summary  descrip- 
tion of  the  twenty-nine  known  species  (Hymenoptera:  Formicidae). 
Austr.  J.  Zool.  27:773-788. 

Wilson,  E.  O. 

1962.  Behavior  of  Daceton  armigerum  (Latreille),  with  a classification  of  self- 
grooming movements  in  ants.  Bull.  Mus.  Comp.  Zool.  127(7):403-421. 

1971.  The  Insect  Societies.  Belknap  Press  of  Harvard  Univ.  Press,  Cambridge, 
Mass. 

1976.  A social  ethogram  of  the  Neotropical  arboreal  ant  Zacryptocerus  varians 
(Fr.  Smith).  Anim.  Behav.  24:(2):354-363. 

1978.  Division  of  labor  in  fire  ants  based  on  physical  castes  (Hymenoptera: 
Formicidae:  Solenopsis ).  J.  Kansas  Entom.  Soc.  51(4):6 15-636. 

1980.  Caste  and  division  of  labor  in  leaf-cutting  ants  (Hymenoptera:  Formici- 
dae: Atta).  Behav.  Ecol.  Sociobiol.  7:143-156. 


TRAIL  COMMUNICATION  IN  THE  DACETINE  ANT 
ORECTOGNATHUS  VERSICOLOR 
(HYMENOPTERA:  FORMICIDAE)* 

By  Bert  Holldobler 

Department  of  Organismic  and  Evolutionary  Biology, 
Harvard  University,  Cambridge,  Mass.  02138 

Although  division  of  labor  within  two  dacetine  species  has  been 
studied  at  length  (Wilson  1962;  Carlin  1982),  very  little  has  hitherto 
been  reported  on  social  communication  in  the  Dacetini,  a myrmi- 
cine  tribe  of  nearly  200  known  species  (Brown  and  Wilson  1959; 
Wilson  1962).  Foraging  habits  have  also  been  studied  in  several 
species  (for  review  see  Brown  and  Wilson  1959;  Wilson  1962).  As 
now  known,  the  dacetines  seem  to  be  individual  foragers;  recruit- 
ment to  food  sources  and  cooperation  during  retrieval  of  prey  have 
not  been  observed.  It  is  therefore  of  some  interest  that  we  have 
recently  discovered  trail  laying  and  trail  following  in  the  dacetine 
species  Orectognathus  versicolor.  Experiments  in  the  laboratory 
further  indicate  that  trail  communication  may  play  an  especially 
important  role  during  nest  emigrations. 

Material  and  Methods: 

A queenright  colony  of  O.  versicolor  was  collected  from  rotting 
wood  near  Eungella,  North  Queensland  (Australia)  and  housed  in  a 
glass  tube  (</>  1 cm),  with  water  trapped  at  its  bottom  behind  a 
cotton  plug.  The  nest  tube  was  laced  into  an  arena  (45  X 30  cm)  in 
which  small  pieces  of  cockroaches  ( Nauphoeta  cinerea),  chopped 
meal  worms  ( Tenebrio  molitor ),  several  species  of  small  flies  and 
honey  water  were  provided  as  food.  The  colony  developed  very  well 
under  these  conditions,  and  when  the  experiments  began  (4  weeks 
after  collection)  it  contained  one  queen,  80  workers  (42  minors,  27 
medias,  11  majors  (see  Carlin,  1982)),  14  freshly  eclosed  males,  and 
brood  of  all  stages. 

For  histological  investigations  live  specimens  were  fixed  in  Car- 
noy  (Romeis  1948),  embedded  in  methyl-methacrylate  and  sectioned 
8 fj,  thick  with  a Jung  Tetrander  I microtome  (Rathmayer  1962).  The 


♦Manuscript  received  by  the  editor  December  1,  1981. 


245 


246 


Psyche 


[Vol.  88 


staining  was  Azan  (Heidenhain).  The  SEM  pictures  were  taken  with 
an  AMR  1000  A scanning  electron  microscope. 

Additional  methodological  details  will  be  given  with  the  descrip- 
tion of  the  individual  experiments. 

Results: 

As  demonstrated  by  Carlin  (1982)  most  of  the  foraging  in  O. 
versicolor  is  conducted  by  the  minor  and  medium  worker  castes;  the 
majors  function  primarily  or  entirely  as  a defense  caste,  for  which 
they  have  unique  modifications  in  the  form  of  the  mandibles. 
Although  workers  of  O.  versicolor  seem  to  forage  individually,  our 
observations  in  the  laboratory  indicate  that  some  sort  of  social 
facilitation  might  be  involved  in  stimulating  foraging  activity  in  the 
colony. 

Often  not  more  than  1-3  workers  roamed  the  foraging  arena.  But 
when  suddently  30-50  flightless  Drosophila  flies  were  released  into 
the  arena,  and  the  first  one  or  two  foragers  had  returned  with 
captured  prey  to  the  nest,  the  number  of  workers  leaving  the  nest 
tube  and  venturing  into  the  foraging  arena  increased  markedly.  We 
did  not,  however,  observe  the  foragers  performing  any  motor 
display  inside  the  nest,  which  might  have  stimulated  the  nestmates. 


Fig.  1.  Part  of  the  colony  of  Orectognathus  versicolor,  the  three  worker  castes 
(minors,  medias,  majors),  males,  and  different  developmental  stages. 


1981]  Holldobler  — Communication  in  Orectognathus 


247 


nor  did  it  appear  that  workers  leaving  the  nest  followed  chemical 
trails. 

On  the  other  hand,  trail  following  was  very  obvious  when  the 
colony  or  fragments  of  the  colony  were  forced  to  move  to  a new  nest 
site.  For  example,  when  we  shook  the  colony  out  of  the  nest  tube 
into  the  arena,  which  had  been  provided  with  a new  papered  floor 
before  each  experiment,  the  “homeless”  colony  soon  gathered  at  one 
spot,  where  it  was  closely  guarded  by  members  of  the  major  worker 
caste  (Fig.  1).  After  varying  intervals  (sometimes  lasting  more  than 
one  hour),  some  of  the  minors  and  medias  began  exploring  the 
arena,  and  eventually  they  discovered  a nest  tube  that  had  been 
provided  at  the  edge  of  the  arena  (usually  30-35  cm  away  from  the 
displaced  colony).  After  exploring  the  nest  tube,  some  of  the  ants 
returned  to  the  colony,  and  after  a while  they  often  moved  again  to 
the  nest  tube  to  continue  to  explore  it  thoroughly. 

Usually  this  procedure  was  repeated  several  times,  before  the  first 
signs  of  a colony  movement  could  be  observed.  It  occurred  when 
several  additional  minors  and  medias  departed  from  the  colony  and 
traveled  directly  to  the  new  nest.  Their  straight  orientation  and  the 
fact  that  during  running  they  kept  the  tips  of  their  antennae  close  to 
the  ground,  suggested  that  these  ants  were  following  a chemical 
trail.  Soon  afterwards  the  traffic  between  the  “homeless”  colony  and 
the  newly  discovered  nest  tube  increased  leading  finally  to  a full- 
scale  colony  emigration. 

All  three  worker  castes  were  involved  in  transporting  brood, 
callow  workers,  and  males  to  the  new  nest,  although  the  minors 
handled  eggs  and  small  larvae  preferentially  while  the  medias  and 
majors  concentrated  on  large  larvae,  pupae  and  adults  (Fig.  2). 
Usually  the  queen  moved  during  the  early  phase  of  the  colony 
movement  and  always  traveled  on  her  own.  On  the  other  hand,  the 
males  were  always  carried  by  the  workers  (Fig.  2),  usually  not  before 
most  of  the  brood  had  already  been  moved.  Only  once  did  we  see  a 
fully  pigmented  worker  being  carried  by  a nestmate.  The  trans- 
ported individual  was  grasped  dorsally  at  the  head  and  lifted 
upwards  with  gaster  tip  pointing  forwards;  it  had  the  appendaes 
folded  in  the  pupal  position.  All  ants  traveled  along  a relatively 
narrow  route  between  colony  and  new  nest  site.  This  strongly 
suggested  that  O.  versicolor  employs  chemical  trail  communication 
during  the  process  of  colony  migration.  The  following  experiments 


gpggp 


248 


Psyche 


[Vol.  88 


Fig.  2.  Colony  emigration  in  Orectognathus  versicolor,  (a)  major  transporting  larva;  (b)  major  transporting  pupa;  (c)  media 
transporting  larva,  accompanied  by  a minor;  (d)  media  transporting  male. 


1981]  Holldobler  — Communication  in  Orectognathus 


249 


were  designed  to  localize  the  anatomical  source  of  a possible  trail 
pheromone  in  O.  versicolor. 

Close-up  cinematography  and  photography  revealed  that  many 
Orectognathus  workers,  when  moving  back  and  forth  between  the 
displaced  colony  and  the  nest  tube,  touched  their  abdominal  tips 
intermittently  to  the  ground,  presumably  depositing  droplets  of  trail 
pheromone.  Three  major  exocrine  glands  open  at  or  near  the 
abdominal  tip  of  O.  versicolor  workers:  the  poison  gland  and 
Dufour’s  gland,  both  of  normal  size,  and  a relatively  large  pygidial 
gland,  which  opens  between  the  6th  and  7th  abdominal  tergites  (Fig. 

3). 

Most  myrmicine  ants  have  a more  or  less  well  developed  pygidial 
gland  (Holldobler  et  al.  1976,  Holldobler  and  Engel  1978;  Kugler 
1978),  but  in  O.  versicolor  this  gland  is  more  complex  than  usually 
found  in  Myrmicinae.  It  more  closely  resembles  the  pygidial  gland 
of  some  ponerine  species,  for  example  Pachycondyla  laevigata , in 
which  it  serves  as  the  source  of  a trail  pheromone.  The  paired 
reservoir  sacs  (invaginations  of  the  intersegmental  membrane  be- 
tween the  6th  and  7th  tergites)  are  filled  with  a clear,  lightly 
brownish  liquid.  Several  ducts  lead  from  paired  clusters  of  glandular 
cells  into  the  reservoir,  penetrating  the  intersegmental  membrane 
(Fig.  4).  The  cuticle  of  the  7th  tergite  has  a grooved  structure  (Fig. 
5),  underneath  of  which  is  a large  glandular  epithelium  (Fig.  4). 
Orectognathus  versicolor  workers,  when  engaged  in  trail  laying 
behavior,  usually  hold  their  gaster  in  an  almost  vertical  position. 
This  brings  the  opening  of  the  pygidial  gland  very  close  to  the  floor 
so  that  part  of  the  grooved  structure  on  the  7th  tergite  can  be  easily 
put  in  contact  with  the  surface  of  the  ground. 

In  the  next  series  of  experiments  we  tested  the  trail  following 
response  of  O.  versicolor  to  artificial  trails  drawn  with  glandular 
secretions  of  the  poison  gland,  Dufour’s  gland  and  pygidial  gland. 
The  glands  were  dissected  out  of  freshly  killed  workers  and  for  each 
trail  test  one  gland  of  a kind  was  crushed  on  the  tip  of  a hardwook 
applicator  stick  and  smeared  once  along  a 20-cm-long  pencil  line. 
The  trails  were  made  to  originate  either  from  the  entrance  of  the 
nest  tube  or  from  the  periphery  of  the  clustered  colony,  which  had 
previously  been  shaken  out  of  their  nest  tube  into  the  test  arena.  As 
a control  a second  trail  was  offered  simultaneously  which  was 
derived  either  from  a droplet  of  water  or  from  one  of  the  other 


250 


Psyche 


[Vol.  88 


Fig.  3.  Gaster  of  a media  of  Orectognathus  versicolor,  (a)  SEM  picture;  arrow 
indicates  opening  of  pygidial  gland,  (b)  Sagittal  section  through  gaster  of  a media. 
PG  = pygidial  gland;  R = reservoir  of  pygidial  gland;  D = Dufour’s  gland;  S = 
stinger. 


1981]  Holldobler  — Communication  in  Orectognathus 


251 


Fig.  4.  (a)  Sagittal  section  through  pygidial  gland  of  a media  of  Orectognathus 
versicolor.  PG  = pygidial  gland;  R = reservoir  of  pygidial  gland;  GE  = glandular 
epithelium,  (b)  Close-up  of  glandular  epithelium  (GE)  under  the  cuticular  structure 
(CS)  of  pygidium. 


252 


Psyche 


[Vol.  88 


Fig.  5.  Above:  SEM  pictures  of  the  gaster  tip  of  a media  of  Orectognathus 
versicolor.  The  slightly  extruded  stinger  is  visible.  It  is  surrounded  by  long  sensory 
setae  (confirmed  by  histology;  probably  mechanoreceptors)  on  the  edge  of  the 
pygidium  and  7th  sternum. 

Below:  Groved  cuticular  structures  on  the  pygidium  associated  with  the  pygidial 
gland.  This  structure  is  usually  covered  by  the  preceding  6th  tergite. 


1981]  Holldobler  — Communication  in  Orectognathus 


253 


glands.  All  ants  following  the  trails  to  the  end  during  a 5-minute 
period  were  counted. 

As  can  be  seen  from  table  1,  trails  drawn  either  with  crushed 
pygidial  glands  or  poison  glands  elicited  a precise  trail  following 
behavior  in  all  three  worker  castes  (Fig.  6),  but  the  ants  did  not 
respond  to  trails  drawn  with  crushed  Dufour’s  glands.  We  noticed, 
however,  several  differences  in  the  reaction  of  the  ants  to  poison 
gland  trails  and  pygidial  gland  trails.  (1)  When  both  trails  were 
offered  simultaneously,  starting  at  the  periphery  of  a “homeless” 
colony,  significantly  more  workers  (Tab.  1)  carrying  brood  moved 
along  the  poison  gland  trail.  (2)  In  all  tests  the  poison  gland  trail  was 
the  more  effective  one  and  lasted  over  a longer  period  of  time.  After 
5 minutes  the  ants’  response  to  pygidial  gland  trails  had  almost 
vanished,  whereas  they  were  still  strongly  following  the  trail  drawn 
with  poison  gland  material.  In  fact,  poison  gland  trails  presented  to 
the  ants  24  hours  after  they  were  drawn  were  still  effective  as 
orientation  cues  for  emigrating  O.  versicolor  workers.  (3)  Although 
we  could  not  detect  a preference  for  either  trails  drawn  with  poison 
glands  or  pygidial  glands,  ants  moving  along  the  pygidial  gland  trail 
seemed  to  gape  their  mandibles  more  frequently  than  ants  moving 
along  poison  gland  trails. 

From  these  observations  we  conclude  that  the  trail  pheromones 
serve  different  functions.  The  poison  gland  trail  is  obviously 
employed  during  nest  emigrations,  where  it  serves  as  a stimulative 
recruitment  signal  as  well  as  a longer  lasting  orientation  cue.  On  the 
other  hand,  the  pygidial  gland  trail  probably  functions  as  a relatively 
short  lasting  alarm-recruitment  signal,  channeling  workers  to  areas 
of  disturbance  near  the  nest.  It  is  also  possible  that  the  pygidial 
gland  pheromone  is  discharged  by  successful  foragers  when  they 
return  to  the  nest,  which  might  cause  the  social  facilitation  of  the 
foraging  activity  mentioned  above.  In  fact,  when  a crushed  pygidial 
gland  is  presented  inside  the  nest  tube,  it  elicits  more  excitement  in 
the  workers  than  any  other  glandular  secretions  (mandibular  gland, 
poison  gland,  Dufour’s  gland),  causing  several  workers  to  move 
toward  the  nest  entrance. 

All  three  worker  castes  have  the  same  glandular  equipment  and 
their  secretions  release  the  same  behavioral  responses. 


254 


Psyche 


[Vol.  88 


Fig.  6.  Trail  test  with  Orectognathus  versicolor.  Artificial  trails  drawn  with 
secretions  from  the  poison  gland  (PoG)  and  Dufour’s  gland  (C),  both  originating  at 
the  opening  of  the  nest  tube,  are  offered  simultaneously.  All  workers  follow  the 
poison  gland  trail. 


1981]  Holldobler  — Communication  in  Orectognathus  255 
Discussion: 

It  has  been  well  documented  that  many  species  of  the  sub-family 
Myrmicinae  employ  secretions  from  the  glands  associated  with  the 
sting  apparatus  (poison  gland,  Dufour’s  gland)  for  chemical  trail 
communication  and  orientation  (for  review  see  Wilson  1971;  Holl- 
dobler  1978).  This  paper  presents  the  first  evidence  of  the  phenom- 
enon in  the  myrmicine  tribe  Dacetini.* 

In  the  dacetine  species  Orectognathus  versicolor  trails  laid  with 
poison  gland  secretions  function  both  as  recruitment  and  orienta- 
tion signals  during  nest  emigration.  In  fact,  many  dacetine  species 
seem  to  construct  relatively  simple  nests  in  soil  or  rotting  wood  and 
it  is  easily  conceivable  that  colonies  frequently  abandon  their  nests 
and  move  to  new  nest  sites.  More  surprising,  however,  was  the 
discovery  that  this  species  possesses  a pygidial  gland  whose  struc- 
ture closely  resembles  that  of  the  pygidial  gland  of  some  ponerine 
species.  The  secretions  of  this  gland  can  also  function  as  a recruit- 
ment trail  pheromone  in  O.  versicolor. 


Table  1.  Number  of  workers  following  artificial  trails  within  5 min.  periods.  The 
means  and  standard  deviations  are  given. 


Trails  presented  at  nest  entrance  (n  = 4) 


Dufour’s 

water 

Poison 

water 

Pygidial 

water 

gland 

control 

gland 

control 

gland 

control 

0 

0 

12.7  ± 3.8 

0 

8.3  ± 2.8 

0 

Trails  presented  simultaneously  at  periphery  of  clustered  colony  (n  = 5) 


Dufour’s  gland 

Poison  gland 

Pygidial  gland 

with 

brood 

without 

with  brood  without 

with  brood 

without 

incl. 

males 

brood 

inch  males  brood 

incl.  males 

brood 

0 

0 

7.8  ± 3.8  7.0  ± 2.2 

2.0  ± 1.6 

9.8  ± 3.1 

*Blum  and  Portocarrero  (1966)  demonstrated  that  three  attine  ant  genera  follow 
trails  drawn  with  poison  gland  secretions  of  Daceton  armigerum,  but  they  could  not 
demonstrate  trail  following  behavior  in  Daceton. 


256 


Psyche 


[Vol.  88 


From  recent  investigations  we  know  that  the  pygidial  gland  is 
quite  common  in  the  Myrmicinae  (Kugler  1978;  Holldobler  and 
Engel  1978).  In  at  least  two  species  its  secretions  serve  as  an  alarm 
pheromone  (Holldobler  et  al.  1976;  Kugler  1979).  On  the  other 
hand,  in  several  ponerine  species  the  pygidial  gland  secretions  have 
been  demonstrated  to  function  as  a recruitment  pheromone  during 
tandem  running  (Holldobler  and  Traniello  1980a)  or  trail  communi- 
cation (Maschwitz  and  Schonegge  1977;  Holldobler  and  Traniello 
1980b).  From  our  findings  in  O.  versicolor  it  appears  now  that  this 
ponerine  trait  might  have  been  preserved  in  the  Dacetini,  whose 
origin  presumably  dates  back  to  early  Tertiary  times  (Brown  and 
Wilson  1959).  If  this  speculation  is  right,  we  should  expect  that  the 
most  primitive  dacetine  species,  Daceton  armigerum  (Brown  and 
Wilson  1959;  Wilson  1962),  has  a well  developed  pygidial  gland, 
resembling  closely  that  found  in  many  ponerine  ants,  and  its 
secretions  presumably  serve  as  an  alarm-recruitment  pheromone.  In 
fact,  Wilson  (1962)  observed  that  workers  of  D.  armigerum  often 
moved  to  areas  of  excitement  and  when  a worker  just  had  dis- 
covered prey  it  moved  in  “excited  broken  running  patterns”  by 
which  other  ants  in  the  vicinity  might  be  attracted.  Wilson  (1962, 
1971)  hypothesized  that  this  running  pattern  might  serve  as  a 
communicative  signal  of  the  kind  of  “Stager’s  kinopsis”,  i.e.  the 
large-eyed  Daceton  workers  might  respond  to  the  visual  stimuli 
produced  by  the  excitedly  moving  nestmate.  We  have  now  to 
consider  the  possibility  that  a Daceton  huntress  which  pursues  a 
prey,  discharges  a short-range  recruitment  pheromone  from  the 
pygidial  gland,  and  that  consequently  the  attraction  of  other 
huntresses  in  the  close  vicinity  is  caused  by  this  chemical  signal. 

Acknowledgments 

I would  like  to  thank  E.  O.  Wilson  for  reading  and  commenting 
on  the  manuscript.  Hiltrud  Engel  and  E.  Seling  (SEM-Lab  of  the 
MCZ,  Harvard)  provided  valuable  technical  assistance.  Many 
thanks  to  R.  W.  Taylor  and  the  Division  of  Entomology  of  CSIRO, 
Canberra  (Australia)  for  their  support  and  hospitality.  R.  W.  Taylor 
identified  the  ants;  voucher  specimens  were  deposited  in  the  Austral- 
ian National  Insect  Collection,  Canberra.  This  work  was  supported 
by  grants  of  the  National  Geographic  Society,  National  Science 


1981]  Holldobler  — Communication  in  Orectognathus 


257 


Foundation  (BNS  80-02613)  and  by  a fellowship  from  the  John 
Simon  Guggenheim  Foundation. 

References 

Blum,  M.  S.  and  C.  A.  Portocarrero 

1966  Chemical  releases  of  social  behavior.  X.  An  attine  trail  substance  in  the 
venom  of  a non-trail  laying  myrmicine,  Daceton  armigerum.  Psyche 
(Cambridge)  73:  150-155. 

Brown,  W.  L.  and  E.  O.  Wilson 

1959  The  evolution  of  the  dacetine  ants.  Quarterly  Rev.  Biology  34,  278-294. 

Carlin,  N.  F. 

1982  Polymorphism  and  division  of  labor  in  the  dacetine  ant  Orectognathus 
versicolor  (Hymenoptera:  Formicidae)  Psyche  (Cambridge)  88:231-244. 

Holldobler,  B.  and  H.  Engel 

1978  Tergal  and  sternal  glands  in  ants.  Psyche  (Cambridge)  85,  285-330. 

Holldobler,  B.  and  J.  Traniello 

1980a  Tandem  running  pheromone  in  ponerine  ants.  Naturwissenschaften  67, 
360. 

Holldobler,  B.  and  J.  F.  A.  Traniello 

1980b  The  pygidial  gland  and  chemical  recruitment  communication  in  Pachy- 
condyla  (=  Termitopone ) laevigata.  J.  Chem.  Ecology  6,  883-893. 

Holldobler,  B.,  R.  Stanton  and  H.  Engel 

1976  A new  exocrine  gland  in  Novomessor  (Hymenoptera:  Formicidae)  and 
its  possible  significance  as  a taxonomic  character.  Psyche  83,  32-41. 

Kugler,  C. 

1978  Pygidial  glands  in  the  myrmicine  ants  (Hymenoptera:  Formicidae). 
Insectes  sociaux  25,  267-274. 


1979  Alarm  and  defense:  a function  for  the  pygidial  gland  of  the  myrmicine 
ant,  Pheidole  biconstricta.  Annals  Entomological  Society  America  72, 
532-536. 

Maschwitz,  U.  and  P.  Schonegge 

1977  Recruitment  gland  of  Leptogenys  chinensis.  Naturwissenschaften  64, 
589-590. 

Rathmayer,  W. 

1962  Methylmetacrylat  als  Einbettungsmedium  fur  Insekten.  Experientia 
(Basel)  18,  47-48. 

Romeis,  B. 

1948  Mikroskopische  Technik,  Miinchen  1948. 

Wilson,  E.  O. 

1962  Behavior  of  Daceton  armigerum  (Latreille),  with  a classification  of  self- 
grooming movements  in  ants.  Bull.  Mus.  Comp.  Zool.  Harvard  Univ. 
127,  401-422. 


1971  The  insect  societies.  The  Belknap  Press  of  Harvard  University  Press, 
Cambridge,  Mass. 


FRANCIS  WALKER  TYPES  OF,  AND  NEW  SYNONYMIES 
FOR,  NORTH  AMERICAN  HYDROPSYCHE  SPECIES 
(TRICHOPTERA,  HYDROPSYCHIDAE)* 

By  Andrew  P.  Nimmo 

Department  of  Entomology,  University  of  Alberta 
Edmonton,  Alberta,  Canada  T6G  2E3 

Introduction 

Recently,  while  assembling  the  manuscript  of  a handbook  to  the 
Arctopsychidae  and  Hydropsychidae  of  Canada,  I had  occasion  to 
examine  the  female  holotypes  of  three  species  of  Hydropsyche 
described  from  North  America  by  Francis  Walker  (1852).  Betten  & 
Mosely  (1940)  record  these  types,  but  in  line  with  the  still  all  too 
prevalent  practice  regarding  unassociated  female  Trichoptera,  they 
did  not  illustrate  the  genitalia.  If  they  had  provided  figures  for  these 
types  the  true  identities  of  at  least  two  North  American  species  of 
Hydropsyche  would  have  been  long  since  established.  The  genital 
segments  of  these  types  are  illustrated  here  for  the  first  time. 

Hydropsyche  confusa  (Walker) 

Philopotamus  confusus  Walker,  1852:  103. 

Hydropsyche  confusa : Milne,  1936:  61;  Betten  & Mosely,  1940:  21. 
Hydropsyche  separata  Banks,  1936:  129;  Ross  & Spencer,  1952:  46 
(as  synonym  of  H.  guttata  Pictet);  Smith,  1979:  10.  new 

SYNONYMY. 

Hydropsyche  guttata  Pictet:  Schuster  & Etnier,  1978:  126. 
Hydropsyche  corbeti  Nimmo,  1966:  688;  Schuster  & Etnier,  1978: 
126  (as  synonym  of  H.  guttata  Pictet),  new  synonymy. 

Fig.  2 depicts  the  genitalia  of  the  female  holotype  of  H.  confusa 
(Walker).  Fig.  1 is  of  the  genitalia  of  a female  which  has  been 
recognised  as  belonging  to  H.  separata  Banks.  The  rather  obscure 
locality  information  recorded  by  Betten  & Mosely  (1940)  indicates 
that  the  type  of  confusa  was  collected  in  the  western  Northwest 
Territories,  adjacent  to  the  northern  boundary  of  Alberta.  The 
female  of  separata  was  collected  at  Empress  in  southeastern  Alberta. 


* Manuscript  received  by  the  editor  October  14,  1981. 


259 


260 


Psyche 


[Vol.  88 


Figures  1-5.  Fig.  1.  Hydropsyche  separata  Banks  [=confusa  (Walker)] — genital 
segments  of  female,  lateral  aspect.  Fig.  2.  H.  confusa  (Walker) — genital  segments 
of  female  holotype,  lateral  aspect.  Fig.  3.  H.  recurvata  Banks  [ =alternans  (Walker)] 
— genital  segments  of  female,  lateral  aspect.  Fig.  4.  H.  alternans  (Walker) — genital 
segments  of  female  holotype,  lateral  aspect.  Fig.  5.  H.  reciproca  (Walker) — genital 
segments  of  female  holotype,  lateral  aspect. 


1981] 


Nimmo  — Walker  Types  of  Hydropsvche 


261 


While  not  absolutely  identical  (the  differences  may  be  attributed 
to  geographic  variation,  laboratory  treatment,  and  observer  vari- 
ables), these  two  specimens  are  much  more  similar  to  each  other 
than  either  is  to  the  females  of  the  most  nearly  related  species  ( H . 
betteni  Ross),  and  I judge  them  to  be  conspecific.  H.  separata  is 
therefore  synonymised  with  H.  confusa  which  has  clear  priority. 

Smith  (1979)  quotes  me  as  considering  the  possibility  that  H. 
corbeti  Nimmo  may  be  a synonym  of  separata.  Prior  to  my 
examination  of  the  type  of  confusa  I had  decided  that  it  was. 
However,  it  must  now  be  entered  as  a synonym  of  confusa. 

In  view  of  the  taxonomic  history  of  this  species  Walker  must  be 
attributed  with  remarkable  insight  in  naming  it  confusa. 


Hydropsyche  alternans  (Walker) 

Philopotamus  alternans  Walker,  1852:  104. 

Hydropsyche  alternans:  Vorhies,  1909:  707  (sp.indet.);  Betten,  1934: 
185  (prob.  H.  bifida ). 

Philopotamus  indecisus  Walker,  1852:  104. 

Hydropsyche  indecisa:  Betten  & Mosely,  1940:  20  (as  synonym  of 
H.  alternans ). 

Hydropsyche  slossonae  var.  recurvata  Banks,  1914:  253. 
Hydropsyche  recurvata : Betten,  1934:  190;  Milne,  1936  73  (as 
synonym  of  H.  slossonae );  Ross,  1944:  99.  new  synonymy. 
Symphitopsyche  recurvata : Schuster  & Etnier,  1978:  34. 
Hydropsyche  codona  Betten,  1934:  187;  Milne,  1936:  73  (as  syn- 
onym of  H.  slossonae );  Ross,  1938:  18  (as  synonym  of  H. 
recurvata). 

Fig.  4 depicts  the  genitalia  of  the  holotype  female  of  H.  alternans 
(Walker)  (from  the  Albany  R.,  far  northern  Ontario).  Fig.  3 depicts 
the  genitalia  of  a female  (from  Wandering  River,  northeastern 
Alberta)  which  has  been  recognised  as  belonging  to  H.  recurvata 
Banks.  Again,  these  two  females  are  not  precisely  identical,  for 
possible  reasons  similar  to  those  given  under  H.  confusa  above.  I 
judge  these  two  specimens  to  be  conspecific.  H.  recurvata  is 
therefore  synonymised  with  H.  alternans  which  has  priority. 


262 


Psyche 


[Vol.  88 


Hydropsyche  reciproca  (Walker) 

Philopotamus  reciprocus  Walker,  1852:  104. 

Hydropsyche  reciproca : Betten  & Mosely,  1940:  22. 

The  genitalia  of  the  female  holotype  do  not  correspond  to  those 
of  any  other  species  known  to  me.  They  are  illustrated  here,  for  the 
first  time,  for  the  future  reference  of  students  of  North  American 
Hydropsyche  species.  The  type  locality  is  given  simply  as  ‘North 
America’. 


Summary 

The  female  holotypes  of  Hydropsyche  confusa  (Walker),  H. 
alternans  (Walker),  and  H.  reciproca  (Walker),  all  from  North 
America,  were  examined,  and  illustrations  of  the  genitalia  are 
provided  for  the  first  time.  It  is  concluded  that  H.  separata  Banks  is 
conspecific  with  H.  confusa  (Walker),  and  that  H.  recurvata  Banks 
is  conspecific  with  H.  alternans  (Walker).  The  Walker  names  have 
priority.  H.  reciproca  (Walker)  cannot  yet  be  equated  with  any 
other  known  species. 


Acknowledgments 

Loan  of  the  three  types  examined  here  was  very  kindly  arranged 
by  Peter  Barnard  of  the  Entomology  Dept,  British  Museum  (Na- 
tural History).  This  paper  is  an  offshoot  of  work  on  preparation  of  a 
Handbook  to  the  Arctopsychidae  and  Hydropsychidae  of  Canada, 
which  was  supported  by  a contract  from  the  Canada  Dept  of 
Agriculture.  The  work  was  carried  out  in  the  Dept,  of  Entomology, 
University  of  Alberta.  Steve  Ashe  of  that  Department  read,  and 
commented  on  the  manuscript.  Publication  costs  were  met  by 
George  E.  Ball  of  the  same  Department,  from  grant  NRC  A- 1399 
held  by  him. 

To  all,  my  sincere  thanks. 


1981] 


Nimmo  — Walker  Types  of  Hydropsyche 


263 


References 


Banks,  N. 

1914.  American  Trichoptera — notes  and  descriptions.  Can.  Ent.  46:  252-258. 
1936.  Notes  on  some  Hydropsychidae.  Psyche,  Camb.  43:  126-130. 

Betten,  C. 

1934.  The  Caddis  Flies  or  Trichoptera  of  New  York  State.  Bull.  N.Y.  St.  Mus., 
Albany  292:1-576. 

Betten,  C.  and  M.  E.  Mosely 

1940.  The  Francis  Walker  Types  of  Tricoptera  in  the  British  Museum.  British 
Museum  (Natural  History),  London,  vi+248  pp. 

Milne,  L.  J. 

1936.  Studies  in  North  American  Trichoptera.  3:56-128,  Cambridge,  Mass. 
Nimmo,  A.  P. 

1966.  A list  of  Trichoptera  taken  at  Montreal  and  Chambly,  Quebec,  with 
descriptions  of  three  new  species.  Can.  Ent.  98:688-693. 

Ross,  H.  H. 

1938.  Lectotypes  of  North  American  Caddis  Flies  in  the  Museum  of  Compara- 
tive Zoology.  Psyche,  Camb.  45:1-61. 

1944.  The  Caddis  Flies,  or  Trichoptera,  of  Illinois.  Bull.  111.  St.  Nat.  Hist. 
Surv.,  Urbana,  23,  326  pp. 

Ross,  H.  H.  and  G.  J.  Spencer 

1952.  A preliminary  list  of  the  Trichoptera  of  British  Columbia.  Proc.  Ent. 
Soc.  Br.  Columb.  48:43-51. 

Schuster,  G.  A.  and  D.  A.  Etnier 

1978.  A manual  for  the  identification  of  the  larvae  of  the  Caddisfly  genera 
Hydropsyche  Pictet  and  Symphitopsyche  Ulmer  in  eastern  and  central 
North  America  (Trichoptera:  Hydropsychidae).  Environmental  Moni- 
toring and  Support  Laboratory,  Cincinnati  EPA-600/ 4-78-060:  xii+129 

pp. 

Smith,  D. 

1979.  The  larval  stage  of  Hydropsyche  separata  Banks  (Trichoptera:  Hydro- 
psychidae). Pan-Pacif.  Ent.  55:10-20. 

Vorhies,  C.  T. 

1909.  Studies  on  the  Trichoptera  of  Wisconsin.  Trans.  Wis.  Acad.  Sci.  Arts 
Lett.  16:647-738. 

Walker,  F. 

1852.  Catalogue  of  the  specimens  of  neuropterous  insects  in  the  collections  of 
the  British  Museum.  Part  1.  192  pp.  London. 


TERRITORIALITY,  NEST  DISPERSION, 

AND  COMMUNITY  STRUCTURE  IN  ANTS. 

By  Sally  C.  Levings1  and  James  F.  A.  Traniello2 
Introduction 

The  dispersion  patterns  of  ant  colonies  have  been  reported  for  a 
variety  of  species  having  very  different  ecological  characteristics 
(Pontin  1961;  Yasuno  1963,  1964a, b,  1965;  Brian  1964;  Brian  et  al. 
1965,  1966;  Greenslade  1971;  Room  1971,  1975a, b;  Bernstein  and 
Gobbel  1979;  Levings  and  Franks  1982),  and  typically,  spacing 
studies  involve  discussions  of  territoriality.  Recently,  Holldobler 
and  Lumsden  (1980),  using  a cost/ benefit  approach,  examined  the 
importance  of  the  economic  defensibility  of  territories  and  its 
influence  on  the  use  of  space  and  dispersion  patterns.  Holldobler 
(1974,  1976a,  1979a, b)  demonstrated  the  relationship  between  re- 
source distribution,  territory  shape  and  nest  spacing.  These  studies 
also  emphasize  that  in  order  to  understand  thoroughly  territoriality 
and  other  intra-  and  interspecific  relationships,  it  is  necessary  to 
comprehend  the  role  of  social  design  in  the  establishment  and 
maintenance  of  territory.  Without  such  a combined  approach  of 
behavior  and  ecology,  it  is  difficult  to  assess  accurately  the  signifi- 
cance of  territoriality  in  social  species  such  as  ants. 

In  many  studies  there  have  been  problems  in  the  application  of 
the  term  territoriality  and  discrepancies  in  the  identification  of 
territorial  phenomena.  Terms  describing  the  use  of  foraging  area 
such  as  territory  and  home  range  have  been  rather  poorly  defined 
and  vary  in  meaning  between  authors.  Territory  to  some  authors 
denotes  a defended  area  (Baroni-Urbani  1979;  Holldobler  1974, 
1976a;  Holldober  and  Wilson  1977a, b;  Holldobler  and  Lumsden 
1980)  whereas  to  others  it  is  synonymous  with  home  range  or  is 
casually  used  (Dobrzanska  1958,  1966).  There  are  also  problems 
with  the  application  of  information  on  territoriality  in  the  interpre- 
tation of  spacing  patterns.  For  example,  mathematical  evidence  of 


Museum  of  Comparative  Zoology  Laboratories,  Harvard  University,  Cambridge, 
Mass.  02138. 

department  of  Biology,  Boston  University,  Boston,  Mass.  02215  [To  whom  reprint 
requests  should  be  sent]. 

Manuscript  received  by  the  editor  June  19,  1981. 


265 


266 


Psyche 


[Vol.  88 


nest  overdispersion  is  frequently  confused  with,  or  taken  as  evidence 
for,  territoriality  although  crucial  behavioral  patterns  are  not 
considered.  However,  sufficient  information  is  available  in  the 
literature  to  suggest  some  of  the  behavioral  and  ecological  factors 
important  in  the  regulation  of  nest  distribution. 

With  the  above  cautions  in  mind,  we  here  present  a simple  model 
of  predicted  spatial  distributions  of  colonies  under  different  ecologi- 
cal conditions.  We  then  survey  the  literature  to  examine  the  fit  of 
available  data  to  our  predictions.  Finally  we  discuss  the  general 
problem  of  the  form  of  interactions  between  colonies  and  some  of 
the  implications  of  this  for  both  field  and  theoretical  considerations. 


Theoretical  Aspects  of  Nest  Distribution  Patterns. 

We  would  first  like  to  develop  a set  of  biologically  realistic 
predicted  spatial  distributions  of  colonies.  We  begin  by  positing 
some  simple  assumptions  about  a hypothetical  ant  population: 

1.  Nest  sites  are  unlimited. 

2.  The  habitat  is  homogenous  and  inhabited  by  a single  species. 

3.  Each  colony  forages  symmetrically  around  the  nest  to  some 
distance  r,  which  forms  the  radius  of  a circle.  Within  this  circle,  no 
other  colonies  can  forage  or  become  established. 

Simberloff  (1979)  derives  the  maximum  foraging  distance,  r,  as 

sly 

4\/3  \/p 

where  p is  the  density  of  nests.  In  this  case,  nests  are  hexagonally 
packed  and  the  array  of  nests  is  overdispersed  (more  regularly 
spaced  than  expected  if  random;  Figure  1,  case  1).  Nests  are  spaced 
2 r apart  and  have  6 equidistant  nearest  neighbors. 

Under  different  ecological  conditions,  the  expected  spatial  dis- 
tribution of  nests  will  change.  In  low  density  populations,  nest 
distribution  should  reflect  the  best  foraging  or  nest  sites;  nests  may 
be  dispersed  in  any  way  and  should  tend  towards  a random 
distribution  (Figure  1,  case  2).  Internest  distance  should  on  the 
average  be  at  least  twice  r and  usually  more;  its  variance  should  be 
high.  If  nest  sites  are  not  uniformly  available,  then  nest  spacing  will 


1981] 


Levings  & Traniello  — Territoriality  in  Ants 


267 


depend  upon  whether  or  not  nest  sites  are  farther  apart  or  closer 
together  than  this  distance.  We  predict  one  of  2 patterns:  (1)  nests 
will  be  more  overdispersed  than  potential  nest  sites  (Case  3a)  or  (2) 
although  nests  may  be  clumped  in  space,  foraging  ranges  which  are 
asymmetric  and  which  partition  foragers  will  develop  (Case  3b).  If 
potential  nest  sites  are  farther  apart  than  twice  r,  then  nests  will  be 
distributed  only  with  respect  to  potential  nest  sites.  The  effects  of 
habitat  heterogeneity  will  depend  upon  the  scale  and  extent  of  the 
patchiness  in  relation  to  the  foraging  range  of  a species.  If  patches 
hold  several  to  many  colonies,  then  clumps  of  nests  which  are 
overdispersed  within  the  clump  are  predicted.  Smaller  patches  in 
complex  mosaics  will  not  generate  predictable  nest  distributions 
unless  the  arrays  of  patches  are  very  regularly  distributed. 

The  effect  of  adding  more  species  to  the  system  will  depend  upon 
the  species.  Generally,  in  multi-species  systems,  the  level  of  repul- 
sion observed  between  co-occurring  species  should  be  directly 
proportional  to  the  amount  of  overlap  in  resource  use.  Species 
utilization  curves  can  range  in  overlap  from  0 to  essentially  com- 
plete ecological  identity  (100%  overlap).  Predicted  spatial  patterns 
will  clearly  depend  on  the  actual  distribution  of  species.  If  two  or 
more  species  with  identical  requirements  and  foraging  radii  occur  in 
the  same  area,  interactions  within  and  between  species  should  be 
equally  strong.  In  this  case,  the  pattern  of  nest  distribution  predicted 
is  random  for  any  one  species  (Franks  1980;  Levings  and  Franks 
1982).  Nests  should  be  overdispersed,  but  each  species  is  distributed 
with  respect  to  every  other  species  (i.e.,  nests  of  all  species  are 
treated  as  equivalents).  In  addition,  there  should  be  no  pattern  in 
the  species  identity  of  nearest  neighbors  (Case  4).  Removal  of  any 
one  species  should  have  the  effect  of  the  removal  of  a nest  at  random 
from  an  overdispersed  array;  the  degree  of  observed  overdispersion 
should  decrease.  The  spatial  dispersion  of  any  one  species  in  such  an 
array  should  tend  to  look  like  a low  density  nest  population  (Case 
2),  but  the  history  of  the  area  may  cause  any  type  of  pattern  under 
different  conditions. 

If  two  or  more  species  have  the  same  foraging  radius  but  do  not 
overlap  100%  in  resource  requirements,  intraspecific  interactions 
should  be  stronger  than  interspecific  interactions  (Case  5).  We 
predict  that  (1)  the  entire  array  will  be  overdispersed  and  (2)  each 
species  will  also  be  overdispersed  from  itself.  Franks  (1980)  and 


268 


Psyche 


[Vol.  88 


FIGURE  1 


Case  1 High  density  population 
Assumptions:  1.  Single  species 

2.  All  nests  have  the  same  r 

3.  Unrestricted  nest  sites 

Predictions:  1.  Overdispersed  nest  array 

2.  Nest  to  nest  distance  ~ 2 r 


Case  2 Low  density  population 
Assumptions:  1,  2,  3 

Predictions:  1.  Nest  distribution  will  tend  to  randomness 

2.  Average  nest  to  nest  distance  >2  r 

3.  High  variance  in  nest  to  nest  distance 


Case  3 Limited  nest  sites 


a.  Assumptions: 
Predictions: 


b.  Assumptions: 
Predictions: 


1,  2 

1.  Nests  more  overdispersed  than  potential  nest  sites 

2.  Nest  spacing  will  vary  with  nest  site  location,  minimum  nest 
to  nest  distance  = 2 r,  average  nest  to  nest 

distance  >2  r 

3.  High  variance  in  nest  to  nest  distance 
1 

1.  Nests  distributed  as  nest  sites 

2.  Asymmetric  foraging  ranges 


Case  4 Intraspecific  = interspecific  interactions 
Assumptions:  2,  3 

Predictions:  1.  Entire  nest  array  overdispersed 

2.  Individual  species  are  more  randomly  dispersed  than  the 
total  array 

3.  No  pattern  in  the  identity  of  nearest  neighbor 

4.  High  variance  in  nest  to  nest  distances  within  a species, 
average  nest  to  nest  distance  >2  r 

5.  Low  variance  in  nest  to  nest  distances  for  the  entire  array, 
average  nest  to  nest  distance  = 2 r 


Case  5 Intraspecific  interactions  > interspecific  interactions 
Assumptions:  2,  3 

Predictions:  1.  Entire  nest  array  overdispersed 

2.  Individual  species  within  the  array  are  also  overdispersed 

3.  Nearest  neighbors  tend  to  be  members  of  other  species 

4.  Low  variance  in  nest  to  nest  distances  within  species,  average 
nest  to  nest  distance  >2  r 

5.  Low  variance  in  internest  distances  for  the  entire  array, 
average  nest  to  nest  distance  = 2 r 


1981] 


Levings  & Traniello  — Territoriality  in  Ants 


269 


CASE  I 


CASE  2 


CASE  3a 


CASE  3b 


CASE  4 


CASE  5 


Figure  1.  Theoretical  nest  dispersion  patterns  under  different  ecological  condi- 
tions. Additional  details  in  text. 


270 


Psyche 


[Vol.  88 


Levings  and  Franks  (1982)  have  reviewed  the  relevant  statistical 
literature  and  give  a suggested  procedure  for  examining  this 
problem. 

In  addition  to  changes  in  the  observed  spatial  array  of  any  one 
species,  in  multi-species  populations,  there  should  be  correlated 
changes  in  expected  internest  distances  under  different  competitive 
regimes.  If  intra-  and  interspecific  interactions  are  equally  strong, 
the  average  internest  distance  within  any  one  species  should  be 
longer  than  twice  the  species’  average  r and  the  variance  in  between 
nest  distances  within  any  one  species  should  be  high  (essentially  a 
low  density  population,  Case  2).  If  intraspecific  interactions  are 
more  important  than  interspecific  interactions,  then  internest  dis- 
tance within  any  one  species  should  be  greater  than  twice  the 
species’  average  r and  their  variance  should  be  relatively  low.  The 
exact  predicted  distance  will  be  a function  of  the  number  of 
interacting  species  and  their  relative  abundances.  It  may  be  possible 
to  use  the  degree  of  departure  from  predicted  intraspecific  spacing 
patterns  as  a measure  of  competition  between  species  in  homog- 
enous habitats.  If  intranest  distances  within  a species  are  2 r,  then  it 
does  not  appear  to  be  interacting  significantly  with  sympatric 
species,  at  least  not  in  ways  which  affect  its  spatial  distribution. 

Detection  of  Overdispersion  and  Methodological  Problems 

There  are  certain  methodological  difficulties  in  applying  any  sort 
of  spatial  analysis  to  previously  published  data  on  nest  distribu- 
tions. In  particular,  the  complicated  structure  of  the  nests  of  many 
species  has  confused  workers,  especially  when  many  nest  entrances 
are  present.  In  Lasius  neoniger,  Headley  (1941)  assumed  that  the 
species  was  unicolonial,  since  he  could  only  occasionally  elicit 
aggression  between  adjacent  nest  entrances.  In  fact,  L.  neoniger 
colonies  are  distinct  and  well  organized,  but  extensive  field  tests  are 
required  to  delineate  colony  boundaries  (Traniello  1980).  Simple 
mapping  of  nest  openings  may  reflect  the  distribution  of  colonies 
fairly  well  (as  it  does  for  many  species  in  the  ground  ant  community 
in  Panama,  Levings  and  Franks  1982;  Levings,  personal  observa- 
tions), but  may  lead  to  confusion  unless  sufficient  data  on  the 
species  are  available  (see,  for  example,  Brough  1976).  Whitford  et 
al.  (1980)  assumed  that  workers  of  Novomessor  cockerelli  were 
entering  an  alien  nest  because  they  did  not  return  to  the  same  nest 


1981] 


Levings  & Traniello  — Territoriality  in  Ants 


271 


entrance  from  which  they  departed.  However,  Holldobler  et  al. 
(1978)  and  Davidson  (1980)  documented  that  this  species  has  nests 
with  multiple  entrances. 

Although  there  are  several  methods  for  the  detection  of  over- 
dispersion (Pielou  1977),  we  have  chosen  to  apply  Clark  and  Evans’ 
(1954)  nearest  neighbor  (NN)  technique  wherever  possible.  It  is 
based  upon  the  ratio  between  the  observed  mean  nearest  neighbor 
distance  and  the  expected  distance  when  a population  is  distribution 
at  random.  The  index  R can  range  from  0 (perfect  aggregation)  to 
2.1491  (perfect  hexagonal  overdispersion).  A value  of  1 indicates  a 
random  dispersion  pattern.  The  significance  of  R is  tested  using  the 
z transformation.  In  an  overdispersed  population,  the  observed 
mean  nearest  neighbor  distance  is  larger  and  the  variance  in  nest  to 
nest  distance  is  lower  than  it  would  be  in  a randomly  distributed 
population.  Thus  a population  which  is  significantly  overdispersed 
using  this  measure  confirms  2 of  our  predictions  (overdispersion 
and  low  variance  in  NN  distance).  Other  methods  do  not  have  this 
property. 

In  our  evaluation  of  spacing  information  in  the  literature,  if  we 
were  unable  to  apply  nearest  neighbor  methods,  but  complete 
quadrat  counts  were  published,  we  calculated  variance/  mean  ratios 
and  tested  them  for  significance  using  X2  statistics  (Pielou  1977).  A 
V/M  ratio  of  less  than  1 indicates  overdispersion  while  values 
greater  than  1 indicate  clumping.  Cases  are  included  in  which  data 
are  not  sufficient  to  test  for  statistical  overdispersion,  but  informa- 
tion on  partitioning  of  resources  or  area  was  published.  We  have 
organized  the  available  data  by  geographic  region,  habitat  and  food 
types  (Table  1).  Methods  used  in  gathering  previously  unpublished 
data  will  be  described  with  the  specific  set  of  data.  In  testing  our 
model  and  spatial  predictions  from  the  literature,  we  are  limited  by 
the  previous  interests  and  focus  of  other  authors.  We  are  able  to  test 
the  spatial  predictions  far  more  thoroughly  than  the  hypotheses 
about  the  actual  expected  distances  between  nests,  but  there  is  no 
empirical  reason  that  they  cannot  be  experimentally  verified  in  the 
field  (see  discussion). 

Data  are  discussed  by  subdividing  reported  cases  into  groups 
according  to  foraging  type:  (1)  species  which  do  not  defend  re- 
sources although  they  may  or  may  not  recruit  to  food,  (2)  species 
which  defend  randomly  and  unpredictably  distributed  resources 
(e.g.,  dead  insects,  which  are  patchy  in  both  space  and  time),  (3) 


272 


Psyche 


[Vol.  88 


species  which  defend  predictable  and  persistent  resources  (e.g., 
honeydew  from  aphids,  resources  which  are  patchy  in  space  but  not 
in  time)  and  (4)  truly  territorial  species  which  defend  area  which  has 
potential  food  resources.  These  divisions  mark  some  ecologically 
important  foraging  types  within  communites. 

Observed  Patterns 

1.  Nest  Defense 

The  data  suggest  that  species  which  display  only  nest  defense  fall 
into  4 major  groups,  depending  upon  the  details  of  their  foraging 
biology.  First,  some  species  forage  only  as  solitary  individuals  for 
food  items  which  a single  forager  can  subdue  and  retrieve  (Group  I 
foragers,  Oster  and  Wilson,  1978).  Examples  of  this  group  include 
most  Dacetini,  many  Ponerinae,  and  some  of  the  non-leaf  cutting 
Attini  (Brown  and  Wilson  1959,  Wilson  1971,  Oster  and  Wilson 
1978). 

There  is  very  little  applicable  data  on  this  group.  The  frequency  of 
dacetine  nests  in  extensive  Berlese  sampling  of  a tropical  deciduous 
forest  fit  a Poisson  distribution  indicating  a random  distribution 
(Levings,  unpublished  data),  but  this  sort  of  data  does  not  differen- 
tiate between  the  suitability  of  the  site  or  other  important  factors  in 
the  distribution  of  nests.  Certainly  there  was  no  indication  that  nests 
were  clumped.  The  maximum  number  of  nests  found  was  6 in  84 
0.25  m2  samples.  When  a truncated  Poisson  was  fit  (0  class 
excluded),  the  distribution  did  not  differ  from  Poisson  expectation 
(p  > 0.5,  x2  test). 

Second,  some  species  may  recruit  nestmates  to  food  resources, 
but  make  no  attempt  to  defend  them,  decamping  if  another,  more 
aggressive,  species  arrives  before  the  food  is  retrieved  (Group  II,  in 
part,  Oster  and  Wilson  1978).  These  species  specialize  in  the  rapid 
location  and  removal  of  food.  Examples  include  Paratrechina 
longicornis  and  Tapinoma  melanocephalum  (Wilson  1971).  No  data 
on  their  nest  distribution  is  available,  but  many  are  known  to  form 
small  fragmented  colonies  which  move  frequently  between  ephem- 
eral nest  sites. 

The  third  set  of  species  have  developed  mechanisms  for  feeding  at 
the  same  resources  as  other,  more  aggressive  ants,  without  eliciting 
defensive  reactions  (Groups  I & II,  in  part,  Oster  and  Wilson  1978, 
Wilson  1971).  It  is  not  known  how  much  of  a colony’s  food  intake 


1981] 


Levings  & Traniello  — Territoriality  in  Ants 


273 


results  from  such  theft  and  how  much  is  independently  gathered. 
Examples  include  Leptothorax  acervorum  and  various  Cardio- 
condyla  species  (Brian  1955;  Wilson  1959a,  1971).  These  species 
usually  recruit  few  other  workers  to  the  food  item;  many  of  these 
species  recruit  only  one  other  nestmate  using  tandem  running 
(Wilson  1959a).  No  spacing  information  is  available  for  these 
species. 

The  fourth  set  of  species  include  the  legionary  ants  (true  group 
foragers)  and  most  of  the  specialists  on  extremely  difficult  prey 
(Groups  IV  and  V,  Oster  and  Wilson  1978).  These  species  defend 
only  their  nest  sites  (which  may  move  often)  and  forage  in  various 
sized  groups.  The  most  spectacular  examples  of  this  type  of  foraging 
are  the  army  ants  (Schneirla  1971).  Specialists  on  difficult  prey 
occur  in  several  genera  (examples,  Pachycondyla  ( =Termitopone ), 
Leptogenys,  Gnamptogenys)\  specialized  retrieval  methods  may 
involve  extensive  cooperative  foraging  (Wilson  1971).  Little  nest 
spacing  information  is  recorded  about  these  groups.  Army  ants  of 
several  genera  have  been  observed  to  avoid  each  other  when  they 
meet  in  the  field,  but  no  similar  information  is  available  for  related 
groups  (Schneirla  1971).  Other  legionary  groups  are  relatively  rare 
on  BCI  and,  in  4 years  of  field  work,  no  interactions  were  observed 
(Levings,  personal  observation). 

In  general,  information  on  spacing  patterns  of  ants  which  defend 
only  their  nests  is  extremely  difficult  to  gather,  since  the  investigator 
must  usually  depend  upon  luck  to  locate  colonies  and  will  never  be 
certain  that  all  colonies  in  an  area  have  been  found.  Because 
information  on  foraging  ranges  for  most  species  is  unavailable,  we 
are  unable  to  test  those  aspects  of  our  hypotheses.  Many  species 
which  are  now  assumed  to  defend  only  their  nest  sites  may  well  be 
found  to  defend  either  resources  or  a foraging  territory. 

2.  Resource  defense 
a.  short  term 

The  defense  of  unpredictable  resources  occurs  on  varying  time 
scales.  Resources  which  persist  for  very  short  periods  (i.e.,  minutes 
for  most  dead  insects)  are  defended  by  many  generalist  or  scavenging 
ants  during  the  recruitment/ retrieval  process  (Groups  II  & III,  in 
part,  Oster  and  Wilson  1978).  Spatial  overdispersion  in  densely 
populated  areas  has  been  shown  in  one  complex  tropical  com- 
munity (Levings  and  Franks  1982).  It  is  probably  typical  of  many 


274 


Psyche 


[Vol.  88 


reported  cases  of  overdispersion  in  temperate  ground  ant  com- 
munities dominated  by  relatively  few  generalist  species  (most  species 
of  the  genera  Myrmica,  Tetramorium,  Lasius,  Aphaenogaster,  some 
Formica;  Table  1).  Some  species  are  placed  here  somewhat  ar- 
bitrarily because  good  foraging  ecology  data  are  not  available. 

In  more  complex  (i.e.,  non-uniform)  habitats,  the  pattern  of  nest 
spacing  is  reported  to  be  directly  related  to  environmental  condi- 
tions. Lasius  flavus,  which  has  been  intensively  studied  in  several 
European  habitats,  displays  different  nest  distributions  between 
locations.  Waloff  and  Blackith  (1962;  Table  1)  found  that  nests  were 
overdispersed  in  a high  density  population  and  tended  toward 
randomness  in  a low  density  population.  In  a wet,  low  pasture  with 
limited  nest  sites,  nests  were  also  overdispersed  (Blackith  et  al., 
1963,  Table  1).  With  Myrmica  rubra  present  in  a low  density 
population,  L.  flavus  was  randomly  distributed  (Elmes  1974). 
However,  the  partial  segregation  of  species  indicated  that  both 
intra-  and  interspecific  interactions  were  present;  M.  rubra  nests 
were  more  overdispersed  than  potential  nest  sites  (Table  1).  Similar 
patterns  have  been  noted  in  other  species.  Petal  (1972)  showed  that 
the  pattern  of  distribution  in  Myrmica  laevinodis  depended  upon 
the  scale  with  which  the  species  was  examined.  Within  the  habitat, 
nests  were  clumped,  but  within  clumps  of  nests  on  a small  scale, 
nests  were  either  overdispersed  or  randomly  distributed.  In  another 
study,  Petal  (1977)  linked  observed  nest  distribution  and  the  avail- 
able food  supply  in  Myrmica  lemanica.  In  a year  with  low  food 
abundance,  nests  were  overdispersed;  when  food  was  abundant,  nest 
distribution  was  random,  tending  to  aggregation.  Petal  did  not  state 
if  she  distinguished  between  nests  and  nest  openings  by  testing 
aggressive  responses  between  colonies.  However,  overall  nest  density 
remained  approximately  the  same.  Most  other  studies  have  assumed 
but  not  demonstrated  the  correlation  between  food  abundance  and 
nest  dispersion  patterns. 

Within  colonies  with  multiple  nest  entrances,  the  distance  be- 
tween nest  entrances  should  be  approximately  2 r and  nest  entrances 
should  be  overdispersed  if  avoiding  redundant  search  is  the  under- 
lying cause  of  polydomy.  This  appears  to  be  the  case  in  Lasius 
neoniger.  Each  nest  is  composed  of  a series  of  nest  entrances  which 
are  overdispersed  within  a colony  (Traniello  1980).  L.  neoniger  is 
unable  to  retrieve  prey  effectively  further  than  approximately  15  cm 
from  any  given  nest  opening  due  to  interference  from  other  species 


1981] 


Levings  & Tranie/lo  — Territoriality  in  Ants 


275 


or  congeners  (Traniello,  1980).  Inter-opening  distances  are  not 
statistically  different  from  30  cm  in  a set  of  12  nests  with  varying 
numbers  of  nest  openings  (P  > 0.10,  t test,  11/12  cases;  range  2-27 
nest  entrances),  fitting  our  predictions  quite  well.  The  only  nest  with 
consistently  closer  inter-opening  spacing  was  hemmed  in  by  3 larger 
nests;  its  openings  occupied  essentially  the  entire  available  area  (18 
cm  between  entrances,  4 entrances).  Although  this  species  fits  our 
predictions,  we  are  unable  to  test  them  further  with  other  species, 
either  within  species  between  nest  openings  or  between  separate 
nests.  Nest  entrance  patterns  of  Paltothyreus  tarsatus,  which  is  also 
a polydomous  species,  appear  to  be  similar  in  function  to  those  of  L. 
neoniger  (Holldobler,  personal  communication).  However,  in  poly- 
domous species  of  Camponotus,  Atta  and  Pheidole,  nest  entrances 
are  often  much  less  than  2 r apart  (Yasuno  1964a;  Holldobler  and 
Moglich  1980).  Therefore, the  association  between  foraging  ecology 
and  nest  structure  probably  depends  on  the  details  of  the  biology  of 
individual  species. 

When  resources  persist  for  slightly  longer  time  periods  (patches 
that  can  be  exploited  in  a few  days  such  as  rotting  fruit),  we  also 
expect  overdispersion  of  nests.  This  pattern  has  been  confirmed  in 
several  species.  Myrmecocystus  mimicus  nests  in  desert  areas  and 
exploits  patchy,  unpredictable  concentrations  of  termites  which 
form  a major  part  of  its  diet  (Table  1,  Holldobler  1976b,  1979a, 
Holldobler  and  Lumsden  1980).  During  the  retrieval  of  these 
patches  of  food,  a nest  will  defend  the  area  by  engaging  surrounding 
nests  in  a complex  ritualized  display  and  battle  (“tournamenting”) 
which  may  result  in  the  destruction  of  incipient  colonies.  Normally, 
the  tournamenting  behavior  persists  until  the  patch  is  exploited; 
searching  in  the  area  continues  during  this  time.  Nests  are  overdis- 
persed (G.  Alpert,  personal  communication).  Nests  of  Prenolepsis 
imparis  are  overdispersed  (Table  1),  and  workers  defend  pieces  of 
fruit  for  1 or  more  days.  This  species  has  also  been  observed  to 
tournament  as  Myrmecocystus  mimicus  does  (Traniello,  unpub- 
lished observations).  It  appears  that  in  these  species  the  cost  of 
allocating  a portion  of  the  worker  force  to  engage  foragers  from  a 
neighboring  nest  in  tournaments  that  prevent  their  access  to  a 
resource  is  less  than  the  benefits  obtained  from  these  patchily 
distributed  food  sources  (Holldobler  and  Lumsden  1980). 
b.  persistent  resources 

Persistent  resources  vary  in  their  importance  to  colonies,  depend- 


276 


Psyche 


[Vol.  88 


ing  on  their  nutritional  value,  and  can  differ  from  a branch  with  a 
few  aphids  to  a large  homopteran  population  which  provides  most 
of  a colony’s  food  intake.  The  degree  to  which  a colony  depends 
upon  persistent  resources  will  approximately  determine  the  intensity 
of  their  defense.  Formica  fusca  tends  only  a very  few  aphids  and  can 
be  chased  from  them  relatively  easily  (Brian  1955),  while  F.  rufa 
colonies  regularly  destroy  each  other  in  battles  for  the  control  of 
specific  trees  (Elton  1932,  Skinner  1980).  Therefore,  the  removal  of 
persistent  resources  can  affect  colonies  differently;  some  nests  will 
die  if  they  are  deprived  of  them  (Elton  1932).  In  this  section  we  will 
only  consider  species  which  are  dependent,  at  least  in  part,  upon 
such  resources  (Groups  II  & III,  in  part,  Oster  and  Wilson  1978). 

Most  studies  on  the  defense  of  persistent  resources  concern  the 
genera  Formica  and  Pogonomyrmex.  The  patterns  of  their  nest 
distribution  depends  upon  colony  structure,  nest  site  requirements 
and  habitat  complexity.  Most  Formica  nest  distributions  are  the 
result  of  the  interaction  between  the  need  for  high  insolation  of  the 
nest  and  the  proximity  of  trees  or  bushes  which  are  suitable  for 
tending  aphids.  Many  species  nest  along  the  ecotones  between  fields 
and  forests,  in  forests,  and  in  forest  clearings  (F.  lugubris,  F. 
schaufussi,  F.  exsectoides,  F.  polyctena,  F.  rufa,  F.  ulkei).  These 
species  will  be  found  in  overdispersed  arrays  only  if  habitat  patches 
are  found  in  rather  predictable  patterns.  These  are  clearly  special 
cases  and  explain  some  of  the  variation  between  authors  for  some 
species  (see  for  example,  F.  lugubris,  Table  1).  We  expect  the  linear 
distance  along  the  ecotone  to  be  relatively  even  in  this  case,  but  we 
have  no  data  to  test  this  hypothesis.  Casual  observations  on  F. 
schaufussi  tend  to  support  this  (Traniello,  personal  observations). 
Formica  species  which  nest  in  fields  should  be  found  in  overdis- 
persed arrays;  the  few  reports  that  exist  indicate  that  they  are  (F. 
uralensis,  F.  opaciventris,  F.  fusca,  F.  pratensis,  Table  1).  In 
addition,  Pogonomyrmex  species  which  defend  patches  of  seeds  are 
found  in  overdispersed  arrays.  These  and  other  species  that  defend 
persistent  resources  and  are  not  nest  site  limited  are  in  general  found 
in  overdispersed  arrays  ( Atta  spp.,  Acromyrmex  octospinosus, 
Lasius  niger,  etc.,  Table  1). 

Colonies  which  depend  upon  persistent  resources  frequently 
organize  resource  defense  and  utilization  with  trunk  trails.  Trunk 
trails  are  long  term  routes  which  are  marked  with  persistent  trail 


1981] 


Levings  & Traniello  — Territoriality  in  Ants 


277 


pheromones  (Holldobler,  1974,  1976a;  Traniello,  1980;  Group  III, 
Oster  and  Wilson,  1978).  Thus  both  the  track  to  the  resource  and 
the  resource  itself  constitute  the  defended  area.  These  foraging 
ranges  are  highly  asymmetric — foragers  from  different  colonies  are 
only  likely  to  interact  when  trail  systems  overlap.  Essentially  all 
foragers  follow  these  trails  in  some  species  (Holldobler,  1976a),  but 
this  varies  a great  deal  from  group  to  group.  In  general,  we  expect 
that  nest  to  nest  distances  will  be  shorter  than  the  distance  to  the 
defended  resource  if  colonies  have  highly  skewed  foraging.  This 
prediction  is  born  out  in  a study  of  three  sympatric  species  of 
Pogonomyrmex  (Holldobler,  1976a).  Between  nest  distances  are 
shorter  in  the  two  interspecifically  defending  species  which  forage 
on  trunk  trails  than  between  nests  of  the  individually  foraging  P. 
maricopa. 

3.  Defense  of  area 

We  consider  defense  of  space  larger  in  area  than  nest  yards  or 
core  areas  (Holldobler  1976a)  to  be  true  territoriality.  This  defense 
of  area  is,  in  essence,  defense  of  potential  foraging  grounds.  Only  a 
few  ant  species,  characterized  by  complex  mechanisms  of  mass 
recruitment  and  high  levels  of  intra-  and  interspecific  aggression,  are 
therefore  truly  territorial  in  our  classification.  Most  dominant 
tropical  canopy  ants  (some  members  of  the  genera  Azteca,  Oeco- 
phylla,  Crematogaster,  Camponotus,  Monads,  Polyrachis,  Anoplo- 
lepis,  Table  1)  and  at  least  one  member  of  the  genus  Solenopsis  are 
truly  territorial.  We  must  point  out  that  in  some  cases  the  distinc- 
tion between  true  territoriality  and  resource  defense  is  not  perfectly 
clear,  and  that  strategies  of  territorial  defense  and  resource  defense 
are  at  times  difficult  to  distinguish. 

Solenopsis  invicta,  an  introduced  species  from  South  America, 
has  been  extensively  studied  in  the  southern  United  States  where  it 
may  form  monocultures  in  fields  (Wilson  et  al.  1972).  Extensive 
mapping  of  one  population  showed  overdispersion  of  nests  main- 
tained over  time  despite  frequent  nest  movement  (Eisenberg  1972, 
Table  1). 

Maps  of  intercolony  dispersion  have  been  published  for  a number 
of  ant  species  in  tree  crops  in  tropical  areas  (Table  1).  Individual 
colonies  hold  territories  in  the  canopy  both  intra-  and  interspecifi- 
cally. The  distribution  of  the  canopy  mosaic  of  dominants  can  have 
a very  complex  structure  (Way  1953;  Greenslade  1971;  Majer 


278 


Psyche 


[Vol.  88 


1976a,b).  Individual  colonies  are  clearly  separated  from  each  other, 
frequently  by  a no  ant’s  land  between  defended  areas  (Holldobler 
1979b). 

However,  the  statistical  dispersion  of  these  colonies  is  difficult  to 
assess.  Territory  size  varies  a great  deal  between  species  because 
population  structure  has  very  strong  effects  on  colony  size  and 
organization.  Only  a few  polydomous,  polygynous  colonies  may 
occupy  extensive  areas  (Steyn  1954,  Greenslade  1971,  Leston  1973, 
Majer  1976a, b).  The  dispersion  of  volumes  in  space  is  difficult  to 
treat  statistically  from  published  data  although  3-dimensional  meth- 
ods exist  (Clark  and  Evans  1979,  Simberloff  1979).  Luckily,  the 
biological  evidence  for  dispersion  and  nonoverlap  of  area  is  over- 
whelming. Territorial  battles  are  commonly  observed  and,  in  popu- 
lations followed  over  a number  of  years,  control  of  a given  area 
shifts  from  colony  to  colony  and  species  to  species.  Although  we 
predict  statistical  overdispersion,  we  are  unable,  for  both  statistical 
and  biological  reasons,  to  test  for  it  in  these  cases.  There  is, 
however,  no  question  about  the  existence  of  true  territorial  defense 
and  the  spatial  separation  of  colonies. 

Ant  plants  are  a special  set  of  cases  of  true  territoriality.  Several 
tropical  tree  species  (Table  1)  are  coevolved  with  certain  species  of 
ants  (some  members  of  the  genera  Pseudomyrmex,  Azteca  or 
Pachysima ) which  protect  the  tree  from  herbivores  or  overgrowth  in 
return  for  food  and  nest  sites.  Few  other  animals  of  any  species  are 
tolerated  on  the  plant;  the  ant  species  are  characteristically  extreme- 
ly aggressive.  The  mutualism  is  sufficiently  old  than  at  least  one 
species  parasitizes  it  by  using  the  plant  without  protecting  it  in 
return  (Janzen  1975).  These  ant  colonies  are  thus  distributed  with 
respect  to  the  distribution  of  their  host  and  form  distinct  territories 
within  the  canopy  mosaic.  They  may  also  help  “grow”  new  host 
plants  by  affecting  the  survival  of  nearby  seedlings  (Janzen  1967, 
1973). 

Intercolony  Spacing  Effects 

Interspecific  overdispersion  is  regularly  reported  in  almost  all 
habitats  (Table  1).  However,  detailed  ecological  studies  indicate  that 
different  mechanisms  operate  in  different  habitats.  In  part,  this  is 
due  to  the  fact  that  the  only  necessary  characteristic  required  to 
generate  overdispersed  arrays  is  the  ability  of  a colony  to  interfere 


1981] 


Levings  & Traniello  — Territoriality  in  Ants 


279 


with  colony  foundation  of  potential  competitors.  Few  studies  have 
examined  the  pattern  of  species  mingling.  Brian  and  his  co-workers 
have  shown  that  nest  sites  probably  limit  many  species  in  England, 
which  apparently  is  a rather  poor  habitat  for  ants.  The  pattern  of 
dominance  over  nest  sites  determines  the  location  and  abundance  of 
many  species  (Brian  1952,  1956b,  Brian  et  ah  1965).  Nest  density 
could  be  increased  by  providing  new  nest  sites,  and  once  established, 
populations  remained  relatively  stable  over  long  periods  (Brian  et 
al.  1966).  Competition  between  species  where  nest  sites  were  not  as 
limiting  tended  to  restrict  individual  species  to  areas  which  were 
close  to  optimal  species  requirements  (Brian  et  al.  1966,  Elmes  1971, 
1974).  These  studies  indicate  that  the  details  of  species  biology  and 
physical  tolerances  may  be  critical,  even  in  very  simple  habitats  like 
heath.  However,  even  in  these  systems,  species  are  definitely  not 
distributed  independently  of  their  competitors. 

Tropical  canopy  dominants  are  associated  with  certain  canopy 
conditions,  and  tend  to  be  found  mostly  in  shade  or  under  certain 
other  limited  environmental  states  (Majer  1976a).  Removal  experi- 
ments indicate  that  colony  foraging  areas  are  competitively  com- 
pressed; when  a dominant  is  removed,  surrounding  colonies  expand 
to  fill  the  available  space.  Species  usually  found  in  one  kind  of 
canopy  may  expand  into  other  types  of  foliage  if  adjacent  domi- 
nants are  extirpated  (Majer  1976a,b).  This  pattern  is  similar  to  that 
found  in  far  simpler  grassland  communities. 

In  a complex  tropical  ground  ant  community  with  at  least  16 
ecologically  similar  species,  Levings  and  Franks  (1982)  have  shown 
that  new  nests  are  not  added  at  random  to  the  nest  array.  Grouping 
all  species,  nests  are  overdispersed  from  each  other.  Each  common 
species  considered  independently  was  also  overdispersed.  This  is 
interpreted  as  evidence  that  species  are  interacting  more  strongly 
intra-  than  interspecifically,  but  that  interspecific  effects  were  still 
important  in  determining  nest  distributions.  Similar  patterns  in 
simpler  communities  indicate  that  this  may  be  common  (Table  1). 
The  worst  neighbor  in  a competitive  sense  should  be  identical  to 
oneself.  In  any  case,  the  simplifying  assumption  that  species  have 
identical  requirements  is  almost  infinitely  unlikely  to  apply;  even 
small  differences  in  requirements  or  tolerances  can  be  important  in 
determining  colony  distributions.  However,  few  adequate  tests  have 
been  done  and,  in  one  published  case,  two  closely  related  congeners, 


280 


Psyche 


[Vol.  88 


Pogonomyrmex  rugosus  and  P.  barbatus,  essentially  act  like  exact 
ecological  equivalents  (Holldobler  1976a).  Davidson  (1977a)  has 
suggested  that  the  distribution  of  several  individually  foraging 
Pogonomyrmex  (maricopa,  californicus,  desertorum  and  magni- 
canthus)  is  consistent  with  the  hypothesis  that  they  replace  each 
other  between  habitats.  The  pattern  could  also  occur  in  some  other, 
less  completely  documented  cases,  perhaps  in  Atta  (Rockwood 
1973). 

Population  Structure  and  Its  Effects  on  the  Spatial 
Distribution  of  Colonies 

Monogynous,  queenright  colonies  are  almost  innevitably  aggres- 
sive to  conspecific  nests  or  foragers,  regardless  of  how  territorial 
they  are  (Holldobler  and  Wilson  1977c).  Polygynous  colonies  may 
or  may  not  display  internest  aggression.  Holldobler  and  Wilson 
(1977c)  point  out  the  importance  of  queen  number  in  the  mainte- 
nance of  clear  territorial  borders.  Species  which  commonly  have 
polygynous  colonies  and/or  those  which  adopt  newly  fertilized 
females  to  augment  or  replace  females  already  in  the  nest  do  not 
always  have  strong  intraspecific  interactions;  some  do  not  form 
distinct  colonies  ( Formica  yessensis,  F.  lugubris,  Table  1).  In  these 
cases  the  location  of  nests  should  be  predominantly  determined  by 
ecological  factors,  in  particular  the  kind  of  resource  defense  the 
colony  shows.  Thus  some  species  should  retain  overdispersed 
patterns  of  nest  distribution  while  other  show  clumped  or  random 
patterns  (see  model  and  predictions). 

Examining  this  issue  is  complicated  by  the  lack  of  population 
structure  data  for  many  species.  Several  Formica  species  which 
form  unicolonial  populations,  but  depend  upon  randomly  and 
unpredictably  distributed  resources,  are  found  in  overdispersed 
arrays  [those  species  found  in  fields:  F.  pratensis  (provisionally),  F. 
uralensis  (provisionally),  F.  opaciventris,  F.  exsectoides,  Table  1]; 
those  which  nest  along  the  margins  of  a habitat  and/or  which 
defend  persistent  resources  tend  to  have  more  random  or  clumped 
distributions  ( F . ulkei,  F.  rufa,  F.  lugubris,  Table  1).  Territory  size 
in  some  tropical  tree  ants  is  partially  a result  of  population 
structure.  Many  dominant  species  are  polygynous  and  are  able  to 
expand  their  territories  almost  indefinitely  under  good  ecological 
conditions  (Greenslade  1971,  Majer  1976a,b).  In  some  cases,  single 


1981] 


Levings  & Traniello  — Territoriality  in  Ants 


281 


queen  species  like  Oecophylla  may  be  at  a disadvantage.  Resistance 
to  invasion  or  persistance  of  the  nest  may  be  limited  by  the  female’s 
egg  production  under  some  conditions,  although  this  does  not 
usually  seem  to  be  the  case  (Holldobler  and  Wilson  1977c;  Holl- 
dobler  and  Lumsden  1980).  We  must  emphasize  that  in  populations 
with  complex  or  variable  structure  it  may  be  very  difficult  to 
determine  the  factors  which  are  controlling  distributions.  Spacing 
may  reflect  foraging  ecology  as  well  as  being  an  aspect  of  territorial- 
ity. More  data  are  needed  before  good  generalizations  can  be  made. 

Behavioral  and  Ecological  Aspects  of  Spacing 

For  the  cases  we  have  been  able  to  examine  statistically,  67  out  of 
80  show  overdispersed  nest  distributions  or  tend  toward  overdisper- 
sed nest  distributions.  The  other  80  cases,  which  cannot  be  treated 
statistically,  mainly  have  either  overdispersed  nest  distributions  or 
tend  toward  overdispersed  nest  distributions.  Thus  the  majority  of 
species  studied  tend  to  have  regular  nest  arrays.  This  pattern  holds 
despite  the  large  number  of  species,  food  types  and  habitats 
considered.  Species  which  defend  only  their  nests  are  too  rare  to 
consider  in  our  sample. 

Our  basic  assumption  is  that  no  colony  can  become  established  or 
forage  within  some  radius  r of  another  colony.  There  is  a biological 
basis  for  this  assumption  in  the  patterns  of  interference  with  colony 
establishment  and  foraging  patterns.  Therefore,  to  understand  nest 
spacing  it  is  important  to  understand  the  different  levels  of  competi- 
tion in  ant  communities.  Fertilized  females  or  incipient  colonies  are 
usually  destroyed  when  they  are  encountered  by  foragers  from 
established  colonies  (Wilson  1971).  The  specificity  of  this  behavior 
varies  between  species  depending  in  part  on  population  structure 
(Holldobler  and  Wilson  1977c,  DeVroey  1979).  There  is  some 
evidence  that  workers  are  more  likely  to  attack  females  from 
conspecific  nests  or  closely  related  species,  especially  in  monogy- 
nous,  queenright  colonies,  as  has  been  shown  in  Pogonomyrmex 
(Holldobler  1976a)  and  Myrmecocystus  (Holldobler,  personal  com- 
munication). The  studies  of  Pontin  (1960)  and  others  (reviewed  in 
Wilson  1971)  suggest  that  such  behavior  is  more  often  directed 
toward  queens  of  the  same  species  as  the  attacking  workers. 

Another  factor  which  may  operate  during  this  period  is  resource 
depletion  mediated  by  either  direct  interference  or  exploitation. 


282 


Psyche 


[Vol.  88 


Within  the  foraging  radius  of  an  established  colony,  there  is  likely  to 
be  less  food  available,  even  if  the  established  colony  ignores 
incipient  colonies.  The  amount  of  depletion  will  depend  on  the 
amount  of  resource  overlap.  Because  destruction  of  females  and 
incipient  colonies  is  frequently  reported  and  resource  depletion 
probably  also  affects  colony  persistence,  the  chance  of  a small 
colony  becoming  established  is  low.  Wilson  (1971)  estimates  that 
only  0.01%  of  all  fertilized  females  survive  to  found  successful  nests. 
Therefore,  established  colonies  tend  to  persist  and  interact  over  long 
periods,  insofar  as  is  known  (Wilson  1971).  Given  this  pattern,  what 
is  the  form  of  the  interaction  and  why  are  patterns  of  interspecific 
overdispersion  so  common? 

According  to  current  theory,  species  can  segregate  a habitat  to 
avoid  or  lessen  competition  in  several  ways:  microhabit  partition- 
ing, food  size  or  type,  and  activity  period  (Pianka  1978).  Further, 
equilibrium  theory  generally  asserts  that  only  a limited  amount  of 
overlap  is  tolerated  on  any  given  niche  axis  (Mac Arthur  and  Levins 
1967;  Colwell  and  Futuyma  1971).  Species  which  are  too  similar 
should  not  be  able  to  coexist  and,  over  a long  enough  period,  the 
superior  competitor  in  the  overlapping  pair  will  drive  the  other 
species  extinct.  Although  there  are  many  problems  with  the  assump- 
tions of  this  argument,  we  will  use  its  basic  divisions  to  examine  the 
patterns  of  overlap  between  co-occurring  ant  species.  Ant  species 
may  be  specialized  along  these  three  major  axes.  We  will  consider 
each  potential  kind  of  specialization  in  turn  and  evaluate  the 
evidence  that  segregation  of  species  along  that  factor  is  usually 
sufficient  to  prevent  strong  competitive  interactions. 

Species  may  be  considered  specialized  on  food  types  in  3 major 
ways:  (1)  restricted  prey  types  (i.e.,  only  centipedes),  (2)  specific  size 
ranges  of  prey  (i.e.,  only  prey  1-3  mm  in  length)  or  (3)  some 
combination  of  (1)  and  (2)  (i.e.,  centipedes  between  5 and  8 mm 
long).  Different  kinds  of  specializations  will  have  different  effects  on 
colony  structure,  nest  size  and  foraging  strategy.  Resource  restric- 
tion is  frequently  based  on  the  matching  of  mandible  or  head  size  to 
food  size  or  type  (the  trophic  appendage,  Schoener  1971,  see  below). 
Resources  which  are  retrieved  by  individual  workers,  not  by 
coordinated  action,  are  especially  likely  to  be  treated  in  this  manner 
(for  example,  seeds  for  desert  ants,  collembolans  for  dacetines).  The 
resistance  of  the  food  item  to  recovery  is  also  important;  items 


1981] 


Levings  & Traniello  — Territoriality  in  Ants 


283 


which  do  not  resist  (seeds)  are  more  likely  to  be  size  matched  than 
items  which  require  more  complex  treatment  from  the  ants.  Nests  of 
specialist  species  may  be  restricted  to  areas  which  contain  concen- 
tration of  suitable  prey  (and  as  such  violate  the  assumptions  of  our 
model).  If  resource  size  is  matched  with  worker  size  class,  then  size 
polymorphism  is  one  way  to  expand  the  resource  spectrum  of  the 
colony  without  any  changes  in  individual  retrieval  patterns  (Oster 
and  Wilson  1978).  The  development  of  coordinated  retrieval  mech- 
anisms can  further  expand  the  accessible  resource  spectrum. 

Almost  all  specialists,  by  definition,  have  less  harvestable  energy 
available  to  them  than  generalists.  Thorne  and  Sebens  (1981) 
suggest  that  species  with  low  habitat  quality  (i.e.,  low  food  density) 
will  have  smaller  nests  than  species  with  high  quality  areas  (high 
food  density).  We  extend  this  argument  to  predict  that  once  a 
species  has  broadened  its  diet,  it  will  include  essentially  all  retriev- 
able food  types  encountered.  Such  an  increase  in  diet  breadth  is 
needed  to  support  large  colony  sizes,  based  on  almost  any  simple 
foraging  efficiency  model.  Although  specific  prey  types,  especially 
those  with  noxious  chemical  defenses,  require  special  handling 
methods,  many  prey  types  may  be  captured  and/or  retrieved  by 
species  with  a limited  behavioral  repertoire.  Certainly  scavenged 
material  can  be  handled  by  all  but  the  most  specialized  mandibular 
types.  Since  ant  colonies  persist  over  years,  they  more  or  less 
continually  require  resources.  Resource  distributions  are  highly 
variable  over  time;  prey  types  appear  and  disappear  seasonally 
(Mabelis  1979;  Levings  and  Windsor  1982).  It  is  a general  con- 
sequence of  this  that  once  a species  generalizes  its  diet,  it  is  likely  to 
overlap  strongly  with  one  or  more  sympatric  species.  The  value  of 
large  colony  size  is  reflected  in  reproductive  output.  Numbers  of 
reproductives  usually  increase  with  colony  size  to  some  upper  limit 
(Wilson  1971).  Since  the  chance  of  success  for  any  given  reproduc- 
tive is  low,  high  production  will  be  likely  to  correlate  with  the  largest 
probability  of  leaving  successful  offspring.  Colonies  which  bud  will 
tend  to  have  higher  rates  of  success  if  the  new  buds  have  large 
worker  forces;  this  is  also  a function  of  energy  intake.  Colonies 
almost  all  require  protein  to  raise  brood  (usually  from  insect  prey  or 
seeds)  and  many  accept  or  require  sugar  to  maintain  adult  workers 
(usually  from  Homoptera,  fruit  or  nectaries).  In  general,  large 
colony  size  is  strongly  associated  with  the  maintenance  of  sugar 


284 


Psyche 


[Vol.  88 


resources.  It  appears  that  when  adults  can  be  fueled  from  sugar, 
more  intensive  foraging  is  possible  and  more  brood  and  workers  can 
be  supported  (Greenslade  1971;  Leston  1973). 

We  do  not  deny  that  species  which  are  specialized  on  prey  types 
evolved  resource  segregation  from  competitive  pressure.  In  fact, 
among  specialized  species  which  forage  individually  for  prey,  we 
expect  some  equilibrium  co-existence  theory  to  apply  (see  for 
example,  Davidson  1980).  We  assert  that  there  is  no  support  for  the 
contention  that  generalists  segregate  the  resource  spectrum  to 
reduce  competition  (Wilson  1971).  Available  empirical  studies 
indicate  that  high  or  essentially  complete  overlap  in  food  type  is 
frequent  (Brian  1956a, b;  Pontin  1961,  1963,  1969;  Yasuno  1964a, b); 
Abe  1971;  Holldobler  1976a;  Levieux  1977,  Levings  and  Franks 
1982).  At  best,  partitioning  of  food  type  can  account  for  only  a 
small  part  of  the  observed  pattern  of  species  distribution  in  most 
habitats. 

Habitat  partitioning  is  a second  possible  method  of  limiting 
competitive  interactions.  Even  in  simple  temperate  grassland  com- 
munities, co-occurring  species  forage  at  slightly  different  heights  in 
the  grass  or  tend  to  move  more  or  less  beneath  the  surface  (Brian 
1952,  1955,  1956b;  Brian  et  al.  1966).  However,  all  these  species  are 
usually  described  as  being  interspeciflcally  territorial.  Tropical 
faunas  are  well  divided  into  arboreal  and  terrestrial  components; 
many  specialized  species  are  further  restricted  to  logs,  rotting  leaves 
or  other  microhabitats  (Wilson  1959b,  1971;  Carroll  and  Janzen 
1973).  Within  these  strata,  high  overlap  between  species  resulting  in 
intra-  and  interspecific  aggression  is  frequently  described  (Carroll 
and  Janzen  1973;  Leston  1973;  Greenslade  1975;  Room  1975a, b). 

Faunas  may  be  further  subdivided  by  time  of  foraging,  if  by 
foraging  at  different  times,  different  resources  are  harvested.  Time 
of  foraging  can  differ  daily  (nocturnal  vs.  diurnal,  Carroll  and 
Janzen  1973),  seasonally  ( Prenolepis  imparts  which  forages  in  early 
spring  and  late  fall,  Talbot  1943,  Lynch  et  al .,  1980)  or  may  track 
environmental  cues,  such  as  desert  species  that  forage  after  rains 
(Bernstein  1974,  1979).  Although  it  has  not  been  proven,  it  is 
probable  that  generalist  and  scavenging  species  forage  on  different 
resources  if  they  forage  at  different  times,  if  there  are  temporal 
components  to  food  availability.  Most  dead  or  readily  captured 
prey  do  not  remain  available  for  long  periods,  few  probably  persist 
even  hours  (Carroll  and  Janzen  1973,  Culver  1974,  Traniello  1980). 


1981] 


Levings  & Traniello  — Territoriality  in  Ants 


285 


Other  resources  may  be  similarly  affected — for  example,  winds  may 
cover  and  uncover  seeds  in  the  desert  (Reichman  1979).  The  option 
to  forage  at  different  times  is  not  uniformly  available  to  ant  species; 
thermal  tolerances  may  severely  limit  foraging  time  in  both  cold  and 
hot  climates  or  may  affect  the  outcome  of  foraging  contests  (Hunt 
1974;  Davidson  1977a, b,;  Holldobler  and  Moglich  1980).  Many 
species  change  the  time  of  their  foraging  in  the  presence  of 
competitors  (Hunt  1974,  Swain  1977).  Thus  foraging  times  may 
separate  some  species,  but  as  in  the  case  of  food  or  habitat,  high 
overlap  between  sets  of  sympatric  species  in  foraging  time  is  the 
norm,  not  the  exception,  in  ant  communities.  The  evolution  of 
intra-  and  interspecific  behaviors  incuding  complex  patterns  of  food 
retrieval  and  defensive  strategies  has  resulted  from  such  high 
overlap. 

The  form  and  outcome  of  interactions  between  species  is  de- 
termined in  large  part  by  the  mechanisms  of  recruitment  and 
communication  within  species.  The  subtleties  of  recruitment  com- 
munication and  their  effects  on  foraging  ecology  and  interference 
competition  are  not  appreciated  by  most  ecologists.  Behavioral 
mechanisms  are  so  critical  that  we  suggest  that  when  examining  the 
diet  of  a species,  an  investigator  first  ask  why  more  items  are  not 
included.  For  many  years  harvester  ants  were  considered  to  forage 
individually  for  seeds,  but  the  field  and  laboratory  studies  of 
Holldobler  (1976a),  Holldobler  and  Wilson  (1970)  and  Holldobler 
et  al.  (1978)  unequivocally  demonstrated  that  species  of  Pogon- 
omyrmex  and  Novomessor  rely  on  a sophisticated  array  of  recruit- 
ment behaviors  in  foraging.  In  Novmessor  cockerelli,  short-range 
recruitment,  mediated  by  both  chemical  and  vibrational  signals, 
allows  workers  to  move  food  sources  quickly  and  thereby  enables 
them  to  compete  with  sympatric  species  (Holldobler  et  al.  1978; 
Markl  and  Holldobler  1978). 

Behavioral  interactions,  not  food  choice,  seem  to  partition  food 
resources  among  generalists.  Protein  foods  (arthropod  prey)  tend  to 
be  highly  unpredictable  in  time,  space  and  size;  adaptations  to  this 
resource  distribution  among  generalists  may  be  behavioral  rather 
than  morphological.  Monomorium  pharonis  and  Solenopsis  fugax 
employ  a chemical  interference  technique  both  defensively  and 
offensively  during  foraging  (Holldobler  1973).  Adams  and  Traniello 
(1981)  have  documented  the  ecological  effects  of  recruitment  and 
resource  defense  in  Monomorium  minimum,  a north  temperate 


286 


Psyche 


[Vol.  88 


open  field  ant.  Monomorium  minimum  is  a small  (head  width  0.47 
mm),  monomorphic  species.  Workers  are  successful  at  retrieving 
food  particles  which  are  either  extremely  small  (less  than  0.5  mg  in 
weight)  or  large  (greater  than  450  mg  in  weight).  Most  items  of 
intermediate  size  are  lost  due  to  either  exploitative  or  interference 
competition  from  other  species.  Detailed  laboratory  and  field 
experiments  on  the  organization  of  foraging  showed  that  M. 
minimum  recruits  other  workers  to  food  sources  using  trail  pher- 
omones. The  quantity  of  pheromone  determines  the  foraging  re- 
sponse of  the  colony.  As  trail  pheromone  concentration  increases, 
more  workers  are  recruited.  The  amount  of  trail  pheromone 
deposited  is  dependent  upon  resource  quality  (in  this  case,  measured 
by  the  investigators  as  weight).  Large  food  items  induce  trail  laying 
by  many  workers  and  therefore  result  in  strong  recruitment.  If  there 
is  interference  from  another  species  while  prey  is  being  dissected, 
workers  display  a specific  posture  (gaster  flagging)  while  extruding 
the  sting  and  discharging  a droplet  of  poison  gland  secretion.  This 
secretion  has  a repellent  effect  on  intruding  ants  and  causes  them  to 
recoil  and  vigorously  groom.  The  effectiveness  of  this  defensive 
behavior  is  dependent  on  the  number  of  workers  recruited.  There- 
fore, large  prey,  which  elicit  strong  trail  pheromone  deposition, 
induce  strong  recruitment  responses  and  this  results  in  a worker 
force  which  can  successfully  defend  the  item  during  retrieval.  The 
result  of  this  feedback  between  prey  size,  pheromone  concentration 
and  colony  response  is  a diet  composed  of  small  individually 
retrieved  items  and  large  items  recovered  by  recruitment  and 
successful  defense. 

Perhaps  the  best  evidence  for  the  importance  of  behavioral 
parameters  in  species  interactions  is  the  phenomenon  of  alarm 
specification.  Certain  ant  species  which  interact  strongly  with  other 
species  may  respond  specifically  to  the  presence  of  the  competitor. 
The  best  studied  case  is  that  of  Pheidole  dentata  and  Solenopsis 
geminata  (Wilson  1975).  Pheidole  dentata  colonies  respond  to  the 
presence  of  Solenopsis  by  a strong  recruitment  of  major  workers. 
Major  workers  proceed  to  attack  and  kill  all  Solenopsis  encountered 
and  to  search  thoroughly  the  area  near  where  the  Solenopsis 
workers  were  found.  They  do  not  respond  to  the  odors  or  presence 
of  other  species  with  major  worker  recruitment.  A similar  pattern  of 
response  is  indicated  in  the  interactions  between  Oecophylla  longi- 


1981] 


Levings  & Traniello  — Territoriality  in  Ants 


287 


noda  and  Camponotus  sp.  in  Kenyan  forest;  alarm/ recruitment 
specification  may  be  the  behavioral  mechanism  responsible  for  the 
structure  and  maintenance  of  the  tropical  canopy  ant  mosaic  (Holl- 
dobler  1979b). 

In  general,  to  defend  any  resource  or  area,  including  the  nest,  an 
ant  must  be  able  to  summon  her  nestmates  to  a particular  location. 
Within  the  nest,  even  quite  primitive  ants  are  able  to  communicate 
alarm  and  attract  reinforcements  (Robertson  1971;  Traniello  un- 
published data).  Outside  the  nest,  recruitment  is  a necessary  com- 
ponent of  effective  defense. 

Ant  species  possess  a wide  diversity  of  recruitment  communica- 
tion techniques  that  are  ecologically  significant  (see  review  by 
Holldobler  1977).  It  is  important  to  understand  the  ethology  of 
social  design  to  comprehend  its  role  in  ecological  interaction.  There 
are  definite  phylogenetic  constraints  and/or  trends  in  the  form  of 
recruitment  communication  within  the  various  subfamilies  of  ants 
(Holldobler  1977).  More  primitive  groups  (some  Ponerinae)  usually 
recruit  few  workers  to  food  sources;  some  group  raiding  species  are 
exceptions.  Mass  recruitment  is  characteristic  of  some  groups  of 
Myrmicinae,  Dolichoderinae  and  Formicinae.  Each  lineage  has 
developed  within  certain  paths  involving  specific  glandular,  physical 
and  behavioral  trends.  These  pathways  are  important  in  considering 
the  evolution  and  development  of  ant  community  structure. 

Summary  and  Conclusions 

We  have  argued  that  a very  simple  hypothesis  is  sufficient  to 
generate  predictions  of  spatial  distributions  of  colonies  under  a 
variety  of  ecological  settings.  The  majority  of  cases  in  the  literature 
(Table  1)  support  the  hypothesis  that  most  ant  species  are  regularly 
distributed  with  respect  to  conspecifics  and  other  co-occurring 
species.  We  assert  that  this  is  a natural  outcome  of  high  overlap  in 
food  utilization  in  many  species,  and  in  particular,  among  general- 
ists. We  have  suggested  that  departures  from  expected  spatial 
patterns  be  used  as  a measure  of  competition  between  species,  but 
too  little  information  on  colony  foraging  radii  in  relation  to  spacing 
patterns  exists  to  test  our  hypotheses  critically.  More  field  measure- 
ments of  colony  foraging  distances  in  relation  to  intercolony  spacing 
are  needed.  Measurements  of  potential  foraging  distances  when 


Nest  Defense  Forager  Nest 

Species  site1  type2  type3  spacing4  Evidence5  Source 


288 


Psyche 


[Vol.  88 


60 

CA 

T3 


<N 

s ? g 

w SO 
- rX  Crt 

C X}  a > 

.2  25  g 

CQ  W 


On 

<N  VO 

u->  ^ On 
ON  — 


^ NO 

C JQ 
.2  2 
'C 
CQ 


.£  o 
(6 

60  3 
cd  X 


O e 

S 4>  « 


£ o 


3 

o 

"O 

<D 

cd 

a 


O • - 

E 1 
♦2  o 
8 ^ 
z 


S?  g ts 

cd  X • — 

^ -C  *« 


<U  X 
> Cd 
O X 

c *2 

S *S 


2 ‘o 

1 ^ ^ 

T3  <u 

.s  § * 

60 

T3  t« 

<u  .£  C 

& £ S> 

& o « 

,§!3 

« 2 « 

j?  <u 

O H 


<L> 

o. 

x" 

x" 

X 

JD 

X 

x' 

B 

4> 

H 

U 

U 

u 

X 

00" 

CQ 

« 

< 

< 

< 

m 

i_ 

< 

< 

< 

w 

Z 

Q 

j 

a: 

<u 

cd 

a 

53 

o 

a -S 

o 

a .5 

.»3 

"5 

o 2 

2 

53  "E* 

X 

2 

o 

o 

£ 

_C 

o 

•Si  g 

■g  2 

.2  C 

S! 

53 

.Si  C 

S -o 

C Q 

5 1 

Nj  C 
1 1 

•JC 

o 

6 

X g 

X g 

X % 

X % 

X -2 

Xl 

& 

a 

■j 

Vh 

3 

O S 


acervorum  distribution  1956a) 

Tetramorium  A b,c?,e?,f?  II  o/+  Morisita’s  index,  tend  to  Baroni-Urbani  (1969) 

caespitum  overdispersion 

Tetramorium  A,B  b,c,e,f  II  o/+  Intra-  and  interspecifically  Brian  et  al.  (1966) 

caespitum  territorial 


Lasius  A,B  b,c,e,f  II  o/+  R = 1.22-1.27,  dense  popula-  Waloff  and  Blackith 

flavus  tion  overdispersed,  low  dens-  (1962) 

ity  population  tends  to 
randomness 


1981] 


Levings  & Traniello  — Territoriality  in  Ants 


289 


m 

NO 

NO 

On 

NO 

ON 

*3 

On 

NO 

NO 

On 

On 

"3 

r~ 

, , 

m 

so 

3 

3 

<D 

os 

• p 

•p 

s 

u 

c 

c 

.2 

3 

>w' 

o 

o 

'C 

£ 

CL 

Oh 

00 

c 

V. 

C -2 
> ^ 

C 

3 

o 

Ih 

<L> 

3 

Ih 

O 

3 

o 

w 1 

o 

T3 

•o 

3 

1 

jd 

o 

X 

X 

T3 

3 

T3  3 

Jd 

<U 

o 

3 

'-3 

00 

c 

op 

<L> 

o 

n-4 

3 lH 

|?3  . o 

3 

CL 

3 

CL 

b 

<u 

'53 

c 

3 

CL 

X 

o 

C/3 

> 

'vi 

1/3 

C/3 

3 

CL 

>. 

<U 

3 

CL> 

>> 

O 

3 00 

>. 

T3 

c/3 

<L) 

T3 

Ih 

£ 

<o 

U g 

’5 

c 

o 

.3 

<U 

c/3 

fl) 

c 

<U 

<L> 

C/3 

Ih 

41 

<D 

c 

C/3 

o 

3 

X 

£ ts 

<u 

C/3 

CL 

g 

CL 

Vi 

"O 

O (U 

T3 

■3 

c/3 

<D 

C/3 

3 

>, 

•3 

4) 

O 

3 

CL 

° 3 

A S 

£ 

C/3 

C/3 

<D 

•3 

<u 

> 

>< 

O 

■3 

Ih 

OJ 

> 

3 

> 

w 

Id 

3 

3 

Vi 

<5 

C/3 

CL 

z 

O 

o 

z 

3 > 

'g  ^ 


•3  s 


3 > 

g ^ 


3 > 

•3  ^ 

3^ 


3 0$ 

g 5 


S5  <L 

3 Oc 

g S 


Lasius  A b,c?,e?,f?  II  o/+  Morisita’s  index,  tend  to  Baroni-Urbani  (1969) 

alienus  overdispersion 

Lasius  A,B  b,c,e,f  II  o/+  Intra-  and  interspecifically  Brian  et  al.  (1966) 

alienus  territorial 


Nest  Defense  Forager  Nest 

Species  site1  type2  type3  spacing4  Evidence5  Source 


290 


Psyche 


[Vol.  88 


</-)  so 

m so 

Os  Os 


rsi 

>rs 

Os 

13 

<N 

r- 

05 

£ & 
2 « 
w o 

s 5 

■§  i 

T3 

C 

cS 

> 

O 

L* 

11 

(50 

Tt 

X 

Os 

c 

.2 

1956a 

o 

c 

.2 

to 

t-l 

«3 

CO 

2 ® 
13  o 

> -3 
2 5 

2 o 

■S  ftS 

> -s 

« N 
2 o 

« f* 

> 

<U 

c3 

X 

<u 

'5 

N 

O 

CC 

1 s 

•g  £ 

O 

3 

3 

CO 

«3 

>- 

ffl 

ffl 

£ 

S 

25 

25 

25 

25 

.5  _o 


£ 2 
I 1 

o O 


O > 
M O 

fc  „ 
« > 


CO 


X)  ”D 

•C  o g 

.2  ^§8 
T3  o « 


O 

60 

T3 

O 

60 

C 

£ 


co 


„ > 
T3  ‘35 

(L>  CO 

*2  D 
3 i_ 
u 00 
5 60 
CL  c« 
o 
C/2 


O) 

s 

ft 


CL  2 
c « co 
3 <u 


.2  0)  O) 

> 0) 

T3  ‘5S  £ 

0)  C/5  *■> 

o o) 

cd  ln  x 

u.  GO 

(<  GO  « 

CL  & g 

CO  £ 


■ (fl 

* •§ 

T3  3 
d>  O 

3 6 

o .2 

O C) 

0> 

>s  g. 

"o 

g S3 
a "S  3 
SP  o o 


<u  0) 

o o' 

X x" 


3 

<3 

<3 

3 

3 

3 

3 

■Si 

3 

•2 

£ 

3 

5 

3 :§ 

3 

•2 

3 

3 

3 

3 

3 

Jk. 

O 

•2 

5! 

3 

o 

8 

O 

a 

tj 

O 

O 

o 

<L> 

Si 

ft, 

O 

O 

3 

6i 

O 

o 

c 

<L> 

e<5 

5 

§ 

5 

§ 

5 

% 

g 

"o 

g 

3 

L. 

5 *§ 

g 

8 

g 

.o 

'g 

3 

L. 

«2 

£ 

< 

Jk. 

£ 

<L) 

t£ 

ft 

£ 

cl 

l.  JS 

Jk. 

£ 

3 

V, 

£ 

ft, 

V. 

£ 

>S 

1981] 


Levings  & Traniello  — Territoriality  in  Ants 


291 


r~  o 

ON  3 

O 6 

2 

3 "O 

E 

1 JS 

_ vi 

3 00 

"*  X 
o 


v~i  © 

m oo 

ON  ON 


O 


"O 

■a 

i.s 


3 <D 

a*  o- 


•5S- 

I S3  .3 

o £ x> 


.3  <D 

* . & 

T3  5 

« 3 

i3  E 

O 

H «5 
§•8 
if  I-  §> 

(u  c 

■t  ’5 
cs  6 si 
??  o o 


3 

.2  x) 


oo  a, 

00  Vi 
3 ^ 

w *- 

Vi  O 

<U  > 
3 O 


•<9  3 


O 14) 

o = 

a 

•2 

c 

u 

J 

<D 

3 

3 

Vi 

3 

Oo 

2 

;3 

w 

J 

0£ 

<U 

3 

c 

§.  a 

o 

o 

’o 

V. 

<4) 

5 

O 

o 

!■* 

*-> 

£ 

1 

o 

«>5 

£ 

| 

o 

Q 

J 

O 

Uh 

CXh 

Q 

hJ 

o 

Uh 

O 

o 

I* 


lugubris  smaller  colonies,  all  (1977) 

mutually  aggressive 

Formica  A b,c,e  III  + R = 1,32,  p <C  0.001  Breen  (1979) 

lugubris 


OLD  WORLD  I 


1 

I 

go 

I 

I 

I 

I 


Nest  Defense  Forager  Nest 

Species  site1  type2  type3  spacing4  Evidence5  Source 


292 


Psyche 


[Vol.  88 


Myrmicinae 

Messor  A b?,c?,e?,f  II?  o/+?  Nonoverlap  of  foraging  trails.  Pickles  (1944) 

barbatus  interspecific  aggression 

Messor  A b?,c?,e?,f  II?  o/+?  Nonoverlap  of  foraging  trails,  Pickles  (1944) 

aegypticus  interspecific  aggression 


1981] 


Levings  & Traniello  — Territoriality  in  Ants 


293 


. - cr> 

vo  2 
r-  w 

ON 

—i  e 
W o 


'e?  J 


. „ m 

? 2 
ON  ^ 


? 2 

— 

2 c 

w o 


5 


. „ r*-> 

S 2 


<U  D 

•J?  J 


£ ^ 
£ 13 


u 3 


60 

C 

'a, 

o 

<u 

Oh 

60 

c 

'O. 

o 

u 

Oh 

so 

60 

c 

’o, 

o 

<u 

o, 

c/5 

00 

c 

'Oh 

o 

<u 

o, 

C/5 

O. 

l_ 

(U 

<u 

> 

o. 

t-H 

a> 

1> 

> 

Oh 

Lh 

<u 

u 

> 

Oh 

jd 

<U 

c 

on 

Uh 

c 

"c/5 

C 

C/5 

c/5 

Th 

c 

aj 

a> 

<L> 

aj 

<u 

<U 

o 

> 

O 

o 

u. 

GO 

> 

o 

o 

i- 

&o 

>. 

o 

o 

ob 

> 

O 

o 

c 

GO 

c 

60 

c 

00 

c 

o 

aj 

cd 

o 

£ 

(S3 

o3 

o 

Z 

S3 

03 

o 

Z 

03 

a 


D 

03 

C 

•2 

Co 

3 

& 2 

05 

Si  -2 

CO 

C 

1 

a 2 

O 3 

o 

'o 

s 

u, 

>N 

1 

"a 

a 

1 ^ 
£ 

a .a 

5 S 

a - 

§ 'C 
S -2 

<o 

castanea  aggressive 

Pheidole  A,C  b,c,d,e,f  II  o/+?  Nonoverlapping  territories,  Greenslade  (1971) 

megacephala  also  interspecifically 

aggressive 


isjiiiiUtMiii  |ii|lj||i  |i  1«  li  Ihi 


I 


D b,c,d,e,f  II, III?  o/+? 
D b,c,d,e,f  11,111?  o /+? 
C b.c.d.c.f  11,111?  o/+? 


A,C  b,c,d,e,f  II  o/+? 


I 

I 

ft* 

f 


I 


s 


Nest  Defense  Forager  Nest 

Species  site1  type2  type3  spacing4  Evidence5  Source 


294 


Psyche 


[Vol.  88 


Camponotus  A b,c,d,e,f  II?  o/+?  Nonoverlapping  territories,  Majer  (1976a); 

acvapimensis  also  interspecifically  Leston  (1973) 

aggressive 

Polyrachis  ? b,c,d,e,f  II, III?  o/+?  Intra-  and  interspecifically  Steyn  (1954) 

schitacea  aggressive 


1981]  Levings  & Traniello  — Territoriality  in  Ants  295 


Nest  Defense  Forager  Nest 

Species  site1  type2  type3  spacing4  Evidence5  Source 


296 


Psyche 


[Vol.  88 


Formica  A b?,c,e,f?  III?  o Mounds  located  along  the  edges  Scherba  (1958); 

ulkei  of  fields  Talbot  (1961) 

Formica  A b?,c,e,f?  Ill  o Mounds  located  along  the  edges  Nipson  (1978) 

exsectoides  of  fields 


Lasius  A b,c,e,f  III  + Nest  craters  overdispersed  Traniello 


1981] 


Levings  & Traniello  — Territoriality  in  Ants  297 


fulva 

Stenamma  A b,e,f?  ? o/+(!)  V/M  = 0.79,  N.S.  Headley  (1952) 

brevicorne 


jijliiliilli  jfifl  W i* j jl^l1  j ih'M 


V/M  = 0.85,  N.S. 


I 

ss 


I 

I 

i 

! 

S’ 

I 


Nest  Defense  Forager  Nest 

Species  site1  type2  type3  spacing4  Evidence5  Source 


298 


Psyche 


[Vol.  88 


rl  •§ 
r-~  S 

ON 

W >> 

>.=  00 


73  s 

cd  CQ 

t-i 

CO 


cd 

i_ 

3 _ 

e 2 

cd  w 


s'  <8 

cd  T3 

S "O 

S C 
8 * 
S*  •a 


$ 


s S 

cd  ^ 

iS  ^3 
S C 

8 «* 
a -o 


•—  <u 
X>  .2: 

c g 


• DO 

2 5510 


s s 

cd  -a 

£ -s 


S-.-2 

5 2 

c 


- ¥ 3 


DO  cd 

eo  <u 
cd  C 


<u 

«5  P 

s S 

s e 

O o 
\ n i: 

c/3  ^ 

Si  ^ 

DO  2 
00  > 
cd  cd 

S c 

II 

II 

c 


Oh  ^3 

,S2  T3  ^ 
^ X C/3 
TD  (D 

£ ” * 
° o o 

TD 

C c/3 
cd  z,  « 
cd  X) 

^ On  Cd 

on  .ti 

© £ 5 

E 2 

» o £ 


2 I 

It 

O V. 


3 

X O 

2 I 
it 

rr  Do 

2 c 
S.O 
4 


* w 
5 § 
2 .1 

II 

■T 


<U  Co 


4J 

T3 


r*  h* 

■S| 
0 « 
f 1 
~©  2 
Q 


<3  ^ 

•2  o 

I" 


Formica  A b,c,e,f  III  +?  Intraspecific  aggression;  Bradley  (1972,  1973); 

obscuripes  nest  movement  away  from  (1968) 

competitors  Bradley  and  Hinks 

(1968) 


Prenolepis  A b,e,f  III  o/+(?)  V/M  = 0.54,  N.S.  Headley  (1952) 

imparis 

Prenolepis  A b,e,f  III  o/+?  Nests  spaced  out  under  fruit  Talbot  (1943); 

imparis  trees,  aggression  over  food  Lynch  et  al.  (1980) 

sources 


1981] 


Levings  & Traniello  — Territoriality  in  Ants 


299 


s •§ 

•1  & 
5 s 


° e 

T3  -^3 


° 2 
T3 


C 

o 

o 

o 

'2 

3 

3 


00 


2 ° 

1 1 
3 


3 

5 1 
o 
c 

o C 

a,  c 


I S 

■2 

II 

I" 


<4< 

c 

a,  2 

•a  "5 

5 a, 


«U  a 

.2  1 

i * 

>> 

3 


a § 

oo  a 


Solenopsis  A b?,c?e?,f?  II?  o R = 0.6-2. 1 Bernstein  and  Gobbel 

xyloni  (1979) 


Nest  Defense  Forager  Nest 

Species  site1  type2  type3  spacing4  Evidence3 


Intraspecific  aggression,  nest 
movement  away  from  competit' 


Bradley  (1972,  1973); 
Bradley  and  Hinks 
(1968) 


Francoeur  and  Pepin 
(1978) 


Bradley  (1972,  1973); 
(1968) 


Nest  Defense  Forager  Nest 

Species  site1  type2  type3  spacing4  Evidence5  Source 


300 


Psyche 


[Vol.  88 


0 

O 

ON 

O 

OO 

0 

O 

W 

ON 

■0 

c 

a 

"O 

c 

r- 

On 

73 

o> 

73 

c 

c 

e 

C 

1 1 

‘C 

ON 

'5 

ON 

’C 

c 

0 

O 

O 

CD 

C/3 

O 

Ui 

03 

r- 

ON 

C/3 

c 

S-i 

<u 

ON 

C/3 

c 

u 

U 

>> 

PQ 

"O 

’> 

ec3 

00 

ON 

c 

0 

ss 

PQ 

PQ 

PQ 

Q 

PQ 

<U  On 

12 


.5 

'C  On 

a - 


- > 


O to 
§ 8 
s 3 

T3  is 

>>  g 


•o 

<U 

Q. 

Cl. 

e<J 


O g 
cu  o 


o g 
o.  .2 
„ o 
00  u 
o’  03 

\©  2 
© 

D. 

II  % 

& 


© 

V 

Oh 


£ 


« « <u 

xT  xf  xf 


£ 


9k. 

O 

03 

§ 

Cn 

"3 

c 

ik. 

0 

03 

03 

1 

c 

O 5 
§ | 

Jk.  ._ 

O ^ 

§ g 

■S 

03  ? 

<43  r 

5 3 
^ -a 
5 5 

03 

2*  ^ 
£ 3 

k .a 

1 s 

5 

1 

<u 

5 

S •§ 

5 | 

0 *■» 

1 

1 1 

8S 

§<§. 
rs  ^ 

p c 

O 

Jk. 

oj 

ft, 

0 

j^ 

<43 

ft, 

? p 

> 0 
£ ^ 

-s: 

So 

CU  V 

-s:  * 

1 8 

|8 

ft. 

ft, 

ft, 

ft. 

Dgonomyrmex  A b,c?,e?,f?  II?  + R = 1.14  Bernstein  and  Gobbel 

occidentalis  (1979) 


1981] 


Levings  & Traniello  — Territoriality  in  Ants 


301 


x 

X 

o 

O 

£ 2 
ON  ^ 
w C 


’5 

X 

c 

2 

a 

c 

2 

ON 

t" 

On 

’S3 

C 

a 

5— < 

<u 

ON 

r~ 

ON 

s 

X 

r- 

ON 

X 

o 

x 

o 

CO 

2 

’> 

1 

X 

r- 

ON 

X 

o 

2 

o 

C/3 

2 

’> 

o 

C/3 

2 

'> 

X 

o 

2 

s_x 

5—i 

CO 

s_^ 

X 

v ✓ 

X 

cd 

JC 

v — / 

X 

cd 

cd 

X 

CO 

£ 

X 

Q 

£ 

X 

Q 

Q 

X 

« *8 

C £ 


X 

a 

Oh 

X 

c 

&, 

co 

Oh 

X 

e 

Cl 

co 

x Ck 

c .2 

cd 

X 

cd 

X 

co 

cd 

X 

kH 

cd  X 

cd 

<u 

> 

1 

cd 

t> 

> 

3 

J3 

i 

cd 

<u 

> 

' <u 
cd  > 

C 

o 

a 

o 

2 

£ 

C 

o 

£s  o 
c 

o 

a> 

D- 

<*> 

« *8 

c 2 

'"H  <u 
X Cl, 
C .2 

Cd  TO 
k. 

' <u 

5 > 

£2  O 
C 


X 

C .2 
cd  to 
■ <u 
2 > 
£3  o 
e 


£ 


X 

cu 

5 

5 £ 

o -c 
R ?\ 
O £ 

£ 5 


5 § 


to 

o 

o 

Co 

O 

o 

o 

bo 

bo 

bo 

3 

C 

go 

3 

&0 

3 

k. 

5 3 

O «T 

c*  O 
§ ^ 
S’  2 


5 

* 3 
5 | 

o « 

R *£ 

|| 


X 

d/ 

5 

l! 

o 2 
r -£ 

|l 

On 


X 

<to 

£ 

i! 

o 2 

R *P 

|| 


<4; 

jS  5 
**• 

5 2 
o t: 

£*  dj 

§ 2 

fti 


3 

5 

k a 

b (3 
S p 
o .2 

o a 

S>  S 


a> 

X 
O 
X 
o 

Q ° 


onomyrma  A b?,c?,f?  II?  o/  + R = 0.8-1. 2 Bernstein  and  Gobbel 

insana  (1979) 


§ 


2 


Nest  Defense  Forager  Nest 

Species  site1  type2  type3  spacing4  Evidence5  Source 


302 


Psyche 


[Vol.  88 


f>  <U 

a.  c 

< 5-§ 

« K.  § 
I s ! 

ON  5 


C bQ  zl 

W = 0 

<U  M 

3Q  2 


m Os 

a £ 


Os  & Os 

p-  ~ r- 

ON  g ON 


_c  ^ 

'C  On 

II 

<D 

PQ 


a>  on 

I 2 


£ 


5 o 

Oh  O 
O o 


oo  »- 
© <u 
m g 
O O 
Oh 

II  u 

11  </3 

C* 


S 3 
c >» 
jd  3 

Oh 

' 1/3 

-S  O 

3 Oh 

s 


b .2 

g o 

I5 


<3 

5 

!■§ 

O o 

a5 


8 | 
o 

b c 


a 

S 

o ^ 

c a 


s s,  §.  s a 


2 ^ 
g 3 

2 .c 

I1 


c 

1 1 

1° 


| t 

S 5 

I" 


« .3 

.o  ^3 

I 1 

a 


Myrmecocystus  A b?,c?  II?  o/+  R = 0.9-1. 6,  depending  upon  Bernstein  and  Gobbel 

kennedei  density  (1979) 


Myrmecocystus  A b,c  II  + R = 1 .3-1.7  Bernstein  and  Gobbel 

mexicanus  (1979) 


1981] 


Levings  & Traniello  — Territoriality  in  Ants 


303 


t: 

0> 

a. 

< 


T3 

<o 

JS 

co 

3 

3 

a, 

c 


tJ 

<U 

a 

< 


S'  -2 

so  3 

r-  3 

2 o. 


3 tc 

O <u 

=3  < 
PC 


2 

-a 

c 

a 

So 

C ON 


T3 

C 

ed 

&p  oo 
c on 


w.  o 

<L> 


J5  | 

g o 
o ° 
U 


4:  -2 

■g  s 

o.  >- 


TO  u 

i:  -a 
c 


> 1 

W w 

T3  O 
c x> 
ed 

Ofi 

^ --N 
t-,  <U 
ed  C 


J3  "P 


ed 


T3  O g 
ed  £ T3 
^ .2?  & 


•3  "P 
C <U 

o .-s 

3 S 

O -J- 


O £» 
•35  £ 

K to 


3 

3 

3 

2,  3 

1 

| 

O § 

o ^ 

O 

3 

& .s> 

CJ 

1 s 

1 U 

1 

5 

X s 

v.  ^ 
^3 

>N 

5 

<3 

I S 

<3  3 


s g 

-§  -§ 

o 


a a 
-s:  £• 

$■ 


Pheidole  A,C  b,c?,e,f  II  + Clark  and  Evans’  nearest  Levings  and  Franks 

biconstricta  neighbor  analysis,  over-  (1981) 

group  species  dispersed 


Defense  Forager  Nest 
tyPe:  type3  spacing 


NEW  WORLD  Tropical  forests:  generalists 


Clark  and  Evans*  nearest 
overdispersed 

neighbor  analysis,  nests 
overdispersed 


Holldobler  (1976b); 
Alpert  (unpublished) 


Levings  and  Franks 
(1981) 


Nest  Defense  Forager  Nest 

Species  site1  type2  type3  spacing4  Evidence5  Source 


304 


Psyche 


[Vol.  88 


M 

£4 

C 

e 

as 

cd 

cd 

4> 

i-i 

Uh 

i_ 

Pm 

O.  ^ 
„ C 

•P 

"O 

r-  o 

C 

c 

r-  m 

ed 

cd 

ON  Cd 

as  cT 

as  rT 

W t 

ftp  oo 

op  oo 

c ^ 

.£ 

> w 
<U 

J 

£ ^ 
’>  w 

3 

ll 

C/3 

^ c 
r-~  o 
r-  •£ 
On  cd 

" t 

C as 

1 » 

C/3 


« > 
« O 


> | 


v-  T3 
T3  O 
C X)  £ 
ed  X3  <u 

2?  b* 

£ C .2 

03  C T3 


C/3  1) 

Si  > 

£ o 

4> 

C as 


cd  g 

> s 

w ” 

u.  T3 
"O  ° <U 
C £ £ 
cd  .c  <d 

£ u .2 

cd  C -O 

0 


o> 

■O  > 

<P  '5 


_>>  cd 
"£  af 

o -a 


C £ 
2^ 
as  . 
ed  >» 
£X 

*5  C 

* S 

aj 

«*  as 
as  ,a 
4>  Cm 
'o  O 
D <L> 

o.  o. 


3 


o 

c 

0 

X3  ed  ^ 

1 6 3 

*J  d T3 
ed  <D  w 
.S  . g 

s « .s 

C cj  — 

0 S!  O 
^ ■*-> 

1 s'  I 


5 ^ 

X)  C 


■°  o 
>»  -o 

•S  cd 

£ 

O © 

"S  ° 

£ U 


£: 


o 

C/J 

£ S 

eu  «J 


^ rn 
£ 


p ^ 


1 1 

C ?3 

.o  > 


.2  g 

II 

S’* 


NEW  WORLD  Tropical  forests:  army  ants 


NEW  WORLD  Tropics:  leaf  cutters  (Attini) 


1981]  Levings  & Traniello  — Territoriality  in  Ants  305 


Azteca  spp.  C b,c,d,e,f  III?  As  Found  only  on  Cordia  spp.  Janzen  (1969) 

plants 


srljtWi'litj  >t{ ||i* 


g 


I 

I 

i 

5' 

I 


g 


Nest  Defense  Forager  Nest 

Species  site1  type2  type3  spacing4  Evidence5  Source 


306 


Psyche 


[Vol.  88 


O 

cd 

H 


-2 

.g, 

6 


o £ 
o 73 

o o 

<+H 

d,  o 

c n 1) 
C/3  Q, 

<3  E 

5 a 

8 .s 


e 5 'S3 
cd  c/3 
' <u 

•a  d u 

c is  M 

is  s 


< jd 

’d 


aj 

6 

£ 

o 

T3 


■§ 

3 


cd  u 

g s 8. 

c/3 

iii. 

^ fc  3 T 
8 ^ S S 
ft, 


a> 

•2  o 

« & 

? 1/5 

v r;  ^ 
c J2  + 
£ g CJ/ 


>> 

s 


I 

o 

•jc 

!§ 


macrops  (in  prep.) 

Myrmecia  A b?,e  I + Intraspecifically  territorial  Muir  (1974) 

tarsata  and  overdispersed 

Myrmecia  A b?,e  I + Intraspecifically  territorial  Muir  (1974) 

simillima  and  overdispersed 


Myrmecia  A b?,e  I + Intraspecifically  territorial  Muir  (1974) 

decipiens  and  overdispersed 


1981] 


Levings  & Traniello  — Territoriality  in  Ants 


307 


c m 

ca  in 

S 2 


cx 

— ' C/5 

ca  • — 
o TJ 
Eh 

'o  > 
x>  O 
cx 

c/5  "O 

nj  c 


rz  o 

ca  T3 


C/5 


ac  ci; 
ca  ^ 

~ *o  g- 
<u  c t3 

s i g 

S £ o 
c 


_c«  x) 
S ca 

C t_ 

o 


X3  O 

C/5 

'•3  2 

e - 
E CX 

S £ 


E ’S 

S -o 


s t 

> g 

[jq  ca 

t- 

T3  O 
E X) 
aj  X 

^ .SP 

t-  X) 

ca  c 


60 

E 

c/5  ‘5b  ^ 

60  2 X) 
C o X) 

•-  V C/5 

e e 

S.  g>  8. 

° leg 

c/3  O Jr; 
<L>  O S> 
c > 

u u ° 

<«  O 


X)  T3  O 

u ea 
ea  ex 

E c/5 

ca  c ca 

CX  3 X) 

<U  XX  E 

c«  o ca 


.o  -2 

S S 
s -a 

X 5, 


fc  5» 

-2  -S; 
^3  ^3 


^ 3 

£ S- 


H 
! = 

! t 

<3 


Nest  Defense  Forager  Nest 

Species  site1  type2  type3  spacing4 Evidence5 


308 


Psyche 


[Vol.  88 


H 


-rt  >-* 

« O 

c/3  O 

1 e 

X 5 

s 

o ^ 

> 


S O 


4>  O 
2 O 
u C 

& a 


w £ 

« to 
<u 

Ul 

s £ 

> 


O 4> 

;r!  c/3 

^ 3 

O 

4>  i— i 

Oh  2 

C/2  1—1 

<a 


*3  £ 

<L)  ‘-3 
Ch  3 


G C 
o ca 
o 


. „ to 
4) 

o 2 
" o 

-T-l  CO 
^ 4) 

O i- 
O 

£ s 

w 4) 

^ 2 
"3  c/3 

> t-i 
<3  4> 

O CU 
X . 
4)  CJ 

8 

Uh  _ „ 

O c/3 

<+-i  4> 

4>  O 

l_ 

Oh  3 


oo 

t 2 

o 


2 o 

C/3 

C/3 

- £ 

03  !> 

a x. 

C/3  C 


OJO 


lo 


Oh 

s § ! 

4>  Jr 
c o 

° X ° 

.-a  ,u  c 
3 5t3  o 
o •§  g 

^ 6 o, 

_J~  W ' C/3 


60  > 
3 O 


Oh  C/3 

3 X 
O 3 

l~  QJ 
60  -* 


I 5 o 

* > i 

4>  O 


2 J3 


g .3  $ -S 

.3  &3  ici  Chj 

•a  -3  "3  -3  ^ 

6 £ o,  o .o 

8 £ $ 


3 

s .3  8 

§ J 


I1 


u 

0 

C/3 

<L> 

4)  fv  4) 
3^1- 

U. 

3 — 4> 

O 3 C4 

M 

4> 

8 

u 

’*  .a  .2 

O H-I  X 

O 

4) 

£ ^3  £ 

t-i 

4> 

U 

4> 

c/3 

g 

&■»  > 
3 3° 

X 

3 

O 

-3 

<U 

C*H 

^ ° + 

3 

C/3 

4) 

X 

§.j 

03 

xT 

0 

u cd  r 

X 

‘0 

4> 

O ’o  4) 

5 0. 
2 g<£ 

3 

'3 

Oh 

4/^3 

3 

0 

C/3 

60  - — 

O 

4) 

2 > 

s 

t- 

4) 

C/3 

3 

3 

<2  ~ r 

° c« 
X « 

c -o 
5 

O C/3 
I-  4» 
60  C 

c« 


- | * 6 

. « .’2  2 c 

4>  > *-  X 

C X ^ « 
,4>  C 3 t- 


5V/M,  variance  to  mean  ratio;  !,  our  measurements  or  statistics  based  on  published  data;  R,  Clark  and  Evans’  nearest 
neighbor  analysis,  R = 1,  random  dispersion,  R > 1,  overdispersed,  R < 1,  clumped 
6Mabelis  (1979)  studied  this  species  in  dunes;  it  is  usually  found  in  woodland. 


1981] 


Levings  & Traniel/o  — Territoriality  in  Ants 


309 


competitors  are  removed  can  provide  additional  evidence  for  com- 
petitive effects  on  spacing  and  foraging  patterns. 

Acknowledgements 

This  paper  is  a synthesis  of  portions  of  the  doctoral  dissertations 
of  both  authors.  We  would  like  to  thank  the  following  people  for 
their  assistance  and  helpful  criticism:  G.  Alpert,  F.M,  Carpenter, 
N.R.  Franks,  S.D.  Garrity,  B.  Holldobler,  R.  Levins,  R.  Lewontin, 
B.L.  Thorne,  and  E.O.  Wilson.  Supported  by  the  Anderson  and 
Richmond  Funds  of  Harvard  University,  NSF  Grant  BNS  80-02613 
(B.  Holldobler,  sponsor),  and  NSF  predoctoral  grants  to  both 
authors. 


Literature  Cited 


Abe,  T. 

1971.  On  the  food  sharing  among  four  species  of  ants  in  a sandy  grassland,  I. 
Food  and  foraging  behavior.  Japanese  Journal  of  Ecology  20:  219-230. 
Adams,  E.,  and  J.  F.  A.  Traniello. 

1981.  Chemical  interference  competition  by  Monomorium  minimum. 
Oecologia  in  press. 

Alloway,  T.  M. 

1980.  The  origin  of  slavery  in  leptothoracine  ants  (Hymenoptera:  Formicidae). 
American  Naturalist  115:  247-261. 

Ashton,  D.  H. 

1979.  Seed  harvesting  by  ants  in  forests  of  Eucalyptus  regnans  F.  Muell  in 
central  Victoria.  Australian  Journal  of  Ecology  4:  265-277. 

Autori,  M. 

1941.  Contribuicao  para  o conhecimento  da  sauva  ( Atta  spp.  Hymenoptera: 
Formicidae)  ( Atta  sexdens  rubripilosa  Forel  1980).  Arquivos  do  Institu- 
te Biologico  12:  197-228. 

Baroni-Urbani,  C. 

1969.  Ant  communities  of  the  high  altitude  Appenine  grasslands.  Ecology  50: 
488-492. 

Baroni-Urbani,  C. 

1979.  Territoriality  in  social  insects.  In:  The  Social  Insects,  H.  R.  Hermann, 
ed.  Academic  Press,  N.Y. 

Bernstein,  R.  A. 

1974.  Seasonal  food  abundance  and  foraging  activity  in  some  desert  ants. 
American  Naturalist  108:  490-498. 

Bernstein,  R.  A. 

1975.  Foraging  strategies  of  ants  in  response  to  variable  food  density.  Ecology 
56:  213-219. 

Bernstein,  R.  A. 

1979.  Schedules  of  foraging  activity  in  species  of  ants.  Journal  of  Animal 
Ecology  48:  921-930. 


310 


Psyche 


[Vol.  88 


Bernstein,  R.  A.,  and  M.  Gobbel. 

1979.  Partitioning  of  space  in  communities  of  ants.  Journal  of  Animal  Ecology 
48:  931-942. 

Blackith,  R.  E.,  J.  W.  Siddorn,  N.  Waloff,  and  H.  F.  van  Emden. 

1963.  Mound  nests  of  the  yellow  ant,  Lasius  flavus  L.  on  waterlogged  pasture 
in  Devonshire.  Entomologist’s  Monthly  Magazine  99:  48-49. 

Bradley,  G.  A. 

1972.  Transplanting  Formica  obscuripes  and  Dolichoderus  taschenbergi  (Hy- 
menoptera:  Formicidae)  colonies  in  jack  pine  stands  of  southeastern 
Manitoba.  Canadian  Entomologist  104:  245-249. 

Bradley,  G.  A. 

1973.  Interference  between  nest  populations  of  Formica  obscuripes  and  Do- 
lichoderus taschenbergi  (Hymenoptera:  Formicidae).  Canadian  Ento- 
mologist 105:  1525-1528. 

Bradley,  G.  A.,  and  J.  D.  Hinks. 

1968.  Ants,  aphids  and  jack  pines  in  Manitoba.  Canadian  Entomologist  100: 
40-50. 

Breen,  J. 

1979.  Nest  sites  of  Formica  lugubris  (Hymenoptera:  Formicidae)  in  Irish 
plantation  woods.  Journal  of  Life  Sciences,  Royal  Dublin  Society  1: 
13-32. 

Brian,  M.  V. 

1952.  The  structure  of  a dense  natural  ant  population.  Journal  of  Animal 
Ecology  21:  12-24. 

Brian,  M.  V. 

1955.  Food  collection  by  a Scottish  ant  community.  Journal  of  Animal 
Ecology  24:  336-351. 

Brian,  M.  V. 

1956a.  The  natural  density  of  Myrmica  rubra  and  associated  ants  in  west 
Scotland.  Insectes  Sociaux  3:  473-487. 

Brian,  M.  V. 

1956b.  Segregation  of  species  of  the  ant  genus  Myrmica.  Journal  of  Animal 
Ecology  25:  319-337. 

Brian,  M.  V. 

1964.  Ant  distribution  in  a southern  English  heath.  Journal  of  Animal 
Ecology.  33:  451-461. 

Brian,  M.  V.,  J.  Hibble,  and  D.  Stradling. 

1965.  Ant  pattern  and  density  in  a southern  English  heath.  Journal  of  Animal 
Ecology  34:  544-555. 

Brian,  M.  V.,  J.  Hibble,  and  A.  F.  Kelly. 

1966.  The  dispersion  of  ant  species  in  a southern  English  heath.  Journal  of 
Animal  Ecology  35;  281-290. 

Brough,  E.  J. 

1976.  Notes  on  the  ecology  of  an  Australian  desert  species  of  Calomyrmex 
(Hymenoptera:  Formicidae).  Journal  of  the  Australian  Entomological 
Society  15:  339-346. 


1981] 


Levings  & Traniello  — Territoriality  in  Ants 


311 


Brown,  W.  L.,  and  E.  O.  Wilson. 

1959.  The  evolution  of  the  dacetine  ants.  Quarterly  Review  of  Biology  34: 
278-294. 

Byron,  P.  A.,  E.  R.  Byron,  and  R.  A.  Bernstein. 

1980.  Evidence  of  competition  between  two  species  of  desert  ants.  Insectes 
Sociaux  27:  351-360. 

Carroll,  R.,  and  D.  H.  Janzen. 

1973.  Ecology  of  foraging  by  ants.  Annual  Review  of  Ecology  and  Systematics 
4:  231-257. 

Cherix,  D.,  and  G.  Gris. 

1977.  The  giant  colonies  of  redwood  ant  in  the  Swiss  Jura  ( Formica  lugubris 
Zet.).  Proceedings  of  the  Eighth  International  Congress  of  International 
Union  for  the  Study  of  Social  Insects.  Wageningen,  the  Netherlands. 
Centre  for  Agricultural  Publishing  and  Documentation,  Wageningen. 

Clark,  P.  J.,  and  F.  C.  Evans. 

1954.  Distance  to  nearest  neighbor  as  a measure  of  spatial  relationships  in 
populations.  Ecology  35:  445-453. 

Clark,  P.  J.,  and  F.  C.  Evans. 

1979.  Generalization  of  a nearest  neighbor  measure  of  dispersion  for  use  in  K. 
dimensions.  Ecology  60:  316-317. 

Colwell,  R.,  and  D.  Futuyma. 

1971.  On  the  measurement  of  niche  breadth  and  overlap.  Ecology  52:  567-576. 

Corn,  M.  L. 

1976.  The  ecology  and  behvior  of  Cephalotes  atratus;  a Neotropical  ant 
(Hymenoptera:  Formicidae).  Unpublished  Ph.D.  dissertation,  Harvard 
University. 

Culver,  D. 

1974.  Species  packing  in  Caribbean  and  north  temperate  ant  communities. 
Ecology  55:  974-988. 

Davidson,  D. 

1977a.  Species  diversity  and  community  organization  in  desert  seed-eating  ants. 
Ecology  58:  711-724. 

Davidson,  D. 

1977b.  Foraging  ecology  and  community  organization  in  desert  seed-eating 
ants.  Ecology  58:  725-737. 

Davidson,  D. 

1980.  Some  consequences  of  diffuse  competition  in  a desert  ant  community. 
American  Naturalist  116:  92-105. 

DeVita,  J. 

1979.  Mechanisms  of  interference  and  foraging  among  colonies  of  the  har- 
vester ant  Pogonomyrmex  calif ornicus  in  the  Mojave  desert.  Ecology  60: 
729-737. 

DeVroey,  C. 

1979.  Aggression  and  Gausse’s  Law  in  ants.  Physiological  Entomology  4: 
217-222. 

Diver,  C. 

1935.  A population  of  wood  ants  ( Formica  sp.)  at  Abernethy.  Journal  of 
Animal  Ecology  4:  32-34. 


312 


Psyche 


[Vol.  88 


Dobrzanska,  J. 

1958.  Partition  of  foraging  grounds  and  modes  of  conveying  information 
among  ants.  Acta  Biologiae  Experimentalis,  (Warsaw)  18:  55-67. 

Dobrzanska,  J. 

1966.  The  control  of  territory  by  Lasius  fuliginosus  Latr.  Acta  Biologiae 
Experimentalis,  (Warsaw)  26:  193-213. 

Duncan-Weatherly,  A.  H. 

1953.  Some  aspects  of  the  biology  of  the  mound  ant  Iridomyrmex  detectus. 
Australian  Journal  of  Zoology  1:  178-192. 

Eisenberg,  R.  M. 

1972.  Partitioning  of  space  among  colonies  of  the  fire  ants,  Solenopsis 
saevissima,  I.  Spatial  arrangement.  Texas  Journal  of  Science  24:  39-43. 

Elmes,  G.  W. 

1971.  An  experimental  study  on  the  distribution  of  heathland  ants.  Journal  of 
Animal  Ecology  40:  495-499. 

Elmes,  G.  W. 

1974.  The  spatial  distribution  of  a population  of  two  ant  species  living  in 
limestone  grassland.  Pedobiologia  14:  412-418. 

Elton,  C.  S. 

1932.  Territory  among  wood  ants  ( Formica  rufa  L.)  at  Picket  Hill.  Journal  of 
Animal  Ecology  1:  69-76. 

Fluker,  S.  S.,  and  J.  W.  Beardsley. 

1970.  Sympatric  associations  of  three  ants:  Iridomyrmex  humilis,  Pheidole 
megacephala  and  Anoplolepis  longipes  in  Hawaii.  Annals  of  the  Ento- 
mological Society  of  America  63:  1290-1296. 

Fra  'coeur,  A.,  and  D.  PeppiN. 

i978.  Productivitep  de  la  four  mi  Formica  dakotensis  dans  la  pessiepre  tour- 
beuse.  2.  Variations  annuelles  de  la  densitep  des  colonies,  de  l’occupation 
des  nids  et  de  la  repartition  spatiale.  Insectes  Sociaux  25:  13-30. 

Franks,  N. 

1980.  The  evolutionary  ecology  of  the  army  ant  Eciton  burchelli  on  Barro 
Colorado  Island,  Panama.  Unpublished  Ph.D.  dissertation,  The  Univer- 
sity of  Leeds,  Leeds,  England. 

Greenslade,  P.  J.  M. 

1971.  Interspecific  competition  and  frequency  changes  among  ants  in  the 
Solomon  Islands  coconut  plantations.  Journal  of  Applied  Ecology  8: 
323-352. 

Greenslade,  P.  J.  M. 

1975.  Dispersion  and  history  of  a population  of  the  meat  ant  Iridomyrmex 
purpureus  (Hymenboptera:  Formicidae).  Australian  Journal  of  Zoology 
23:  495-510. 

Headley,  A.  E. 

1941.  A study  of  the  nest  and  nesting  habits  of  the  ant  Lasius  niger  subsp. 
alienus  var.  americanus  Emery.  Annals  of  the  Entomological  Society  of 
America  34:  649-657. 

Headley,  A.  H. 

1952.  Colonies  of  ants  in  a locust  woods.  Annals  of  the  Entomological  Society 
of  America  45:  435-442. 


1981] 


Levings  & Traniello  — Territoriality  in  Ants 


313 


Higashi,  S.,  and  K.  Yamauchi. 

1979.  Influence  of  a supercolonial  ant  Formica  (Formica)  yessensis  on  the 
distribution  of  other  ants  in  Ishikari  Coast.  Japanese  Journal  of  Ecology 
29:  257-264. 

Holldobler,  B. 

1973.  Chemische  strategic  beim  nahrungserwerb  der  Diebsameise  ( Solenopsis 
fugax  Latr.)  und  der  Pharaoameise  ( Monomorium  pharoanis  L.)  Oeco- 
logia  11:  371-380. 

Holldobler,  B. 

1974.  Home  range  orientation  and  territoriality  in  harvesting  ants.  Proceed- 
ings of  the  National  Academy  of  Sciences,  USA  71:  3274-3277. 

Holldobler,  B. 

1976a.  Recruitment  behavior,  home  range  orientation  and  territoriality  in 
harvesting  ants.  Behavioral  Ecology  and  Sociobiology  1:  3-44. 

Holldobler,  B. 

1976b.  Tournaments  and  slavery  in  a desert  ant.  Science  192:  912-914. 

Holldobler,  B. 

1977.  Communication  in  social  Hymenoptera.  Pp.  418-471  in  T.  A.  Sebeok, 
ed.,  How  animals  communicate.  Indiana  University  Press,  Bloomington. 

Holldobler,  B. 

1979a.  Territoriality  in  ants.  Proceedings  of  the  American  Philosophical  So- 
ciety 123:  211-218. 

Holldobler,  B. 

1979b.  Territories  of  the  African  weaver  ant  ( Oecophylla  longinoda  (Latreille)): 
A field  study.  Zeitschrift  fur  Tierpsychologie  51:  201-213. 

Holldobler,  B.,  and  C.  Lumsden. 

1980.  Territorial  strategies  in  ants.  Science  210:  732-739. 

Holldobler,  B.,  and  M.  Moglich. 

1980.  The  foraging  system  of  Pheidole  militicida.  Insectes  Sociaux  27:  237- 
264. 

Holldobler,  B.,  and  E.  O.  Wilson. 

1970.  Recruitment  trails  in  the  harvester  ant  Pogonomyrmex  badius.  Psyche 
77:  385-399. 

Holldobler,  B.,  and  E.  O.  Wilson. 

1977a.  Weaver  ants:  social  establishment  and  maintenance  of  territory.  Science 
195:  900-902. 

Holldobler,  B.,  and  E.  O.  Wilson. 

1977b.  Colony-specific  territorial  pheromone  in  the  African  weaver  ant  Oeco- 
phylla longinoda  (Latreille).  Proceedings  of  the  National  Academy  of 
Science,  USA  74:  2072-2075. 

Holldobler,  B.,  and  E.  O.  Wilson. 

1977c.  The  number  of  queens:  an  important  trait  in  ant  evolution.  Naturwissen- 
schaften  64:  8-15. 

Holldobler,  B.,  Stanton,  R.  C.,  and  H.  Markl. 

1978.  Recruitment  and  food-retrieving  behavior  in  Novomessor  (Formicidae, 
Hymenoptera)  I.  Chemical  signals.  Behavioral  Ecology  and  Socio- 
biology 4:  163-181. 


314 


Psyche 


[Vol.  88 


Hunt,  J. 

1974.  Temporal  activity  patterns  in  two  competing  ant  species  (Hymenoptera: 
Formicidae).  Psyche  81:  237-242. 

Ito,  M.,  and  S.  Imamura. 

1974.  Observations  on  the  nuptial  flight  and  internidal  relationship  in  a 
polydomous  ant,  Formica  (F.)  yessensis  Forel.  Journal  of  the  Faculty  of 
Science,  Hokkaido  University,  Series  VI,  Zoology  19:  681-694. 

Janzen,  D.  H. 

1967.  Interaction  of  the  bull’s-horn  acacia  ( Acacia  cornigera  L.)  with  an  ant 
inhabitant  ( Pseudomyrmex  ferruginea  F.  Smith)  in  eastern  Mexico. 
Kansas  University  Science  Bulletin  47:  315-558. 

Janzen,  D.  H. 

1969.  Allelopathy  by  myrmecophytes:  the  ant  Azteca  as  an  allelopathic  agent 
of  Cecropia.  Ecology  50:  147-153. 

Janzen,  D.  H. 

1972.  Protection  of  Barteria  (Passifloraceae)  by  Pachysima  ants  (Pseudo- 
myrmicinae)  in  a Nigerian  rain  forest.  Ecology  53:  885-892. 

Janzen,  D.  H. 

1973.  Evolution  of  polygynous  obligate  acacia  ants  in  western  Mexico. 
Journal  of  Animal  Ecology  42:  727-750. 

Janzen,  D.  H. 

1975.  Pseudomyrmex  nigropilosa:  A parasite  of  a mutualism.  Science  188: 
936-937. 

Leston,  D. 

1973.  The  ant  mosaic  in  tropical  tree  crops  and  the  limiting  of  pests  and 
diseases.  Pest  Articles  and  News  Summaries  19:  311-341. 

Levieux,  J. 

1971.  Mise  en  epvidence  de  la  structure  des  nids  et  de  l’implantation  des  zones 
de  chasse  de  deux  espeQces  de  Camponotus  (Hym.,  Form.)  aQ  l’aide  de 
radio-isotopes.  Insectes  Sociaux  18:  29-48. 

Levieux,  J. 

1977.  La  nutrition  des  fourmis  tropicales:  V.  Elepments  de  syntheQse.  Les 
modes  d’exploitation  de  la  biocoenose.  Insectes  Sociaux  24:  235-260. 

Levieux,  J.,  and  D.  Louis. 

1975.  La  nutrition  des  fourmis  tropicales.  II.  Comportement  alimentaire  et 
regime  de  Camponotus  vividus  (Smith)  (Hymenoptera:  Formicidae). 
Comparison  intragenerique.  Insects  Sociaux  22:  391-404. 

Levings,  S.  C.,  and  N.  R.  Franks. 

1982.  Patterns  of  nest  dispersion  in  a tropical  ground  ant  community.  Ecology, 
63  in  press. 

Levings,  S.  C.,  and  D.  M.  Windsor. 

1982.  Seasonal  and  annual  variation  in  litter  arthropod  populations.  In  E.  G. 
Leigh,  A.  S.  Rand,  and  D.  M.  Windsor,  eds.,  The  ecology  of  a tropical 
forest:  seasonal  rhythms  and  longer  term  changes.  Smithsonian  Institu- 
tion Press,  Washington,  D.C.  In  press. 

Lewis,  T. 

1975.  Colony  size,  density  and  distribution  of  the  leaf  cutting  ant,  Acromyr- 


1981] 


Levings  & Traniello  — Territoriality  in  Ants 


315 


mex  octospinosus  (Reich)  in  cultivated  fields.  Transactions  of  the  Royal 
Entomological  Society  of  London  127:  51-64. 

Lewis,  T.,  G.  V.  Pollard,  and  G.  C.  Dibley. 

1974.  Rhythmic  foraging  in  the  leaf  cutting  ant  Atta  cephalotes.  Journal  of 
Animal  Ecology  43:  129-141. 

Lynch,  J.,  E.  C.  Balinsky,  and  S.  G.  Vail. 

1980.  Foraging  patterns  in  three  sympatric  forest  ant  species,  Prenolepis 
imparts,  Paratrechina  melanderi  and  Aphaenogaster  rudis  (Hymenop- 
tera:  Formicidae).  Ecological  Entomology  5:  353-371. 

Mabelis,  A.  A. 

1979.  The  relationship  between  aggression  and  predation  in  the  red  wood  ant 
(Formica  polyctena  FoRrst.).  Netherlands  Journal  of  Zoology  29: 
451-620. 

MacArthur,  R.  H.,  and  R.  Levins. 

1967.  The  limiting  similarity,  convergence  and  divergence  of  coexisting  species. 
American  Naturalist  101:  377-385. 

Majer,  J.  D. 

1976a.  The  maintenance  of  the  ant  mosaic  in  Ghana  cocoa  farms.  Journal  of 
Applied  Ecology  13:  123-144. 

Majer,  J.  D. 

1976b.  The  ant  mosaic  in  Ghana  cocoa  farms:  further  structural  considerations. 
Journal  of  Applied  Ecology  13:  145-153. 

Markl,  H.,  and  B.  Holldobler. 

1978.  Recruitment  and  food-retrieving  behavior  in  Novomessor  (Formicidae, 
Hymenoptera)  II.  Vibration  Signals.  Behavioral  Ecology  and  Sociobiol- 
ogy 4:  183-216. 

Marikovsky,  P.  I. 

1963.  The  ant  Formica  sanguinea  Latr.  as  pillagers  of  Formica  rufa  Lin.  nests. 
Insectes  Sociaux  10:  119-128. 

Morglich,  M.,  and  G.  Alpert. 

1979.  Stone-dropping  by  Conomyrma  bicolor:  a new  kind  of  interference 
behavior  in  ants.  Behavioral  Ecology  and  Sociobiology  6:  105-114. 

Muir,  R.  J. 

1974.  A study  of  the  coexistence  of  four  sympatric  species  of  Myrmecia 
(Formicidae).  Unpublished  Ph.D.  dissertation,  University  of  New  Eng- 
land, Armidale,  New  South  Wales,  Australia. 

Nipson,  H.  E. 

1978.  Inbreeding  in  the  ant  species  Formica  exsectoides.  Unpublished  Ph.D. 
dissertation,  Harvard  University. 

Oster,  G.,  and  E.  O.  Wilson. 

1978.  Caste  and  ecology  in  the  social  insects.  Princeton  University  Press, 
Princeton,  N.J. 

Petal,  J. 

1972.  Methods  of  investigating  the  productivity  of  ants.  Ekologia  Polska  20: 
9-22. 

Petal,  J. 

1977.  The  effect  of  food  supply  and  intraspecific  competition  in  an  ant 


316 


Psyche 


[Vol.  88 


population.  Proceedings  of  the  Eighth  International  Congress  of  the 
International  Union  for  the  Study  of  Social  Insects,  5-10  September 
1977,  Wageningen,  The  Netherlands,  pp.  60-61. 

PlANKA,  E.  R. 

1978.  Evolutionary  ecology,  2nd  ed.  Harper  and  Row,  New  York. 

Pickles,  W. 

1944.  Territories  and  interrelations  of  two  ants  of  the  genus  Messor  in  Algeria. 
Journal  of  Animal  Ecology  13:  128-129. 

PlELOU,  E.  C. 


1977.  Mathematical  ecology,  2nd  ed.  John  Wiley,  New  York. 

Pisarski,  B. 

1972.  La  structure  des  colonies  polycaliques  de  Formica  (Coptoformica) 
exsect  a Nyl.  Ekologia  Polska  20:  113-116. 

Pontin,  A.  J. 

1960.  Field  experiments  on  colony  foundation  by  Lasius  niger  (1.)  and  Lasius 
flavus  (F.)  (Hym.,  Formicidae).  Insectes  Sociaux  7:  227-230. 

Pontin,  A.  J. 


1961.  Field  experiments  on  colony  foundation  by  Lasius  niger  (L.)  and  L. 
flavus  (F.)  (Hym.:  Formicidae).  Journal  of  Animal  Ecology  38:  47-54. 

Pontin,  A.  J. 

1963.  Further  consideration  of  competition  and  the  ecology  of  the  ants  Lasius 
flavus  (F.)  and  L.  niger  (L.).  Journal  of  Animal  Ecology  32:  565-574. 

Pontin,  A.  J. 

1969.  Experimental  transplantation  of  nest  mounds  of  the  ant  Lasius  flavus 
(F.)  in  a habitat  containing  also  L.  niger  (L.)  and  Myrmica  scabrinodis 
Nyl.  Journal  of  Animal  Ecology  38:  747-754. 

Poole,  R.  E. 

1974.  An  introduction  to  quantitative  ecology.  McGraw-Hill,  New  York. 

Reichman,  O.  J. 

1979.  Desert  granivore  foraging  and  its  impact  on  seed  densities  and  distribu- 
tion. Ecology  60:  1085-1092. 

Robertson,  P.  L. 

1971.  Pheromones  involved  in  aggressive  behavior  in  the  ant  Myrmecia 
gulosa.  Journal  of  Insect  Physiology  17:  691-715. 

Rockwood,  L. 

1973.  Distribution,  density  and  dispersion  of  two  species  of  Atta  (Hymenop- 
tera:  Formicidae)  in  Guanacoste  Province,  Costa  Rica.  Journal  of 
Animal  Ecology  42:  803-807. 


Room,  P.  M. 

1971.  The  relative  distributions  of  ant  species  in  Ghana’s  cocoa  farms.  Journal 
of  Animal  Ecology  40:  735-751. 

Room,  P.  M. 

1975a.  Relative  distributions  of  ant  species  in  cocoa  plantations  in  Papua,  New 
Guinea.  Journal  of  Applied  Ecology  12:  47-61. 


Room,  P.  M. 

1975b.  Diversity  and  organization  of  the  ground  foraging  ant  faunas  of  forest, 
grassland  and  tree  crops  in  Papua,  New  Guinea.  Australian  Journal  of 
Zoology  23:  71-89. 


1981] 


Levings  & Traniello  — Territoriality  in  Ants 


317 


Sanders,  C.  J. 

1970.  The  distribution  of  carpenter  ant  colonies  in  the  spruce  fir  forests  of 
northwestern  Ontario.  Ecology  51:  865-873. 

Scherba,  G. 

1958.  Reproduction,  nest  orientation  and  population  structure  of  an  aggrega- 
tion of  mound  nests  of  Formica  ulkei,  Emery  (Formicidae).  Insectes 
Sociaux  5:  201-213. 

Scherba,  G. 

1964.  Species  replacement  as  a factor  affecting  distribution  of  Formica 
opaciventris  Emery  (Hymenoptera:  Formicidae).  Journal  of  the  New 
York  Entomological  Society  72:  231-237. 

Schneirla,  T.  C. 

1971.  Army  ants:  a study  in  social  organization.  W.  H.  Freeman  and  Co.,  San 
Francisco. 

Schoener,  T. 

1971.  Theory  of  feeding  strategies.  Annual  Review  of  Ecology  and  Systematics 
2:  369-404. 

SlMBERLOFF,  D.  S. 

1979.  Nearest  neighbor  assessments  of  spatial  configurations  of  circles  rather 
than  points.  Ecology  60:  679-685. 

Skinner,  G.  H. 

1980.  Territory,  trail  structure  and  activity  patterns  in  the  wood  ant,  Formica 
rufa  (Hymenoptera:  Formicidae)  in  limestone  woodland  in  northwest 
England.  Journal  of  Animal  Ecology  49:  381-394. 

STEBAEV,  I.,  AND  J.  I.  REZNIKOVA. 

1972.  Two  interaction  types  of  ants  living  in  steppe  ecosystem  in  South 
Siberia,  USSR.  Ekologia  Polska  20:  103-109. 

Steyn,  J.  J. 

1954.  The  pugnacious  ant  ( Anoplolepis  custodiens  Smith)  and  its  relation  to 
the  control  of  citrus  scales  at  Letaba.  Memoirs  of  the  Entomological 
Society  of  South  Africa  3:  1-96. 

Sudd,  J.  H.,  T.  M.  Douglas,  T.  Gaynard,  D.  M.  Murray,  and  J.  M.  Stockdale. 

1977.  The  distribution  of  wood  ants  ( Formica  lugubris  Zetterstedt)  in  a 
northern  English  forest.  Ecological  Entomology  2:  301-313. 

Swain,  R. 

1977.  The  natural  history  of  Monads,  a genus  of  Neotropical  ants.  Unpub- 
lished Ph.D.  dissertation,  Harvard  University. 

Talbot,  M. 

1943.  Population  studies  of  the  ant  Prenolepis  imparis.  Ecology  24:  31-44. 

Talbot,  M. 

1951.  Population  and  hibernating  conditions  of  the  ant  Aphaenogaster  (At- 
tomyrma)  rudis  Emery  (Hymenoptera:  Formicidae).  Annals  of  the 
Entomological  Society  of  America  44:  302-307. 

Talbot,  M. 

1954.  Populations  of  the  ant  Aphaenogaster  (Attomyrma)  treatae  Forel  on 
abandoned  fields  on  the  Edwin  S.  George  Reserve.  Contribution  from 
the  Laboratory  of  Vertebrate  Biology  of  the  University  of  Michigan,  no. 
69,  pp.  1-9. 


318 


Psyche 


[Vol.  88 


Talbot,  M. 

1957.  Populations  of  ants  in  a Missouri  woodland.  Insectes  Sociaux  4: 
375-384. 

Talbot,  M. 

1961.  Mounds  of  the  ant  Formica  ulkei  at  the  Edwin  S.  George  Reserve, 
Livingston  County,  Michigan.  Ecology  42:  202-205. 

Talbot,  M. 

1967.  Slave  raids  of  the  ant  Polyergus  lucidus  Mayr.  Psyche  74:  299-313. 

Talbot,  M.  and  C.  H.  Kennedy. 

1940.  The  slave  making  ant,  Formica  sanguinea  subintegra  Emery,  its  raids, 
nuptial  flights  and  nest  structure.  Annals  of  the  Entomological  Society 
of  America  32:  304-315. 

Thorne,  B.  L.,  and  K.  P.  Sebens. 

1981.  Regulation  of  social  insect  colony  size:  a model  based  on  foraging 
energetics.  Unpublished  manuscript. 

Traniello,  J.  F.  A. 

1980.  Studies  on  the  behavioral  ecology  of  north  temperate  ants.  Unpublished 
Ph.D.  dissertation,  Harvard  University. 

Waloff,  N.,  and  R.  E.  Blackith. 

1962.  The  growth  and  distribution  of  mounds  of  Lasius  flavus  (F.)  (Hymenop- 
tera:  Formicidae)  in  Silwood  Park,  Berkshire.  Journal  of  Animal 
Ecology  31:  421-437. 

Ward,  P.  S.,  and  R.  W.  Taylor. 

1981.  Allozyme  variation,  colony  structure  and  genetic  relatedness  in  a 
primitive  ant  Nothomyrmecia  macrops  (Hymenoptera:  Formicidae). 
Unpublished  manuscript. 

Way,  M.  J. 

1953.  The  relationship  between  certain  ant  species  with  particular  reference  to 
biological  control  of  the  coreid,  Theraptus  sp.  Bulletin  of  Entomological 
Research  45:  93-112. 

Whitford,  W.  G.,  E.  Depree,  and  P.  Johnson. 

1980.  Foraging  ecology  of  two  Chihuahuan  desert  ant  species:  Novomessor 
cockerelli  and  N.  albisetosus.  Insectes  Sociaux  27:  148-156. 

Whitford,  W.  G.,  P.  Johnson,  and  J.  Ramirez. 

1976.  Comparative  ecology  of  the  harvester  ants  Pogonomyrmex  barbatus  and 
P.  rugosus.  Insectes  Sociaux  23:  117-132. 

Wilson,  E.  O. 

1959a.  Communication  by  tandem  running  in  the  ant  genus  Cardiocondyla. 
Psyche  66:  29-34. 

Wilson,  E.  O. 

1959b.  Some  ecological  characteristics  of  ants  in  New  Guinea  rain  forests. 
Ecology  40:  437-447. 

Wilson,  E.  O. 

1971.  The  insect  societies.  Belknap  Press  of  Harvard  University  Press,  Cam- 
bridge, Mass. 

Wilson,  E.  O. 

1975.  Enemy  specification  in  the  alarm-recruitment  system  of  an  ant.  Science 
190:  798-800. 


1981] 


Levings  & Traniello  — Territoriality  in  Ants 


319 


Wilson,  N.  L.,  J.  H.  Diller,  and  G.  P.  Markin. 

1972.  Foraging  territories  of  imported  Fire  ants.  Annals  of  the  Entomological 
Society  of  America  64:  660-665. 

Yasuno,  M. 

1963.  The  study  of  the  ant  population  in  the  grassland  at  Mt.  Hakkoda.  I.  The 
distribution  and  the  abundance  of  ants  in  the  grassland.  Ecological 
Review  16:  83-91. 

Yasuno,  M. 

1964a.  The  study  of  the  ant  population  in  the  grassland  at  Mt.  Hakkoda,  II.  The 
distribution  pattern  of  ant  nests  at  the  Kayano  grassland.  Science 
Reports  of  the  Tohoku  University,  Sendai,  Japan,  ser.  4 (Biol.)  39: 
43-55. 

Yasuno,  M. 

1964b.  The  study  of  the  ant  population  in  the  grassland  at  Mt.  Makkoda,  III. 
The  effect  of  the  slave  making  ant,  Polyergus  samurai,  upon  the  nest 
distribution  pattern  of  the  slave  ant,  Formica  fusca  japonica.  Science 
Reports  of  the  Tohoko  University,  Sendai,  Japan,  ser.  4 (Biol.)  30: 
167-170. 

Yasuno,  M. 

1965.  Territory  of  ants  in  the  Kayano  grassland  at  Mt.  Hakkoda.  Science 
Reports  of  the  Tohoko  University,  Sendai  Japan,  ser  4 (Biol.)  31: 
195-206. 


THE  EFFECT  OF  FLOWER  OCCUPANCY  ON  THE 
FORAGING  OF  FLOWER-VISITING  INSECTS* 


By  V.  J.  Tepedino  and  F.  D.  Parker 
Bee  Biology  & Systematics  Laboratory,  Agricultural  Research, 
Science  & Education  Admin., 

USDA 

Utah  State  University,  UMC  53 
Logan,  Utah 

Introduction 

To  locate  flowers,  insects  use  a variety  of  visual  and  olfactory  cues 
such  as  flower  color,  shape,  movement  and  scent  (Faegri  and  van 
der  Pijl  1971).  In  addition,  other  insects  on  the  flowers  may  also 
serve  as  cues  that  either  attract  or  repel  prospective  foragers.  First, 
foragers  might  avoid  occupied  inflorescences  because:  1)  there  is  a 
high  probability  that  other  flowers  on  the  inflorescence  have  been 
recently  exploited  (Pleasants  and  Zimmerman  1979,  Zimmerman 
1981);  2)  of  the  potential  loss  of  time  and  energy  due  to  aggressive 
encounter  with  the  occupant  (Kikuchi  1963,  Decelles  and  Laroca 
1979);  3)  the  occupant  might  be  an  enemy  (e.g.,  thomisids,  phyma- 
tids,  etc.).  Thus,  when  flowers  are  abundant,  unoccupied  inflor- 
escences may  yield  a greater  quantity  of  energy  and/or  nutrients  per 
unit  effort.  If  so,  the  distribution  of  foragers  across  inflorescences 
should  be  regular  or  underdispersed,  i.e.,  there  should  be  more 
inflorescences  with  only  one  insect  than  expected  on  the  assumption 
of  a random  distribution. 

Existing  evidence  also  suggests  that  a second  hypothesis  is 
tenable.  Prospective  foragers  may  be  attracted  by  floral  occupants 
because:  1)  the  presence  of  other  foragers  indicates  that  resources 
are  available  on  the  inflorescence;  2)  the  occupants  themselves  are 
sources  of  pollen  to  some  foragers  (Laroca  and  Winston  1978, 
Thorp  and  Briggs  1980).  If  insects  are  attracted  to  occupied 
inflorescences,  then  their  distribution  across  inflorescences  should 
be  over-dispersed. 


* Manuscript  received  by  the  editor  October  20,  1981 


321 


322 


Psyche 


[Vol.  88 


In  this  paper  we  use  data  for  insects  foraging  on  plantings  of 
commercial  sunflowers  ( Helianthus  annuus  L.)  and  onions  {Allium 
cepa  L.)  to  test  these  hypotheses. 

An  additional  question  of  interest  is  whether  bee  species  differ  in 
their  distribution  across  flowers.  For  example,  Benest  (1976)  has 
suggested  that  honeybees  {Apis  mellifera  L.)  are  more  tolerant  of 
joint  foraging  than  are  bumblebees  {Bombus  sp.)  and  Kalmus  (1954) 
reported  that  honeybees  tend  to  form  clusters  at  artificial  feeding 
sites.  Group  foraging,  leading  to  clumped  distributions  on  flowers 
has  also  been  reported  for  several  tropical  bee  species  (Frankie  and 
Baker  1974).  To  ascertain  if  the  distribution  of  the  multispecies 
assemblage  obscured  differences  among  the  component  species,  we 
compared  the  distributions  of  the  more  abundant  species  with  the 
balance  of  other  foraging  individuals  on  the  inflorescences. 


Methods 

Five  cultivars  of  sunflower  and  2 of  onions  were  grown  at  the 
Greenville  Farm  Agricultural  Research  Station  in  North  Logan, 
Utah.  Sunflowers  were  planted  in  5 adjacent  40m  rows,  1 row  per 
cultivar.  The  2 onion  cultivars  were  planted  alternately  in  4 adjacent 
rows,  2 rows  per  cultivar. 

Counts  of  floral  visitors  were  made  several  times  during  the 
flowering  period  as  1 observer  (FDP)  walked  along  each  row.  A 
tape  recorder  facilitated  observations.  Only  heads  with  some  open 
flowers  were  censused. 

The  data  were  transcribed  to  number  of  flower  heads  with  zero, 
one,  two,  etc.  insects  and  then  compared  with  values  expected  on 
the  assumption  of  a Poisson  distribution  (Southwood  1978).  The 
Poisson  series  describes  a random  distribution  and  is  written  Px(k)  = 
e"x(xk/K!)  where  e = base  of  Napierian  logarithms,  and  Px  is  the 
expected  number  of  flower  heads  with  k insects  (k  = 0,  1, 2,—).  The 
parameter  x is  estimated  by  the  mean  number  of  insects  per  flower 
head.  For  the  Poisson  distribution,  the  mean  and  variance  are 
equal,  and  an  indication  of  the  dispersion  of  insects  across  flowers  is 
given  by  the  coefficient  of  dispersion  (C.D.  = s2/x).  When  C.D.  is 
>1.0  the  dispersion  is  clumped  or  contagious;  and  when  <1.0 
dispersion  is  regular  or  repulsed  (Southwood  1978).  The  expected 


1981] 


Tepedino  & Parker  — Flower-Visiting  Insects 


323 


and  observed  distributions  were  tested  for  significance  using  the  x2 
test  (Zar  1974). 

The  distributions  of  more  abundant  species  across  sunflower 
heads  was  compared  with  the  balance  of  the  foraging  assemblage  as 
follows:  each  individual  recorded  was  assigned  to  one  of  two 
mutually  exclusive  categories,  according  to  whether  it  foraged  alone 
or  with  at  least  one  other  insect  (irrespective  of  species)  on  the 
inflorescence.  A chi-square  test  of  independence  was  used  to  com- 
pare each  species  represented  by  >10  individuals  with  the  balance  of 
the  assemblage. 


Results 

Bees  were  the  predominant  visitors  to  sunflowers;  we  recorded  15 
species  in  5 families  (Appendix).  The  species  were  similar  to  that 
reported  previously  by  Parker  (1981)  for  the  same  study  site.  Onion 
visitors  included  many  species  of  wasps  and  flies  that  did  not  forage 
on  sunflowers.  In  contrast  to  sunflowers,  there  were  more  non-bee 
than  bee  visitors  to  onions. 

For  all  sunflower  censuses  the  distribution  of  total  insects  across 
flower  heads  did  not  differ  significantly  from  a Poisson  distribution, 
i.e.,  insects  appeared  to  be  foraging  independently  of  other  insects. 
The  coefficients  of  dispersion  were  mostly  around  1.0.  There  was  no 
tendency  for  C.D.’s  to  be  greater  or  less  than  1;  for  8 censuses  C.D. 
was  >1.0  and  for  6 censuses  C.D.  <1.0.  (Table  1). 

Only  2 of  7 censuses  of  onions  deviated  significantly  from  a 
random  distribution  (Table  1).  Both  deviations  occurred  on  the 
same  day  and  were  in  the  direction  of  under-dispersion;  more  heads 
with  single  visitors  were  recorded  than  expected.  There  was  a 
general  tendency  for  insects  visitors  to  be  under-dispersed  on 
onions;  in  all  tests  C.D.  > 1.0. 

There  was  no  indication  that  any  particular  species  foraged  other 
than  randomly,  with  respect  to  other  occupants  of  sunflower  heads. 
The  results  of  34  comparisons  of  the  distribution  of  individuals  of 
abundant  species  with  the  balance  of  foragers  for  the  single  and 
joint  foraging  categories  are  shown  in  Table  2.  Only  one  comparison 
yielded  significant  results;  another  closely  approached  significance 
(7/31  Peredovik,  AM,  Halictus  ligatus , P.  = 0.051).  It  is  likely  that 
these  two  instances  were  due  to  chance. 


Table  1.  Total  flower  heads,  mean  insects  per  head,  coefficient  of  dispersion  (C.D.  = s2/x)  and  probability  levels  (X2  test)  for 
insects  visiting  sunflower  and  onion  heads.  All  counts  made  between  1000-1 100  hrs  except  those  with  asterisks  which  were  made 


324 


Psyche 


[Vol.  88 


A A A v V A A 


OO  Tf  Tf  — ' Tfr  SO  <N 

Os  OK  ON  Os  OO  OS  Os 

© o’  o’  o'  © © o’ 


OO  cn  SO  Os  SO  g 
O © O O © © 


N 'O  ^ O O O W 

r-  ->3-  < <n  u-s  os 

r-~  oo  ^ so  m t}-  m 


$ 

K -rt  ( N 't  (N  t 
N IT)  CL  n fl,  IT) 

> CU  CU  -1  Cu  -tL  Cu 


osno^nsriosnsnoosnoo 

in  (N  Id  (N  (S  - ■ CN  © — ■ — ' <N  </"S  — I 
OOOOOOOOOOOOO 

AAAAAAAAAAAAA 


_„,_0_000„0„0__ 


«__0— 'OOOOOOOO 


cd  -o 

2 8 

f— 1 JS 


r-Ttommoo— 'OOTtr-^so 
sfinnxici-o-<doso0'o 
<N  — - <N  — I <N  CN 


M M 
‘>  ‘> 
o o 

S T3  T3 

•Si  a>  <u 

U L*  •—  M 

Q D « r/s  a 

> CU  CU  Oh 

^ m o —i 


M 

> 

O 

0-0  0 
00  <L>  oo  o 
00 


> 

5 S - * 

<L>  iP  o o 

S3  £ % £ » 

:>  oo 


C/3  .3*  CU  C/3  <N 

C/3  — — 


> * 

I | 


/ 254*  160  0.45  1.03  >0.25 

/ W501*  195  0.61  1.04  >0.50 


1981] 


Tepedino  & Parker  — Flower-Visiting  Insects 


325 


Discussion 

In  this  study,  foraging  insects  did  not  appear  to  react  to  the 
presence  of  other  insects  in  choosing  flowers.  Only  2 of  the  censuses 
on  onions  and  none  of  the  censuses  on  sunflowers  displayed  a 
significant  departure  from  a random  distribution  (Table  1).  Sun- 
flower foragers  (Apis,  Perdita,  Halictus)  frequently  entered  the 
flower  by  landing  on  the  back  of  the  petals  or  on  the  involucral 
bracts  and  then  crawling  onto  the  head.  If  occupancy  by  another 
insect  were  important,  this  would  be  an  inefficient  method  of 
choosing  a flower.  In  a similar  study  Waddington  (1976)  also 
concluded  that  halictid  bees  were  foraging  independently  on  bind- 
weed ( Convolvulus  arvensis).  None  of  the  abundant  species 
present  appeared  to  forage  other  than  randomly  with  respect  to 
other  flower  occupants.  This  was  especially  surprising  for  honeybees 
which  have  been  reported  to  more  readily  tolerate,  or  even  form, 
clumped  distributions  (Kalmus  1954,  Benest  1976).  However,  con- 
tagious distributions  of  honeybees  may  occur  only  under  unusual 
circumstances;  the  data  of  Kalmus  (1954)  were  gathered  from  a 
small  number  of  feeding  dishes  and  are  quite  artificial.  Benest’s 
(1976)  suggestion  that  honeybees  are  more  tolerant  of  joint  foraging 
than  bumblebees  does  not  stand  close  examination.  Additional 
study  is  required  before  such  conclusions  are  warranted. 

Instead  of  using  the  presence  of  insects  on  inflorescences  as  cues, 
some  flower-visiting  insects  may  make  selections  based  on  the 
number  of  open  flowers  or  the  amount  of  nectar  or  pollen  available. 
Although  all  heads  censused  had  some  open  flowers,  some  had  more 
open  flowers  than  others  and  insects  may  have  been  choosing  those 
heads  with  more  flowers  irrespective  of  other  visitors.  Even  if  heads 
were  equivalent  in  number  of  flowers,  continuous  removal  of  nectar 
and  pollen  by  foragers  would  cause  variation  in  resource  availability 
between  heads  (e.g.,  Pleasants  and  Zimmerman  1979)  and  insects 
may  be  responsive  to  such  variation  prior  to  landing  on  a flower. 
For  example,  Thorp  et  al.  (1975)  have  suggested  that  the  fluorescent 
nectar  (and  perhaps  pollen)  of  many  species  with  open  flowers  may 
be  used  as  a cue  by  foraging  insects  (see  also  Kevan  1976,  Thorp  et 
al.  1976);  and  onion  nectar  is  intensely  fluorescent  (Thorp  et  al. 
1975).  Recently  Heinrich  (1979)  has  shown  that  bumblebee  foragers 
reject  many  more  nectar  depleted  (recently  visited)  white  clover 
( Trifolium  repens)  heads  than  heads  with  abundant  nectar.  Rejec- 


326 


Psyche 


[Vol.  88 


3 

■©  o, 

5 £ 


3:  a 

~S  | 
I1  a3 


Q 

V. 

•2 

ci, 

5 


1 1 
^ - 


“3 

§ -2 


I I ll  l 2 l I I l I l l I i 


I I I I I 


l I I 


l l I 


I o | «n  <n 
I O loo 


— « "3-  O oo  r*-> 
O o’  <N  m'  <N 


I I I 


I I I 


I II  II  M I I 


m — o 
O o'  o' 


<n  rn  r-  o o m o 
O — <’  — o'  <N  © o’ 


I I I I I I 


I I 


I II  I I I 


o m m - n 
o’  — < © (N  — ' o’ 


"NO 

o'  o'  d 


M M 
> > 
o o 
•o  "O 

u «U 


M 
’> 
o 

o -g 
60  .S3  „ 


> 

° © 
O O "g  »/"> 
6>0  SO  t-  ^ 


> * 

o o g ® 
60  S3  £ 


<UCL»  4J  4>  <D  <U  ^ 


Lp  <0.025 


1981] 


Tepedino  & Parker  — Flower-Visiting  Insects 


327 


tion  was  accomplished  without  landing  and  the  cue  was  probably 
scent  of  nectar  (Heinrich  1979).  Future  field  studies  should  explore 
the  use  of  these  more  subtle  cues  by  foraging  insects. 

Acknowledgments 

We  thank  Drs.  Ivan  Palmbad  and  Richard  Rust  for  reviewing  the 
manuscript. 

Appendix 

Insect  taxa  visiting  sunflower  and  onion  plantings. 


Sunflowers 

Hymenoptera: 

Bees — Andrenidae  [ Andrena  helianthi  Robert- 
son, Perdita  sp.,  Pseudopanurgus  sp.,  Ptero- 
sarus  sp.];  Anthophoridae  [Melissodes  agilis 
Cresson,  Svastra  obligua  (Say),  Triepeolus  heli- 
anthi (Robertson)];  Apidae  [ Apis  mellifera  Lin- 
naeus, Bombus  spp.];  Halictidae  [Agapostemon 
sp.,  Dialictus  sp.,  Halictus  farinosus  Smith, 
Halictus  ligatus  Say];  Megachilidae  [Megachile 
paralella  Smith,  Megachile  pugnata  Say]. 

Diptera: 

Lepidoptera: 

Syrphidae 

Hesperiidae 

Onions 

Hymenoptera: 

Bees — Apidae  [Bombus  sp.];  Halictidae  [Evy- 
laeus  sp,  Halictus  farinosus  Smith,  Halictus 
ligatus  Say];  Megachilidae  [Hoplitis  fulgida 
(Cresson),  Megachile  pacifica  (Panzer),  Mega- 
chile sp.]. 

Wasps — Eumenidae  [Euodynerus  sp.,  Ptero- 
cheilus  sp.];  Ichneumonidae;  Sphecidae  [Am- 
mophila  sp.,  Astata  sp.,  Cerceris  sp.,  Philanthus 
sp.,  Podalonia  sp.,  Sphex  sp.,  Tachytes  sp.] 

Diptera: 

Muscidae;  Nemestrinidae;  Sarcophagidae;  Syr- 
phidae; Tachinidae. 

Lepidoptera: 

Hesperiidae 

328 


Psyche 


[Vol.  88 


Literature  Cited 


Benest,  G. 

1976.  Relations  interspecifiques  et  intraspecifiques  entre  butineuses  de  Bom- 
bus  sp.  et  d 'Apis  mellifica  L.  Apidologie,  7:113-127. 

Decelles,  P.  and  S.  Laroca 

1979.  Behavioral  interactions  among  solitarily  foraging  bees  (Hymenoptera: 
Apoidea).  J.  Kansas  Entomol.  Soc.,  52:483-488. 

Faegri,  K.  and  L.  Van  Der  Pijl 

1971.  The  Principles  of  Pollination  Ecology.  2nd  Ed.  Rev.  London.  Pergamon 
Press.  291  pp. 

Frankie,  G.  W.  and  H.  G.  Baker 

1974.  The  importance  of  pollinator  behavior  in  the  reproductive  biology  of 
tropical  trees.  An.  Inst.  Biol.  Univ.  Nal.  Auton.  Mexico  45  Ser. 
Botanica,  ( 1 ):  1 — 10. 

Heinrich,  B. 

1979.  Resource  heterogeneity  and  patterns  of  movement  in  foraging  bumble- 
bees. Oecologia,  40:235-245. 

Kalmus,  H. 

1954.  The  clustering  of  honeybees  at  a food  source.  Brit.  J.  Anim.  Behav., 
2:63-71. 

Kevan,  P.  G. 

1976.  Fluorescent  nectar.  Science,  194:341-342. 

Kikuchi,  T. 

1963.  Studies  on  the  coaction  among  insects  visiting  flowers.  III.  Dominance 
relationship  among  flower-visiting  flies,  bees  and  butterlies.  Sci.  Rep. 
Tohoku  Univ.  Ser.  IV  (Biol.),  29:1-8. 

Laroca,  S.  and  M.  L.  Winston 

1978.  Interaction  between  Apis  and  Bombus  (Hymenoptera:  Apidae)  on  the 
flowers  of  tall  thistle:  honeybees  gather  pollen  from  bodies  of  bumble- 
bees. J.  Kansas  Entomol.  Soc.,  51:274-275. 

Parker,  F.  D. 

1981.  Sunflower  pollination:  abundance,  diversity  and  seasonality  of  bees  and 
their  effect  on  yields.  J.  Apic.  Res.,  20:49-61. 

Pleasants,  J.  M.  and  M.  Zimmerman 

1979.  Patchiness  in  the  dispersion  of  nectar  resources:  evidence  for  hot  and 
cold  spots.  Oecologia,  41:283-288. 

Southwood,  T.  R.  E. 

1978.  Ecological  Methods,  2nd  Ed.  Rev.  London:  Chapman  & Hall.  524  pp. 

Thorp,  R.  W.,  D.  L.  Briggs,  J.  R.  Estes,  and  E.  H.  Erickson 

1976.  Fluorescent  nectar.  Science,  194:341-342. 

Thorp,  R.  W.,  D.  L.  Briggs,  J.  R.  Estes  and  E.  H.  Erickson 

1975.  Nectar  fluorescence  under  ultraviolet  radiation.  Science,  189:476-478. 

Thorp.  R.  W.  and  D.  L.  Briggs 

1980.  Bees  collecting  pollen  from  other  bees  (Hymenoptera:  Apoidea).  J. 
Kansas  Entomol.  Soc.,  53:166-170. 


1981] 


Tepedino  & Parker  — Flower-Visiting  Insects 


329 


Waddington,  K.  D. 

1976.  Foraging  patterns  of  halictid  bees  at  flowers  of  Convulvulus  arvensis. 
Psyche,  83:112-119. 

Zar,  J.  H. 

1974.  Biostatistical  Analysis.  Englewood  Cliffs,  N.  J.:  Prentice-Hall. 
Zimmerman,  M. 

1981.  Patchiness  in  the  dispersion  of  nectar  resources:  Probable  causes. 
Oecologia,  49:154-157. 


ABDOMINAL  TROPHALLAXIS  IN  THE  SLAVE-MAKING 
ANT,  HARPAGOXENUS  AMERICANUS 
(HYMENOPTERA:  FORMICIDAE)* 

Robin  J.  Stuart 

Department  of  Zoology,  Erindale  College, 

University  of  Toronto, 

Mississauga,  Ontario,  L5L  1C6  Canada 

Abdominal  trophallaxis  refers  to  the  passage  of  fluids  from  the 
abdominal  tip  of  one  individual  to  the  mouthparts  of  another.  It  is 
common  among  lower  termites  (Kalotermitidae  and  Rhinotermiti- 
dae)  where  it  functions  in  the  vital  transmission  of  intestinal 
flagellates  to  newly  molted  individuals.  However,  it  has  rarely  been 
documented  among  ants  (Wilson,  1971,  1976).  By  strict  definition, 
the  term  abdominal  trophallaxis  should  be  applied  only  when 
alimentary  fluid  is  being  transmitted  (Wilson,  1971).  Nevertheless, 
in  practice,  the  origin  of  the  fluid  is  often  unknown,  at  least  initially. 
Indeed,  in  all  cases  where  this  behavior  has  been  described  in  ants, 
the  fluid  is  either  suspected  of  being,  or  has  since  been  shown  to  be, 
ovarian  in  nature.  For  example,  workers  of  certain  Eciton  species 
(Dorylinae)  readily  feed  from  droplets  secreted  from  the  tip  of  the 
queen’s  abdomen,  but  this  behavior  has  been  observed  only  during 
egg-laying  bouts  (Schneirla,  1944;  Rettenmeyer,  1963).  So-called 
“proctodeal  feeding”  has  also  been  described  among  the  Doli- 
choderinae  ( Dolichoderus  quadripunctatus,  Tapinoma  erraticum 
and  Iridomyrmex  humilis)  (Torossian,  1958,  1959,  1960,  1961). 
However,  at  least  in  the  case  of  D.  quadripunctatus,  the  fluid  has 
been  identified  as  the  yolky  remnants  of  abortive  trophic  eggs 
(Torossian,  1978,  1979).  Among  the  Myrmicinae,  Zacryptocerus 
varians  exhibits  a similar  behavior  which  is  also  thought  to  be 
associated  with  egg-laying  (Wilson,  1976).  This  paper  reports  an 
unusual  and  interesting  case  of  abdominal  trophallaxis  in  colonies 
of  the  socially  parasitic  myrmicine  ant,  Harpagoxenus  americanus.  I 
have  followed  Wilson  (1976)  and  tentatively  applied  the  term 
abdominal  trophallaxis,  because  the  origin  of  the  fluid  is  unknown. 

H.  americanus  is  an  obligatory  slave-maker  and  forms  mixed 
colonies  with  members  of  certain  Leptothorax  species  in  eastern 

* Manuscript  received  by  the  editor  November  10,  1981. 


331 


332 


Psyche 


[Vol.  88 


North  America  (Alloway,  1979).  The  observations  reported  here 
took  place  in  artificial  nests  in  the  laboratory  (Alloway,  1979)  and 
utilized  colonies  of  H.  americanus  containing  one  or  both  of  its  host 
species,  L.  ambiguus  and  L.  longispinosus.  The  colonies  were 
collected  in  the  regional  municipalities  of  Halton  and  Peel,  Ontario. 

Intermittent  observations  of  activity  inside  H.  americanus  nests 
revealed  that  slave-maker  queens  and  workers  occasionally  convey 
fluids  to  their  slaves  by  means  of  abdominal  trophallaxis.  Donors 
characteristically  raise  their  abdomens  and  assume  a stereotyped 
posture,  similar  to  that  seen  during  “sexual  calling”  (“Locksterzeln”) 
in  other  leptothoracine  ants  (Buschinger  and  Alloway,  1979).  I was 
unable  to  ascertain  whether  the  sting  is  exposed  in  the  present 
context,  as  it  is  during  “sexual  calling”  and  “tandem  calling”,  a 
similar  behavior  used  during  nest-mate  recruitment  in  some  lepto- 
thoracine ants  (Moglich  et  al.,  1974).  While  maintaining  this 
posture,  the  donor  secretes  a droplet  of  clear  fluid  from  the  tip  of 
her  abdomen,  and  holds  it  there,  at  times  for  several  minutes.  Slaves 
do  not  seem  to  be  attracted  from  any  appreciable  distance  by  this 
behavior,  but  those  close  by  turn  and  antennate  the  donor’s 
abdomen,  apply  their  mouthparts  to  the  tip,  and  consume  the 
droplet.  As  many  as  three  slaves  have  been  observed  to  attend  a 
donor  simultaneously  in  this  manner,  clustered  about  her  abdominal 
tip  and  attempting  to  consume  the  droplet.  On  one  occasion,  the 
droplet  was  removed  from  the  donor’s  abdomen  by  three  workers  in 
concert,  held  between  their  mandibles  momentarily,  and  then 
consumed.  Once  the  droplet  is  removed,  the  donor  lowers  her 
abdomen,  and  both  donor  and  recipients  appear  to  resume  normal 
activities.  There  is  no  indication  that  slaves  ever  solicit  this  fluid; 
and  to  date,  the  reverse,  slaves  donating  to  slave-makers,  has  not 
been  observed.  Similarly,  this  behavior  has  never  been  observed  in 
laboratory  colonies  of  the  host  species.  The  nature  of  the  fluid 
transmitted  is  unknown.  It  may  be  ovarian  in  origin,  since  H. 
americanus  workers  will  lay  eggs,  even  in  queenright  colonies 
(Buschinger  and  Alloway,  1977).  The  frequency  of  this  behavior  is 
uncertain.  It  appears  to  be  rare,  since  frequent  observations  of 
colonies  for  other  purposes  have  seldom  encountered  it.  However, 
no  detailed  behavioral  repertoire  study  of  this  ant  has  been  con- 
ducted. 

The  fact  that  H.  americanus  employs  a characteristic  posture 
during  abdominal  trophallaxis  suggests  that  this  behavior  may  have 


1981] 


Stuart  — Trophallaxis  in  Harpagoxenus 


333 


important  biological  consequences.  Furthermore,  the  apparent  ab- 
sence of  this  behavior  in  free-living  leptothoracine  ants,  and  the  fact 
that  transmission  is  consistently  from  slave-maker  to  slave,  suggests 
that  abdominal  trophallaxis  may  in  some  way  contribute  to  this 
species’  particular  socially  parasitic  relationship.  The  discovery  of 
this  behavior  in  a slave-making  ant  opens  a previously  unknown 
avenue  for  consideration  in  discussions  of  the  means  by  which  slave- 
makers  may  affect  the  behavior  of  their  slaves. 


Acknowledgments 

I thank  T.M.  Alloway  for  his  comments  on  the  manuscript. 
Financial  assistance  was  provided  by  an  Ontario  Graduate  Scholar- 
ship to  the  author,  and  a grant  from  the  Natural  Sciences  and 
Engineering  Research  Council  of  Canada  to  T.M.  Alloway. 


Literature  Cited 

Alloway,  T.  M.  1979.  Raiding  behaviour  of  two  species  of  slave-making  ants, 
Harpagoxenus  americanus  (Emery)  and  Leptothorax  duloticus  Wesson 
(Hymenoptera:  Formicidae).  Anim.  Behav.  27:  202-210. 

Buschinger,  A.,  and  T.  M.  Alloway.  1977.  Population  structure  and  poly- 
morphism in  the  slave-making  ant  Harpagoxenus  americanus  (Emery) 
(Hymenoptera:  Formicidae).  Psyche  83:  233-242. 

1979.  Sexual  behaviour  in  the  slave-making  ant  Harpagoxenus  cana- 
densis M.  R.  Smith,  and  sexual  pheromone  experiments  with  H.  canadensis, 
H.  americanus  (Emery),  and  H sublaevis  (Nyl.)  (Hymenoptera:  Formicidae). 
Z.  Tierpsychol.  49:  113-119. 

Moglich,  M.,  U.  Maschwitz,  and  B.  Holldobler.  1974.  Tandem  calling: 
a new  kind  of  signal  in  ant  communication.  Science  186:  1046-1047. 

Rettenmeyer,  C.  W.  1963.  Behavioral  studies  of  army  ants.  Kan.  Univ.  Sci.  Bull. 
44:  281-465. 

Schneirla,  T.  C.  1944.  The  reproductive  functions  of  the  army-ant  queen  as 
pace-makers  of  the  group  behavior  pattern.  J.  N.  Y.  Entomol.  Soc.  52:  153-192. 

Torossian,  C.  1958.  L’aliment  proctodeal  chez  la  fourmi  Dolichoderus  quadri- 
punctatus  (Dolichoderidae).  C.  R.  Acad.  Sci.  246:  3524-3526. 

1959.  Les  echanges  trophallactiques  proctodeaux  chez  la  fourmi  Doli- 
choderus quadripunctatus  (Hymenoptere-Formicoidea).  Ins.  Soc.  6:  369-374. 

1960.  Les  echanges  trophallactiques  proctodeaux  chez  la  fourmi  Tapi- 

noma  erraticum.  Ins.  Soc.  7:  174-175. 

1961.  Les  echanges  trophallactiques  proctodeaux  chez  la  fourmi  d’ Ar- 
gentine: Iridomyrme x humilis  (Hym.  Form.  Dolichoderidae).  Ins.  Soc. 

8:  189-191. 


334 


Psyche 


[Vol.  88 


1978.  La  ponte  d’oeufs  abortifs  chez  les  ouvrieres  de  la  fourmi  Doli- 

choderus  quadripunctatus.  Soc.  Hist.  Nat.  Toulouse  Bull.  114:  207-211. 

1979.  Importance  quantitative  des  oeufs  abortifs  d’ouvrieres  dan  le 

bilan  trophique  de  la  colonie  de  la  fourmi  Dolichoderus  quadripunctatus. 
Ins.  Soc.  26:  295-299. 

Wilson,  E.  O.  1971.  The  insect  societies.  Harvard  Univ.  Press,  Cambridge,  Mass. 

x + 548  pp. 

1976.  A social  ethogram  of  the  neotropical  arboreal  ant  Zacryptocerus 

various  (Fr.  Smith).  Anim.  Behav.  24:  354-363. 


NEW  NAME  FOR  THE  EXTINCT  GENUS  MESAGYRTES 
PONOMARENKO  (COLEOPTERA:  SILPHIDAE  S.L.) 


By  Alfred  F.  Newton,  Jr. 

Museum  of  Comparative  Zoology 
Harvard  University,  Cambridge,  Mass.  02138 

Mesagyrtes  communis  Ponomarenko,  a new  beetle  genus  and 
species  attributed  to  Silphidae,  has  recently  been  described  from 
fossil-bearing  beds  of  Jurassic  age  from  the  locality  of  Novospassk, 
USSR  (Arnoldi  et  al.,  1977:  117).  Unfortunately  the  generic  name  is 
preoccupied  by  Mesagyrtes  Broun  (1895:  95),  proposed  for  a Recent 
New  Zealand  species  originally  placed  in  Silphidae;  this  genus  is 
now  considered  a subgenus  of  the  genus  Colon  Herbst  of  the  family 
Leiodidae  (Szymczakowski  1964). 

I have  brought  the  homonymy  to  the  attention  of  Dr.  Ponoma- 
renko, who  has  kindly  allowed  me  to  propose  a replacement  name 
for  use  in  publications  on  the  family  Silphidae  now  in  preparation. 
Accordingly,  I propose  Mesecanus,  new  name,  to  replace  Mesa- 
gyrtes Ponomarenko  (not  Broun).  The  new  name  alludes  to  the 
resemblance  in  habitus  between  the  extinct  genus  and  the  Recent 
agyrtine  silphid  genus  Ecanus  Stephens. 

Literature  Cited 

Arnoldi,  L.  V.,  V.  V.  Zherikhin,  L.  M.  Nikritin  and  A.  G.  Ponomarenko 
1977.  “Mesozoic  Coleoptera”  [in  Russian].  Trudi  Paleont.  Inst.  Akad.  Nauk 
SSSR  161,  204  pp. 

Broun,  T. 

1895.  Descriptions  of  new  Coleoptera  from  New  Zealand.  Ann.  Mag.  Nat. 
Hist.  (6)15:  67-88. 

Szymczakowski,  W. 

1964.  Revision  des  Colonidae  (Coleoptera)  des  regions  orientale  et  austra- 
lienne.  Acta  Zool.  Cracov.  9(8):  1e59. 
lienne.  Acta  Zool.  Cracov.  9(8):  1-59. 


335 


HISTORICAL  DEVELOPMENT  OF  BEE  FORAGING 
PATTERNS  IN  CENTRAL  NEW  YORK  STATE 

By  Howard  S.  Ginsberg* 

Department  of  Entomology 
Cornell  University 
Ithaca,  New  York  14853 


Introduction 

The  bee  fauna  of  the  northeastern  United  States  has  changed 
markedly  in  the  past  few  centuries.  The  impetus  for  this  change 
came  largely  from  human  activities,  notably  from  introductions  of 
foreign  species  and  modifications  of  the  regional  flora.  Several  bee 
species,  most  notably  the  honey  bee  ( Apis  mellifera),  were  intro- 
duced into  this  region  (Crane  1975;  Linsley  1958).  Honey  bees  can 
powerfully  influence  the  foraging  patterns  of  native  bees  (Pearson 
1933;  Eickwort  and  Ginsberg  1980).  Replacement  of  forests  over 
large  areas  by  cities  and  farms  (Ferguson  and  Mayer  1970;  Vaughan 
1929)  and  numerous  introductions  of  alien  plant  species  (Wiegand 
and  Eames  1925)  have  resulted  in  major  changes  in  northeastern 
plant  communities. 

How  broad  were  these  changes  and  how  have  they  influenced  the 
foraging  ecology  of  northeastern  bees?  What  was  this  area  like 
before  the  European  settlers  arrived?  The  answers  to  these  questions 
are  vital  to  an  understanding  of  contemporary  bee  foraging  patterns 
and  of  community  level  interactions  between  flowers  and  their 
pollinators.  The  purpose  of  this  paper  is  to  describe  some  general 
trends  in  the  foraging  patterns  of  Apoidea  in  central  New  York 
State,  and  to  interpret  them  in  terms  of  the  historical  development 
of  the  flora  and  bee  fauna  of  the  region. 

Materials  and  Methods 

The  study  site  was  a 5.8  hectare  abandoned  field  (last  cultivated 
about  1956)  located  near  Ithaca,  New  York.  It  was  bordered  by 


♦Present  Address:  Department  of  Ecology  and  Evolution,  State  University  of  New 
York  at  Stony  Brook,  Stony  Brook,  New  York  11794 
Manuscript  received  by  the  editor  August  8,  1981 


337 


338 


Psyche 


[Vol.  88 


wooded  areas  and  cultivated  fields.  The  soils  were  well-drained  and 
flower  bloom  was  profuse.  More  than  150  entomophilous  species 
bloomed  on  the  field.  The  most  common  woody  plants  were  red 
maple  (Acer  rubrum),  staghorn  sumac  (Rhus  typhina ),  and  various 
willows  (Salix  spp.),  dogwoods  (Cornus  spp.),  and  brambles  (Rubus 
spp.).  The  dominant  herbaceous  plants  included  several  entomo- 
philous species  and  the  grasses  timothy  (Phleum  pratense)  and 
orchard  grass  (Dactylis  glomerata). 

I sampled  Apoidea  by  walking  transects  and  capturing  bees  from 
flowers.  There  were  10,  30  m transects  randomly-placed  on  the  field. 
I took  transect  samples  during  times  of  maximum  foraging  activity 
(1000-1600  hours)  throughout  the  season  (at  least  3 samples  in  each 
2-week  period,  late  May-October,  1974  and  1975).  I used  all-day 
samples  from  randomly-selected  patches  of  common  flower  species 
(throughout  the  growing  season,  1975  and  1976)  to  confirm  the 
results  from  the  transect  samples  and  to  study  spatial  distributions 
of  foraging  bees.  Voucher  specimens  of  the  bee  species  are  placed  in 
the  Cornell  University  Insect  Collection,  lot  number  1039. 

I counted  the  number  of  flowers  of  each  species  at  anthesis  in  100, 
lm2  subquadrats.  The  subquadrats  were  arranged  in  groups  of  10, 
randomly-placed  within  30  m X 30  m quadrats  (the  bee  transects 
were  also  within  these  quadrats).  There  were  10  quadrats  randomly- 
placed  on  the  field.  Flowers  were  sampled  once  every  2 weeks 
throughout  the  season.  Voucher  specimens  of  the  plant  species  are 
placed  in  the  Bailey  Hortorium  Herbarium,  Cornell  University. 
Details  of  the  field  techniques  are  given  by  Ginsberg  (1979). 

I used  the  records  of  Fernald  (1950)  and  Wiegand  and  Eames 
(1925)  to  determine  whether  flower  species  were  native  or  were 
introduced  into  the  area.  Their  determinations  were  based  largely 
on  the  records  of  early  botanical  explorers  (e.g.  Pursh  1923)  and  on 
previous  species  lists  for  the  area  (e.g.  Dudley  1886).  Admittedly, 
there  is  some  margin  for  error  in  these  judgements,  but  because  of 
the  large  number  of  entomophilous  species  on  the  sample  site, 
mistakes  about  the  points  of  origin  of  a few  species  should  not 
influence  the  major  arguments. 

Results 

Red  maple  was  the  first  abundant  flower  species  to  bloom  on  the 
field  in  spring.  Several  willows  and  rosaceous  trees  (Prunus  cerasus, 


1981] 


Ginsberg  — Bee  Foraging  Patterns 


339 


Pyrus  malus ) bloomed  soon  after,  as  did  several  roadside  weeds 
such  as  dandelion  ( Taraxacum  officinale ) and  yellow  rocket  (Bar- 
barea  vulgaris).  The  spring  species  were  typically  clustered  in 
distribution  at  roadsides  and  forest  edges,  and  the  woody  species 
had  relatively  short  blooming  times.  Of  16  species  recorded  on  the 
field  in  spring  (late  April  and  early  May  in  1975)  half  were  native 
and  half  were  introduced.  I do  not  include  any  of  the  several  species 
that  bloomed  in  the  woods  nearby. 

Flower  bloom  increased  on  the  field  to  a maximum  in  early 
summer  (late  June,  early  July).  Most  of  the  species  in  bloom  at  this 
time  of  the  year  were  introduced  (Fig.  1).  Table  I lists  the  most 
common  of  these  species  and  gives  their  frequencies  of  occurrence  in 
the  subquadrats.  Note  that  the  most  common  flowers  at  this  time 
were  those  of  introduced  herbaceous  species.  Most  flowers  of  these 
species  were  past  blooming  by  midsummer. 

In  August,  goldenrods  ( Solidago  spp.)  predominated  on  the  field. 
These  late  summer  flowers  are  native  to  this  region  (Table  I).  Aster, 
another  native  genus  of  composites,  predominated  after  goldenrod 
passed  bloom  in  the  fall.  Late  season  flowers,  therefore,  were  mostly 
native  species  (Fig.  1). 


DATE 

Fig.  1.  Number  of  introduced  and  native  flower  species  blooming  over  the  summer, 
1974,  in  an  old  field  near  Ithaca,  New  York. 


340 


Psyche 


[Vol.  88 


Table  I.  Frequencies  of  common  flower  species,  1974 


Flower  species 

Origin1 

Time  of 
maximum 
bloom 

Frequency2 

No. 

inflores- 
cences/m2^ 

Hieracium  pratense 

I 

mid  June 

70 

12.55  ± 2.06 

Chrysanthemum 

leucanthemum 

j 

late  June 

54 

2.21  ±0.32 

Cornus  racemosa 

N 

late  June 

12 

2.10  ± 1.10 

Satureja  vulgaris 

N 

late  July 

30 

12.11  ± 3.08 

Achillea  millefolium 

I 

late  August 

20 

1.04  ±0.34 

Daucus  carota 

I 

late  August 

30 

0.87  ± 0.20 

Solidago  juncea 

N 

late  August 

60 

12.86  ±2.08 

S.  graminif olia 

N 

early  Sept. 

54 

9.16  ± 1.85 

S.  rugosa 

N 

early  Sept, 

44 

8.42  ± 2.03 

S.  altissima 

N 

early  Sept. 

62 

14.69  ± 3.05 

1 N = native  species;  I 

= introduced 

species 

2 Number  of  1 m2  subquadrats  (out  of  100)  in  which  species  was  flowering  during 


period  of  maximum  bloom. 

3 Mean  number  of  inflorescences  (sprays  for  Solidago ) per  subquadrat  during 
period  of  maximum  bloom  ± standard  error. 


This  flowering  trend  of  early-summer  introduced  species  and  late- 
summer-fall  native  species  probably  holds  for  central  New  York  as  a 
whole.  In  Figure  2 I plotted  the  number  of  open-habitat,  entomo- 
philous  species  blooming  in  the  entire  Cayuga  Lake  Basin  during 
each  2-week  period  over  the  season  (compiled  from  Wiegand  and 
Eames  1925).  Again,  introduced  species  predominate  in  early 
summer.  Later  in  the  summer,  native  and  introduced  species  are 
about  equal  in  number,  but  the  tremendous  abundance  of  goldenrod 
(Table  I;  also  Ginsberg  1979,  Hurlbert  1970)  results  in  a preponder- 
ance of  native  flowers  late  in  the  season. 

Foraging  phenologies  of  the  most  common  bee  species  indicate  a 
partitioning  of  the  season  according  to  foraging  times.  Native  wild 
bees  (mostly  primitively  social  halictines)  predominated  in  early 
summer,  while  Apis  mellifera  predominated  in  late  summer  (Table 
II).  This  presents  the  interesting  situation  that  native  bees  foraged 
primarily  on  introduced  flowers  in  early  summer,  while  the  intro- 
duced honey  bees  foraged  on  native  flowers  in  late  summer  and  fall 
(Table  III). 


341 


1981]  Ginsberg  — Bee  Foraging  Patterns 


M J J A S O 
DATE 


Fig.  2.  Number  of  introduced  and  native  flower  species  blooming  in  the  Cayuga 
Lake  Basin  (compiled  from  Wiegand  and  Eames  1925). 


Table  II.  Percent  of  honey  bees  in  transect  samples,  1974 


Period 

Dates 

% honey  bees1 

N 

1 

22  May-4  June 

2.1 

48 

2 

5 June-18  June 

1.9 

52 

3 

19  June-2  July 

9.0 

67 

4 

3 July- 16  July 

7.6 

79 

5 

17  July-30  July 

13.8 

29 

6 

31  July-13  August 

15.9 

44 

7 

14  Aug-27  August 

79.8 

119 

8 

28  Aug-10  September 

95.2 

230 

9 

11  Sept. -24  Sept. 

89.7 

78 

Percent  of  bees  captured  in  transect  samples  that  were  Apis  mellifera.  Other 
bees  in  these  samples  were  native  wild  bees  (except  for  3 individuals  of  Andrew 
willcella  captured  on  28  May,  12  June,  and  8 July — this  species  was  probably 
introduced  into  the  region). 


342 


Psyche 


[Vol.  88 


Table  III.  Flower  species  most  commonly  visited  by  bees  during  the  summer,  1974 


Bee  species1 

Flower  species1 

% of  visits2 

sample  size3 

Apis  mellifera  (I) 

Solidago  altissima  (N) 

26.2 

409 

S.  graminifolia  (N) 

24.0 

S.  juncea  (N) 

18.8 

Ceratina 4 (N) 

Rubus  allegheniensis  (N) 

28.0 

50 

Halictus  ligatus  (N) 

Chrysanthemum  leucan- 

themum  (I) 

51.2 

43 

Halictus  confusus  (N) 

Potentilla  recta  (I) 

47.6 

21 

Auguchlorella  striata  (N) 

Chrysanthemum  leucan- 

themum  (I) 

37.5 

24 

Dialictus  rohweri  (N) 

Potentilla  recta  (I) 

47.4 

19 

1 Point  of  origin  given  in  parentheses;  N = native  to  North  America;  I = introduced. 

2 Percent  of  individuals  of  that  bee  species  in  samples  that  were  on  named 
flower  species. 

3 Number  of  bees  of  that  species  in  transect  samples,  1974. 

4 Includes  Ceratina  dupla  and  C.  calcarata.  Females  of  these  species  are  indis- 
tinguishable at  present. 


Spring-flying  bees  were  not  included  in  Table  II  because  they 
foraged  on  flowers  that  were  most  common  off  the  field  and  could 
not  be  sampled  by  the  transect  technique.  All-day  samples  from 
patches  of  common  spring  flowers  revealed  a great  diversity  of 
native  bees,  primarily  solitary,  univoltine  species  of  Andrena, 
Dialictus,  and  Ceratina.  Honey  bees  were  also  common  in  spring, 
especially  on  willows,  rosaceous  trees,  and  on  large  clusters  of 
dandelion  and  yellow  rocket. 

Discussion 

The  fact  that  native  bees  foraged  on  introduced  flowers  in  early 
summer,  while  introduced  bees  predominated  on  native  flowers  in 
late  summer,  suggests  that  this  type  of  old  field  association  is  quite 
recent  in  origin.  Indeed,  the  development  of  this  curious  pattern  can 
be  clarified  by  tracing  the  recent  biotic  history  of  the  Ithaca  area. 

Early  explorers  in  the  region  (up  until  the  early  1800’s)  reported 
extensive  forested  areas  that  were  thickest  near  the  head  of  Cayuga 
Lake  and  to  the  south  of  Ithaca  (Dudley  1886).  The  Indians  cleared 
considerable  acreages  for  villages,  corn  fields,  etc.  (Day  1953)  and 
kept  corridors  of  land  clear  for  stalking  deer  by  annual  burning 


1981] 


Ginsberg  — Bee  Foraging  Patterns 


343 


(Dudley  1886).  These  cleared  areas  were  probably  far  less  extensive 
than  present-day  open  habitats.  Also,  the  deer-stalking  grounds 
differed  from  modern  old  fields  because  they  were  burned  each  year, 
and  because  they  lacked  many  of  the  introduced  flower  species  that 
are  now  common.  Some  of  these  species  were  introduced  by  1807, 
when  the  explorer  Frederick  Pursh  passed  through  Ithaca  (Dudley 
1886;  Pursh  1923). 

The  first  settlers  arrived  in  Ithaca  about  1789  (Dudley  1886).  By 
the  mid  1800’s  extensive  areas  of  land  had  been  cleared  for  farming 
and  settlements.  Total  acreage  used  for  farming  reached  a peak  in 
New  York  State  (approximately  23,780,754  acres)  about  1880.  Since 
then,  gradual  abandonment  of  farmland  has  given  rise  to  many 
abandoned  fields.  By  1925,  only  19,269,926  acres  of  farmland 
remained  (Vaughan  1929).  By  the  late  1960’s  the  area  of  crop  and 
pasture  land  in  New  York  State  totalled  only  about  8,771,800  acres 
(Ferguson  and  Mayer  1970).  Much  of  this  farm  land  was  lost  to 
villages  and  cities,  but  a considerable  amount  was  left  as  abandoned 
fields.  In  the  late  1800’s  and  early  1900’s  several  weedy  species  were 
introduced,  and  many  others  increased  in  abundance  in  central  New 
York.  Among  the  species  that  became  common  at  this  time  were 
Hieracium  pratense  and  Potentilla  recta  (Wiegand  and  Eames 
1925),  both  important  species  at  my  sample  site  (Tables  I and  III). 
Taken  together,  these  facts  suggest  that  the  current  floral  composi- 
tion of  old-field  communities  in  central  New  York  is  on  the  order  of 
100  years  old. 

As  a result  of  these  changes  in  the  local  flora,  at  least  three  new 
classes  of  abundant  flower  forage  have  become  available  to  bees.  In 
spring,  the  introduced  rosaceous  trees  and  roadside  weeds  provide 
considerable  forage.  Second,  the  increased  acreage  of  abandoned 
fields,  along  with  introductions  of  several  plant  species,  results  in  an 
historically  novel  flower  bloom  in  early  summer.  Finally,  the  large 
acreage  of  open  fields  results  in  an  unprecedented  profuse  bloom  of 
goldenrod  in  late  summer. 

The  honey  bee  was  introduced  into  North  America  by  the  early 
colonists  (Crane  1975).  The  Italian  strain  {Apis  mellifera  ligustica), 
which  now  predominates  in  New  York  State,  was  not  introduced 
until  1859  (Ruttner  1975).  Some  more  recent  introductions  into  the 
Ithaca  area  include  the  megachilids  Megachile  rotundata  (Mitchell 
1962)  and  Anthidium  manicatum  (Pechuman  1967),  and  the  andre- 


344 


Psyche 


[Vol.  88 


nid  Andrena  wilkella  (Linsley  1958).  At  my  study  site,  the  honey  bee 
is  far  the  most  abundant  of  these  species  (Ginsberg  1979).  In  the 
1950’s,  honey  bee  populations  declined  sharply  in  New  York  State 
due  to  the  increased  use  of  pesticides  and  the  decline  in  farm  acreage 
devoted  to  buckwheat,  an  important  food  source  for  honey  bees 
(Morse  1975).  Before  1950,  therefore,  honey  bees  were  even  more 
common  than  at  present. 

Apis  mellifera  is  a high-density  specialist  in  flower  foraging.  Its 
large  colony  size  and  recruitment  capabilities  facilitate  this  special- 
ization (Eickwort  and  Ginsberg  1980;  Sakagami  1959).  In  spring, 
honey  bees  forage  on  high-density  resources  such  as  rosaceous  trees, 
willows,  and  clusters  of  roadside  herbs.  In  late  summer,  honey  bees 
forage  on  the  super-abundant  goldenrods,  also  high-density  re- 
sources. 

In  early  summer,  honey  bees  are  relatively  rare  on  the  old  field 
(Table  II).  At  this  time  of  season  they  forage  primarily  off  the  field 
on  high-density  resource  species  in  forests  and  on  cultivated  fields 
(Farrar  1944;  Ginsberg  1979).  The  introduced  herbs  that  bloom  at 
this  time  are  exploited  by  primitively  social  halictines  (Table  III). 
The  multivoltine  seasonal  cycles  of  these  bees  allow  them  to  build 
up  their  populations  over  the  season,  thus  they  can  exploit  the 
recently  introduced  flower  species  that  are  now  abundant  in  early 
summer.  Ceratina,  which  is  probably  univoltine  in  the  Ithaca  area, 
is  also  common  in  early  summer,  but  it  forages  somewhat  earlier 
than  the  halictine  bees,  and  is  most  common  on  native  flowers  such 
as  Rubus  spp.  (Table  III). 

An  interesting  result  of  this  analysis  is  that  each  of  the  major 
historically  novel  instances  of  resource  abundance  is  exploited  by 
social  bees.  Honey  bees  forage  on  rosaceous  trees  and  roadside 
weeds  in  spring,  and  on  goldenrods  in  late  summer.  Native  bees 
forage  on  these  flowers  also,  but  honey  bees  predominate  because  of 
their  high  populations  and  recruitment  ability,  both  features  related 
to  their  social  behavior.  Social  halictines  predominate  on  intro- 
duced herbs  in  early  summer  because  of  their  broad  host  ranges  and 
their  multivoltine  seasonal  cycles,  also  related  to  their  sociality.  Ap- 
parently, the  ability  to  adapt  to  landscape-level  changes  in  resource 
availability  is  an  important  advantage  that  accompanies  social 
behavior  in  bees.  This  does  not  mean  that  only  social  insect  species 
can  adapt  rapidly  to  changes  in  resource  levels.  It  does  suggest  that 
in  bees,  sociality  facilitates  this  rapid  adaptability. 


1981] 


Ginsberg  — Bee  Foraging  Patterns 


345 


Conclusions 

In  an  abandoned  field  in  central  New  York  State,  native  bees 
foraged  predominantly  on  introduced  flower  species  in  early  sum- 
mer, while  the  introduced  honey  bee  predominated  on  native 
goldenrods  in  late  summer.  This  situation  results  from  recent 
changes  in  the  flora  and  fauna  of  the  region. 

The  activities  of  European  settlers  have  caused  large-scale  changes 
in  the  flora  of  the  northeastern  United  States.  These  changes  result 
primarily  from  introductions  of  alien  species,  and  from  clearing  of 
land  for  farming  with  subsequent  abandonment.  At  present,  there 
are  at  least  three  instances  of  profuse  flowering  over  the  season  that 
are  historically  novel  to  this  area.  These  are  the  abundant  bloom  of 
introduced  trees  and  roadside  weeds  in  spring,  the  flowering  peak  of 
introduced  weeds  in  early  summer,  and  the  profuse  flowering  of 
native  goldenrods  in  late  summer.  In  all  three  of  these  cases,  the 
predominant  foragers  are  social  bees;  honey  bees  in  spring  and  late 
summer,  and  social  halictines  in  early  summer.  The  ability  of  these 
bees  to  exploit  historically  novel  pulses  of  flowering  results  from 
features  related  to  their  social  behavior;  large  colony  size  and 
recruitment  ability  in  Apis  mellifera,  and  the  multivoltine  seasonal 
cycle  in  the  social  halictines. 

Acknowledgments 

I thank  G.  C.  Eickwort,  F.  C.  Evans,  R.  A.  Morse,  and  R. 
Nowogrodzki  for  constructive  comments  on  early  drafts  of  the 
manuscript.  F.  J.  Rohlf  provided  useful  advice.  I also  thank  W. 
Denison  and  family  for  providing  their  land  as  a study  site.  This 
study  was  conducted  in  partial  fulfillment  of  the  requirements  for 
the  Doctor  of  Philosophy  at  Cornell  University.  Funding  was 
provided,  in  part,  by  NSF  grant  no.  BMS-72-02386  to  G.  C. 
Eickwort. 


Literature  Cited 


Crane,  E. 

1975.  The  world’s  beekeeping — past  and  present.  In  Dadant  and  sons  (eds.) 
The  Hive  and  the  Honeybee.  Hamilton,  111.:  Dadant  and  sons,  pp.  19-38. 

Day,  G.  M. 

1953.  The  Indian  as  an  ecological  factor  in  the  northeastern  forest.  Ecology 
34:329-346. 


346 


Psyche 


[Vol.  88 


Dudley,  W.  R. 

1886.  The  Cayuga  flora.  Bull.  Cornell  Univ.  (Science)  2:1-132. 

Eickwort,  G.  C.  and  H.  S.  Ginsberg 

1980.  Foraging  and  mating  behavior  in  Apoidea.  Ann.  Rev.  Entomol.  25:421- 
446. 

Farrar,  C.  L. 

1944.  Productive  management  of  honeybee  colonies  in  the  northern  states. 
USDA  circ.  No.  702.  28  pp. 

Ferguson,  R.  H.  and  C.  E.  Mayer 

1970.  The  timber  resources  of  New  York.  USDA  Forest  Service  Resource 
Bull.  NE-20,  193  pp. 

Fernald,  M.  L. 

1950.  Gray’s  Manual  of  Botany.  8th  ed.  New  York:  D.  Van  Nostrand.  1632  pp. 
Ginsberg,  H.  S. 

1979.  Foraging  ecology  of  pollen  utilizing  insects  on  an  old  field  in  central  New 
York  State.  Ph.D.  Thesis,  Cornell  Univ.  Ithaca,  N.Y.  221  pp. 
Hurlbert,  S.  H. 

1970.  Flower  number,  flowering  time,  and  reproductive  isolation  among  ten 
species  of  Solidago  (Compositae).  Bull.  Torrey  Bot.  Club  97:189-195. 
Linsley,  E.  G. 

1958.  The  ecology  of  solitary  bees.  Hilgardia  27:543-599. 

Mitchell,  T.  B. 

1962.  Bees  of  the  eastern  United  States,  Vol.  II.  N.  Carolina  Agr.  Exp.  Sta., 
Tech.  Bull.  No.  152.  557  pp. 

Morse,  R.  A. 

1975.  Bees  and  Beekeeping.  Ithaca,  N.Y.:  Cornell  Univ.  Press.  295  pp. 
Pearson,  J.  F.  W. 

1933.  Studies  on  the  ecological  relations  of  bees  in  the  Chicago  region.  Ecol. 
Monogr.  3:375-441. 

Pechuman,  L.  L. 

1967.  Observations  on  the  behavior  of  the  bee  Anthidium  manicatum  (L.).  J. 
N.  Y.  Entomol.  Soc.  2:68-73. 

PURSH,  F. 

1923.  Journal  of  a botanical  excursion  in  the  northeastern  parts  of  the  states  of 
Pennsylvania  and  New  York  during  the  year  1807.  Dehler  Press.  1 13  pp. 
Ruttner,  F. 

1975.  Races  of  bees.  In  Dadant  and  sons  (eds.)  The  Hive  and  the  Honeybee. 
Hamilton,  111.:  Dadant  and  sons,  pp.  19-38. 

Sakagami,  S.  F. 

1959.  Some  interspecific  relations  between  Japanese  and  European  honeybees. 
J.  Animal  Ecol.  28:51-68. 

Vaughan,  L.  M. 

1929.  Abandoned  farm  areas  in  New  York.  Cornell  Univ.  Agr.  Exp.  Stat.  Bull. 
490:1-285. 

WlEGAND,  K.  M.  AND  A.  J.  EAMES 

1925.  The  flora  of  the  Cayuga  Lake  Basin.  Cornell  Univ.  Agr.  Exp.  Sta.  Mem. 
92:1-491. 


MYRMECOPHILIC  RELATIONSHIP  OF  PELLA 
(COLEOPTERA:  STAPHYLINIDAE)  TO 
LASIUS  FULIGINOS US  (HYMENOPTERA:  FORMICIDAE) 


By  B.  Holldolber*,  M.  Moglich**,  U.  Maschwitz*** 
Introduction 

A large  number  of  staphylinid  beetles  are  closely  associated  with 
ants  and  termites  (for  review  see  Wilson  1971,  Kistner  1979).  Those 
species  living  with  ants  are  commonly  called  myremcophiles.  At 
least  a few  ( Atemeles , Lomechusa ) have  “broken”  the  communica- 
tion code  of  their  host  species  and  are  thereby  able  to  become 
completely  integrated  in  the  social  system  of  the  ants  (Holldobler 
1967,  1970,  1971).  In  an  attempt  to  understand  the  evolutionary 
pathways  of  this  highly  specialized  social  parasitic  behavior,  we 
studied  closely  related  staphylinid  species  that  do  not  live  within  the 
ant  society  but  instead  occupy  the  foraging  trails  and  garbage 
dumps  of  an  ant  nest. 

Many  of  such  myrmecophilous  staphylinids  can  be  found  with  the 
formicine  ant  Lasius  fuliginosus  and  most  of  them  belong  to  the 
genus  Pella.  Apparently  these  beetles  are  not  endowed  with  the 
behavioral  repertory  that  would  enable  them  to  live  within  the  ant 
colony,  although  they  seem  to  have  a close  ecological  association 
with  ants  (Holldobler  1972). 

Kistner  (1971)  redefined  the  genus  Zyras  and  raised  the  former 
subgenus  Pella  to  generic  rank.  The  first  behavioral  observations 
concerning  Pella  ( =Zyras , Myrmedonia ) were  published  by  Was- 
mann  (1886,  1930).  He  stated  that  these  beetles  feed  on  dead  or 
disabled  ants,  but  that  they  also  lie  in  wait  near  the  entrance  and 
hunt  ants  returning  to  the  nest.  Furthermore,  Wasmann  pointed  out 
that  because  of  their  generalized  and  primitive  structure  these 
beetles  can  be  regarded  as  close  to  the  ancestral  forms  from  which 
some  of  the  more  specialized  staphylinid  myrmecophiles  were 
derived. 


♦Department  of  Biology,  Harvard  University,  Cambridge,  Mass.,  USA. 
♦♦Present  address:  Am  Lowentor  15,  Darmstadt,  W. -Germany. 
♦♦♦Fachbereich  Biologie  (Zoologie)  der  Universitat  Frankfurt,  W. -Germany. 
Manuscript  received  by  the  editor  October  19,  1981. 


347 


348 


Psyche 


[Vol.  88 


Material  and  Methods 


At  our  major  study  sites  near  Ochsenfurt,  Riederau  (both  Bavaria, 
W. -Germany)  and  Gravenbruch  (Hessen,  W. -Germany)  we  found 
12  staphylinid  species  associated  with  Lasius  fuliginosus  (Tab.  1). 
Our  investigations  concentrated  on  the  genus  Pella  (mainly  P. 
funesta,  P.  laticollis  and  P.  cognata ).  Since  the  myrmecophilous 
behavior  of  these  species  was  found  to  be  very  similar  we  will  refer 
to  the  individual  species  only  where  necessary.  In  fact,  when 
observing  the  beetles  in  the  field  it  was  usually  not  possible  to 
identify  the  species  precisely.  We  made  additional  observations  with 
Pella  humeralis,  which  can  be  found  with  L.  fuliginosus,  but  which 
often  also  occurs  near  the  nests  of  Formica  polyctena  (Wasmann 
1920;  Kolbe  1971). 

The  field  observations  were  conducted  throughout  the  years 
1967-1969,  and  sporadically  in  1970-1973.  In  an  attempt  to  follow 
the  life  cycle  of  the  beetles  in  the  laboratory  we  set  up  a large  colony 
of  Lasius  fuliginosus  in  a laboratory  nest.  The  culture  and  mainte- 
nance of  these  ants  over  a longer  period  of  time  was  particularly 
difficult,  because  L.  fuliginosus  constructs  carton  nests  with  the  aid 
of  a special  symbiontic  fungus  ( Cladosporium  myrmecophilum) . A 
detailed  description  of  the  nest  building  behavior  of  L.  fuliginosus 
and  of  the  laboratory  nest  is  given  in  Maschwitz  and  Holldobler 
(1970). 

In  order  to  measure  quantitatively  possible  trophallactic  feeding 
of  the  myrmecophiles  by  their  host  ants,  tracer  experiments  were 
carried  out  using  the  radioisotope  32P  mixed  with  honey-water.  The 
quantity  of  marked  food  taken  up  by  the  ants  was  reflected  in  the 
counts  per  minute  which  were  determined  with  a standard  Geiger- 


Table  l 


Staphylinids  found  near  one  nest  of  Lasius  fuliginosus. 


Pella  laticollis 
Pella  lugens 
Pella  cognata 
Pella  funesta 


Pella  humeralis 


Oxypoda  vittata 
Rugilus  rufipes 
Thiasophila  inquilina 
Homoeusa  acuminata 
Sipalia  circellaris 
A theta  fungi 
A theta  sodalis 


1981] 


Holldobler,  Moglich,  & Maschwitz  — Pella 


349 


Muller  counter  combined  with  an  automatic  sample  changer 
(Philips).  For  further  information  concerning  the  tracer  techniques 
applied  in  this  study  see  Gosswald  and  Kloft  1958;  Kloft  1959. 

For  histological  investigations  live  specimens  were  fixed  in  alco- 
holic Bouin  (Dubosq  Brasil)  or  Carnoy  (Romeis  1948),  embedded  in 
Methyl  Methacrylate,  and  sectioned  5-8^  thick  with  a Jung  Tet- 
rander  microtome  (Rathmayer  1962).  The  staining  was  Azan 
(Heidenhain). 

For  the  chemical  analysis  of  the  defensive  secretions  of  Pella , 
liquid  material  was  collected  with  glass  capillaries  from  the  dissected 
glandular  reservoirs  or  washed  with  water  from  the  surface  of  the 
irritated  beetles.  The  quinones  were  identified  by  thinlayer  chroma- 
tography as  2,4  - dinitrophenylhydrazine  in  2 N hydrochloric  acid  or 
by  reduction  with  sulphurous  acid.  The  dinitrophenylhydrazones 
were  separated  on  alumina  F 254  (Merck)  with  chloroform  meth- 
anol (19:1)  as  mobile  phase  and  on  silica  gel  F 254  (Merck)  with 
benzene-ethyl  acetate  (4:1)  as  mobile  phase  (Moore  1968).  The 
hydroquinones  were  separated  on  silica  gel  F254  with  benzene 
dioxane  (3:1)  as  mobile  phase  and  then  sprayed  with  a solution  of 
0.5%  hydrogen  peroxide  and  a solution  of  peroxidase.  The  newly 
formed  quinones  were  made  visible  by  spraying  with  DNP  and 
treating  with  ammonia  vapour  (Schildknecht  and  Kramer  1962). 
Hydrocarbons,  terpenes  and  carbonic  acids  were  analyzed  by  GLC. 
We  used  a Perkin  Elmer  chromatograph,  model  300,  equipped  with 
a flame  ionization  detector.  Columns:  1.8m  X 2.7mm  stainless  steel, 
packed  with  a)  4%  polypropylene  glucol  on  Chromosorb  G (100°C 
column  temperature);  b)  4%  polyethyleneglycol  1500  on  Chromo- 
sorb G (70°  C);  c)  25%  diethylhexyl  sebacinate  plus  sebacinic  acid 
on  Kieselgur  60-100  (140°  C)  (30  ml  N2/min;  FID). 

Results 

The  life  cycle  of  Pella  funesta 

The  following  description  of  the  life  cycle  of  Pella  funesta  is  based 
on  field  observations  and  on  data  obtained  from  laboratory  cul- 
tures. Pella  laticollis  appears  to  have  a similar  life  cycle,  but  our 
observational  data  are  not  as  complete  as  for  P.  funesta. 

In  late  March  and  early  April  a large  number  of  P.  funesta  beetles 
were  typically  found  in  the  excavation  material  on  the  base  of  the 
trunks  of  L.  fuliginosus  nest  trees.  At  this  time  most  of  the  beetles 


350 


Psyche 


[Vol.  88 


were  lying  motionless  in  the  loose  material  and  showed  a kind  of 
“dormance  posture”:  the  abdomen  was  bent  over  its  back,  with  the 
legs  and  antennae  folded  tightly  to  the  body.  On  warmer  days, 
however,  the  beetles  exhibited  high  locomotory  and  flight  activity, 
and  in  the  laboratory  they  showed  a strong  positive  phototaxis. 
During  this  period  we  frequently  observed  beetles  copulating  in  the 
laboratory  nests.  Toward  the  end  of  April  the  sexual  behavior  and 
flight  activities  ceased.  In  the  laboratory  as  well  as  in  the  field  the 
beetles  were  now  active  primarily  during  the  night,  while  during  the 
daytime  they  clustered  under  shelters  near  the  Lasius  fuliginosus 
nest.  Only  occasionally  were  we  able  to  spot  a beetle  outside  the 
shelters  at  daytime. 

Also  near  the  end  of  April  we  found  the  first  beetle  eggs  in  the 
“garbage  dumps”  of  the  laboratory  nest  of  L.  fuliginosus , and  by 
early  May  the  first  Pella  larvae  had  hatched.  The  larvae  developed 
quite  rapidly,  so  that  in  mid-May  we  found  the  first  pupae  in  the 
“garbage  dumps”  of  the  ant  nests,  even  though  larvae  could  still  be 
found  throughout  the  months  of  June  and  July.  In  June  the 
mortality  of  adult  beetles  in  our  laboratory  nest  increased  markedly 
and  in  late  July  and  August  the  first  young  beetles  eclosed  from 
their  pupae.  These  beetles,  as  well  as  those  collected  in  the  field  in 
early  August,  exhibited  strong  positive  phototaxis  and  high  flight 
activity  for  a few  days.  After  this  short  period,  however,  the  beetles 
were  primarily  active  at  night  and  during  the  day  they  stayed  in 
shelters.  Finally,  in  October,  the  number  of  beetles  found  outside 
the  ant  nest  declined  markedly  and  by  November  no  more  beetles 
could  be  found  outside  the  nest.  In  December  we  excavated  to  L. 
fuliginosus  nests.  In  both  nests  we  found  several  Pella  beetles  in 
“dormance  position”  covered  by  loose  nest  material  of  the  peripher- 
al nest  chambers  and  on  the  ground  inside  the  nest  tree  trunk. 
Presumably  these  beetles  were  overwintering  within  the  Lasius 
fuliginosus  nest  until  their  activity  period  would  start  again  in  early 
spring  the  coming  year. 

From  these  observations  we  propose  the  following  life  cycle  for 
Pella  funesta : in  early  spring  the  adult  beetles  deposit  eggs  near  the 
ants’  “garbage  dump”  area.  The  larvae  develop  in  the  “garbage 
dump”,  pupate  during  the  period  from  May  to  July  and  between 
July-August  the  adult  beetles  eclose.  After  eclosion  the  young 
beetles  apparently  migrate,  as  indicated  by  the  short  period  of  high 
diurnal  locomotory  and  flight  activity.  After  this  period  the  beetles 


1981] 


Holldobler,  Moglich,  & Maschwitz  — Pella 


351 


forage  near  the  L.  fuliginosus  nest  during  the  night  and  stay  in 
shelters  during  the  day.  They  overwinter  in  dormancy  inside  the  L. 
fuliginosus  nest.  With  the  end  of  winter  the  beetles  enter  a second 
diurnal  activity  phase  during  which  mating  takes  place.  After  repro- 
duction the  beetles  die,  normally  a few  weeks  before  the  new  beetle 
generation  ecloses  in  June. 

The  behavior  of  the  larvae  of  Pella 

The  description  of  the  behavior  of  the  larvae  is  primarily  based  on 
observations  in  the  laboratory.  In  the  field  and  in  the  laboratory 
nest,  the  larvae  were  almost  exclusively  found  near  or  in  the 
“garbage  dumps”  of  the  L.  fuliginosus  nests.  We  frequently  ob- 
served the  larvae  feeding  on  dead  ants  (Fig.  IB).  During  feeding  the 
larvae  always  attempted  to  stay  “out  of  sight”  either  by  remaining 
entirely  beneath  the  booty  and  devouring  it  from  below  or  by 
crawling  inside  the  dead  ant’s  body.  Occasionally  2-4  larvae  could 
be  observed  feeding  on  the  same  ant  cadaver.  But  when  they  became 
too  crowded  they  frequently  attacked  each  other,  sometimes  result- 
ing in  one  larvae  eating  the  other  (Fig.  1C). 

When  ants  encountered  the  larvae  they  usually  attacked  them. 
Almost  invariably  the  larvae  reacted  with  a typical  defense  beha- 
vior. They  raised  their  abdominal  tip  towards  the  head  of  the  ants. 
Usually  the  ant  responded  by  stopping  the  attack  and  palpating  the 
larva’s  tip  (Fig.  1A).  In  most  cases  this  short  interruption  was 
enough  to  allow  the  larvae  to  escape.  We  observed  hundreds  of  such 
encounters  between  ants  and  Pella  larvae;  only  a few  ended  fatally 
for  the  larvae. 

Histological  invesigations  revealed  that  the  Pella  larvae  have  a 
complex  dorsal  glandular  structure  with  a reservoir  near  the 
abdominal  tip  in  the  second  last  segment.  In  addition  we  found 
single  cell  glands  positioned  dorso-laterally  in  each  segment.  Similar 
glandular  structures  had  previously  been  found  in  larvae  of  the 
myrmecophile  staphylinids  Atemeles  and  Lomechusa,  and  circum- 
stantial evidence  strongly  indicated  that  in  these  species  the  glands 
produce  so-called  pseudopheromones  which  release  adoption  beha- 
vior in  the  host  ants  (Holldobler  1967).  We  have  no  evidence  to 
suggest  that  these  glands  have  a similar  function  in  Pella.  However, 
it  is  possible  that  the  more  complex  glandular  structure  at  the 
abdominal  tip,  produces  an  appeasement  secretion  by  which  the 
aggressiveness  of  attacking  ants  can  be  briefly  blunted. 


352 


Psyche 


[Vol.  88 


Fig.  1 Behavior  of  the  larvae  of  Pella.  A)  Larva  presenting  abdominal  tip  to  an 
attacking  worker  of  Lasius  fuliginosus.  The  ant  interrupts  attack  and  licks  at  the 
larva.  B)  Larva  feeds  on  dead  ant.  C)  Cannibalistic  behavior  of  Pella  larvae. 


1981] 


Holldobler,  Moglich,  & Maschwitz  — Pella 


353 


In  any  case,  Pella  larvae  are  able  to  come  into  close  contact  to 
workers  of  L.  fuliginosus  without  being  attacked,  especially  when 
the  temperature  is  low  (14-17°  C).  Under  those  circumstances  we 
have  seen  the  beetle  larvae  licking  the  cuticle  of  live  ants,  including 
even  the  mandibles  and  mouthparts.  This  led  us  to  the  question  of 
whether  the  beetle  larvae  occasionally  solicit  regurgitation  in  ants. 
In  order  to  investigate  this  possibility  ants  were  fed  with  honey- 
water  labeled  with  the  isotope  32P  and  then  housed  together  with 
beetle  larvae.  For  each  sample  we  kept  30  radioactive  ants  with  5 
beetle  larvae  in  plastic  containers  (10  X 15  cm)  with  a moist  gypsum 
bottom.  One  experimental  series  was  conducted  in  a temperature 
range  of  14.5-16.5°  C,  the  other  in  20-23°  C.  After  24  hours  we 
measured  the  amount  of  radioactivity  in  each  individual  ant  and 
larva.  No  significant  amount  of  radioactivity  had  been  transferred 
from  the  ants  to  the  larvae,  except  in  container  6,  where  one  ant  was 
found  dead  and  obviously  partly  eaten  by  the  larvae.  Since  the 
amount  of  radioactivity  carried  by  some  of  the  larvae  was  only  very 
slightly  above  the  background  activity,  we  concluded  that  it  was 
transferred  by  contamination.  From  this  experiment  it  appears  that 
the  Pella  larvae  do  not  solicit  regurgitation  in  ants.  Their  main  food 
source  seems  to  be  dead  ants  or  debris  of  the  ants.  In  fact,  they  can 
easily  be  raised  by  keeping  them  entirely  separated  from  living  ants, 
just  by  feeding  them  regularly  with  dead  ants. 

Predatory  behavior  of  adult  beetles 
Since  Wasmann’s  early  observations  (1886,  1920,  1925)  very  little 
has  been  reported  concerning  the  biology  of  the  myrmecophilous 
Pella.  Wasmann  reported  that  all  species  he  had  studied  (P. 
humeralis,  P.  funesta,  P.  cognata,  P.  similis,  P.  lugens  and  P. 
laticollis ) live  with  Lasius  fuliginosus , and  only  P.  humeralis  can 
also  be  found  with  species  of  the  Formica  rufa  group.  According  to 
Wasmann  all  these  Pella  species  prey  on  ants,  concentrating  espe- 
cially on  disabled  ants.  In  addition  Wasmann  observed  that  the 
beetles  are  active  primarily  during  the  night.  In  a more  recent 
publication  Kolbe  (1971)  failed  to  find  a predatorial  behavior  in  P. 
humeralis  and  concluded  that  this  species  primarily  feeds  on  dead 
ants.  Similar  observations  were  made  with  Pella  japonicus,  which 
lives  with  Lasius  spathepus  (=  L.  fuliginosus  var.  spathepus 
Wheeler)  (Yasumatsu  1937;  Kistner  1971).  Kistner  also  observed 
that  these  beetles  “ate  small  insects  that  are  being  transported  by  the 


354 


Psyche 


[Vol.  88 


ants”.  However,  he  could  not  “see  the  Pella  eating  live  ants  or 
fighting  any  of  the  ants  on  the  trail”. 

Our  observations  of  Pella  funesta,  P.  laticollis  and  P.  humeralis 
confirmed  that  these  species  live  as  scavengers,  feeding  on  dead  or 
disabled  ants  and  debris  discarded  by  the  ants.  However,  we  also 
observed  these  beetles  acting  as  very  effective  predators  on  the  ants. 
Most  of  the  following  studies  were  made  with  P.  laticollis  and  P. 
funesta. 

During  the  main  foraging  season  from  May  to  October  Lasius 
fuliginosus  is  active  day  and  night.  Foragers  travel  along  well 
established  trunk  trails  to  feeding  sites  which  are  sometimes  more 
than  40  m distant.  At  daytime  we  only  occasionally  saw  Pella  beetles 
moving  along  or  nearby  the  trail.  However,  when  we  watched  the 
trunk  trails  with  a flash  light  at  night  many  Pella  were  seen  running 
along  the  ants’  foraging  routes.  Although  most  beetles  were  found 
within  a range  of  5 m from  the  nest  tree  of  L.  fuliginosus , we  also 
found  beetles  on  the  trunk  trail  as  far  as  22  m away  from  the  nest. 

On  6 different  occasions  we  witnessed  Pella  beetles  hunting  L. 
fuliginosus  workers  at  night  on  the  foraging  trail.  When  an  ant  was 
killed  it  was  dragged  a few  centimeters  away  from  the  trail  and  eaten 
under  a shelter,  sometimes  by  several  beetles  simultaneously. 

More  detailed  observations  on  the  behavioral  interactions  of 
Pella  and  L.  fuliginosus  were  made  in  the  laboratory.  As  long  as 
enough  dead  ants  were  available  at  the  ants’  nest  midden,  the  beetles 
showed  no  predatory  behavior  at  all,  limiting  themselves  to  a diet  of 
ant  cadavers  (Fig.  2A).  But  when  the  beetles  were  starved  for  a few 
days  and  then  placed  together  with  ants  in  an  observation  arena,  the 
predation  by  Pella  became  strikingly  prominent,  although  the  time 
of  onset  was  often  very  unpredictable.  We  saw  the  beetles  hunting 
during  the  daytime,  but  we  observed  such  activity  most  frequently  in 
the  evening  or  at  night.  The  beetles  chased  after  individual  ants  and 
pursued  them  through  approximately  2-6  cm  (very  rarely  through 
longer  distances  than  that).  When  the  beetle  moved  directly  behind 
the  ant  it  attempted  to  mount  it  and  insert  its  head  between  the  ant’s 
head  and  thorax.  When  attacked  the  ant  usually  reacted  by 
suddenly  stopping  and  pressing  the  femur  rapidly  and  tightly  to  its 
body  (Fig.  5).  Often  this  reaction  threw  the  beetle  off  the  back  of  the 
ant,  allowing  the  ant  to  escape.  In  one  series  of  observations  we 
counted  178  beetle  onslaughts  on  L.  fuliginosus  workers  within  a 
period  of  3 hours;  of  these,  only  9 attempts  (5%)  were  successful. 


1981] 


Holldobler,  Moglich,  & Maschwitz  — Pella 


355 


Fig.  2 A)  Pella  beetle  feeding  on  dead  L.  fuliginosus  worker.  Frequently  the 
beetles  lick  first  the  mouth  parts  of  the  ants,  before  tearing  the  cadavers  apart.  They 
might  be  attracted  to  mouthparts  by  sweet  remainders  of  honeydew.  B)  Cluster  of 
Pella  beetles  around  a prey  object. 


356 


Psyche 


[Vol.  88 


Fig.  3 Photograph  and  drawing  of  Pella  attacking  live  worker  of  L.  fuliginosus. 


1981] 


Holldobler,  Moglich,  & Maschxvitz  — Pella 


357 


The  hunting  behavior  of  the  beetles  was  always  the  same  in  P. 
laticollis,  P.  funesta  and  P.  humeralis : the  beetle  attacked  from 
behind  and  always  attempted  to  insert  its  head  between  the  head 
and  thorax  of  the  ant.  We  inspected  several  ants  which  had  just  been 
immobilized  by  an  attack  of  a beetle  and  found  that  in  most  cases 
the  pronotum  was  widely  separated  from  the  head  and  usually  the 
oesophagus  and  connectives  of  the  nervous  system  were  cut. 

Occasionally  we  observed  2-3  beetles  chasing  behind  one  ant 
(Fig.  4).  Once  the  ant  was  caught  by  a beetle  the  other  beetles  joined 
in  subduing  and  killing  the  ant.  Although  individual  beetles  often 
tried  to  drag  the  prey  away  from  the  rest  of  the  “hunting  pack”, 
usually  several  beetles  fed  on  the  prey  simultaneously.  No  aggres- 
sion among  the  beetles  was  observed  in  this  situation.  However, 
when  the  beetles  were  starved  for  several  days  and  were  kept 
without  ants,  they  occasionally  chased  each  other,  jumping  on  each 
other’s  back  as  they  normally  did  when  hunting  ants.  But  we  never 
saw  cannibalistic  behavior  among  the  adult  beetles,  even  when  the 
beetles  were  densely  crowded  around  a prey  object  (Fig.  2B). 

Defense  and  appeasement  behavior  in  adult  beetles 
Defense  with  tergal  gland  secretion: 

Usually  the  Pella  beetles  run  around  with  their  abdomen  curved 
slightly  upwards.  When  encountering  an  ant,  the  beetles  flex  the 
abdomen  even  more  strongly.  This  is  a typical  and  frequently 
described  behavior  of  many  staphylinid  myrmecophiles  and  is 
commonly  considered  a defense  response  (Wasmann  1886,  1920; 
Jordan  1913;  Patrizi  1948;  Koblick  and  Kistner  1965;  Pasteels  1968; 
Holldobler  1970,  1972;  Kolbe  1971).  It  has  been  suggested  that 
during  this  abdominal  flexing  the  beetles  discharge  secretions  from 
their  tergal  gland  (Jordan  1913;  Kistner  and  Blum  1971). 

The  tergal  gland  is  located  between  the  sixth  and  seventh 
abdominal  tergites  (Fig.  6),  and  is  unique  to  the  subfamily  Aleo- 
charinae  (Jordan  1913;  Pasteels  1968).  The  chemistry  of  the  tergal 
gland  secretions  of  several  species  has  been  investigated  and  found 
to  be  extraordinarily  diverse  (Blum  et  al.  1971;  Brand  et  al.  1973; 
Kolbe  and  Proske  1973). 

Kistner  and  Blum  (1971)  suggested  that  Pella  japonicus  and 
possibly  also  P.  comes , both  of  which  live  with  Lasius  spathepus, 
produce  citronellal  in  their  tergal  glands.  This  substance  is  also  a 


358 


Psyche 


[Vol.  88 


Fig.  4 Sequence  of  group  hunting  behavior  of  Pella.  A)  Two  beetles  chase  a 
forager  of  L.  fuliginosus.  One  of  the  beetles  is  jumping  on  the  back  of  the  ant.  B)  The 
ant  has  been  captured  and  subdued  by  both  beetles.  C)  A third  beetle  is  joining  the 
hunting  group. 


1981] 


Holldobler,  Moglich,  & Maschwitz  — Pella 


359 


Fig.  5 A)  Pella  in  “death  feigning”  position.  B)  A “death  feigning”  beetle  is 
carried  around  by  a L.  fuliginosus  worker.  C)  Pella  presents  abdominal  tip  to 
attacking  L.  fuliginosus  worker.  The  ant  licks  at  the  abdominal  tip. 


360 


Psyche 


[Vol.  88 


major  compound  of  the  mandibular  gland  secretions  of  their  host 
ants,  for  which  it  functions  as  an  alarm  pheromone.  Although  no 
Pella  tergal  gland  contents  were  available  for  chemical  analysis, 
because  irritated  beetles  seemed  to  smell  like  the  ants’  mandibular 
gland  secretion,  Kistner  and  Blum  speculated  that  Pella  produce  in 
their  tergal  glands  citronellal  and  thereby  mimic  the  alarm  phero- 
mone of  their  host  ants.  They  suggested  that  in  this  way  the  beetles 
can  “cause  the  ants  to  reverse  their  direction;  a reaction  which 
allows  the  myrmecophiles  to  escape”. 

Our  investigations  of  the  defensive  strategy  employed  by  the 
European  Pella  towards  their  host  ants  Lasius  fuliginosus  led  to 


m ]j 


Fig.  6 Schematical  drawing  of  a Pella  beetle  indicating  the  position  of  the 
exocrine  glandular  complexes.  TG:  tergal  gland;  AG:  appeasement  gland  complex. 


1981] 


Holldobler,  Moglich,  & Maschwitz  — Pella 


361 


different  results.  Pella  laticollis , when  irritated  mechanically  dis- 
charges a pungent  smelling  brownish  secretion  from  its  tergal  gland 
that  shows  acidic  reactions.  Only  when  the  beetles  were  severely 
attacked  and  firmly  grasped  on  their  appendages  by  the  ants,  could 
we  smell  the  tergal  gland  secretion.  We  never  observed  the  beetles 
employing  tergal  gland  secretion  when  they  were  attacking  ants. 
Ants  contaminated  with  tergal  gland  secretion  usually  exhibited  a 
repellent  reaction,  releasing  the  grip  on  the  beetles  and  grooming 
and  wiping  their  mouth  parts  and  antennae  on  the  substrate.  But  the 
beetles  had  to  escape  quickly,  because  other  ants  close  by  became 
alerted  and  were  rapidly  approaching  the  scene,  apparently  alarmed 
by  the  ants’  alarm  pheromone.  We  noticed  that  beetles  that  were 
attacked  by  L.  fuliginosus  workers  often  smelled  somewhat  like  the 
ants’  mandibular  secretions,  but  the  beetles’  tergal  gland  secretions 
clearly  smelled  differently.  Conceivably,  some  of  the  attacked 
beetles  were  contaminated  with  the  ants’  strongly  smelling  mandib- 
ular gland  secretions. 

Our  chemical  analysis  of  the  tergal  gland  secretions  of  P.  laticollis 
did  not  reveal  a resemblance  to  the  mandibular  gland  secretions  of 
L.  fulginosus,  whose  major  compounds  are  farnesal,  6-methyl-5- 
hepten-2-one;  perillene  and  dendrolasin,  a furan  (Quilico  et  al  1957; 
Bernardi  et  al  1967).  When  we  treated  the  tergal  gland  section  with 
2,4 — dinitrophenyl-hydrazine,  we  obtained  an  orange-yellowish  pre- 
cipitate. This  was  subjected  to  thinlayer  chromatography  in  two 
separate  systems.  In  each  system  we  obtained  two  spots.  The  Rf 
values  and  the  color  reaction,  when  treated  with  ammonia  vapour, 
identified  them  as  dinitrophenylhydrazones  of  p-benzoquinone  and 
p-toluquinone.  Furthermore,  the  chromatography  of  the  hydro- 
quinones  obtained  from  the  secretion  by  reduction  with  SO2  also 
demonstrated  the  presence  of  p-benzo-  and  p-toluquinone  in  the 
tergal  gland  secretion. 

For  comparison  we  used  thinlayer  chromatography  to  analyze  the 
dinitrophenylhydrazones  of  the  tergal  gland  secretion  of  several 
other  aleocharine  staphylinds  found  near  the  nests  of  L.  fuliginosus. 
Pella  humeralis,  A theta  fungi  and  Sipalia  cireellaris  also  produce 
benzo-  and  toluquinone;  in  Oxypoda  vittata  we  found  only  tolu- 
quinone. 

In  addition  Kolbe  and  Proske  (1973)  identified  isovaleric  acid  in 
the  tergal  gland  secretion  of  P.  humeralis , and  with  the  aid  of  gas 
chromatography  we  detected  saturated  hydrocarbons  and  short 


362 


Psyche 


[Vol.  88 


,o  O 
ir  C 

>->  D 

3 ’rt 

Xi  > 

c C 


I 


1981] 


Holldobler,  Moglich,  & Maschwitz  — Pella 


363 


chained  fatty  acids  in  the  secretion  of  P.  laticollis  (Tab.  2).  However, 
in  none  of  the  species  could  we  find  citronellal. 

The  common  presence  of  quinones  in  the  tergal  gland  secretions 
of  Pella  and  the  related  aleocharine  species  agree  with  the  previous 
findings  by  Blum  et  al  (1971),  who  found  the  tergal  gland  secretion 
of  Lomeehusa  strumosa  to  contain  benzoquinone,  methyl-benzo- 
quinone,  ethyl-benzoquinone  and  n-tridecane,  the  latter  substance 
accounting  for  more  than  80%  of  the  volatiles  detected  in  the 
secretion.  In  addition  Brand  et  al.  (1973)  analyzed  the  tergal  gland 
secretion  of  Drusilla  canaliculata,  also  an  aleocharine  beetle,  finding 
quinones  and  hydroquinones  together  with  alkanes,  saturated  and 
unsaturated  aliphatic  aldehydes.  Pasteels  (1968)  demonstrated  that 
D.  canaliculata  effectively  employs  the  tergal  gland  secretion  as  a 
repellent-defense  weapon  against  ants  in  a similar  fashion  as  we 
described  it  for  Pella. 

Although  we  could  not  find  any  resemblance  of  the  Pella  tergal 
gland  secretions  to  the  mandibular  gland  secretions  of  Lasius 
fuliginosus,  it  was  noteworthy  that  the  Pella  secretions  contained 
undecane,  a hydrocarbon  commonly  found  in  the  Dufour’s  glands 
of  formicine  ants  (for  review  see  Blum  and  Hermann  1978)  and 
considered  to  be  an  alarm  pheromone  in  L.  fuliginosus  (Dumpert 
1972).  However,  isolated  tergal  gland  secretions  of  P.  laticollis 
elicited  a repellent  reaction  rather  than  an  alarm  response  in  L. 
fuliginosus.  Apparently  the  repellent  effect  of  the  quinones  in  the 
secretions  is  stronger  than  a possible  alarming  effect  released  by 
undecane.  In  fact,  when  the  ant’s  antennae  were  directly  contam- 
inated with  the  beetles’  tergal  gland  secretions  the  antennae  were 
hanging  almost  motionless  and  flabby  and  the  ant  appeared  dis- 
oriented for  several  minutes.  From  all  our  laboratory  tests  it 
appears  obvious  that  the  tergal  gland  secretions  of  Pella  functions  as 
a powerful  chemical  defense  weapon  against  attacks  by  ants. 

Appeasement  behavior: 

When  foraging  on  the  ants’  “garbage  dumps”  or  running  along 
the  ants’  trails,  Pella  frequently  encounter  ants.  Yet  we  were 
impressed  by  the  scarcity  of  their  application  of  the  tergal  gland 
defensive  system.  Much  more  frequently  the  beetles  employed  an 
appeasing  defensive  strategy,  and  the  repellent  defense  seemed  to  be 
employed  only  as  a last  resort. 


364 


Psyche 


[Vol.  88 


Fig.  7 A,B)  Transversal  section  through  glandular  epithelium  in  9th  sternum  of 
P.  humeralis.  C)  Sagital  section  through  sternal  gland  in  7th  sternum  of  P. 
humeralis.  GC:  glandular  cell;  P:  pore  in  cuticle. 


1981] 


Holldobler,  Moglich,  & Maschwitz  — Pella 


365 


It  was  especially  common  in  early  spring,  when  most  of  the 
beetles  were  found  close  to  the  entrance  of  the  ants’  nest,  that  the 
beetles  showed  “death  feigning”  behavior,  when  attacked  by  ants. 
They  fell  to  the  side,  the  legs  and  antennae  folded  tightly  to  the  body 
and  the  abdomen  curved  upwards  (Fig.  5A).  The  ants  either  ignored 
these  motionless  beetles  or  carried  them  around  and  finally  dis- 
carded them  on  the  “garbage  dump”.  But  only  rarely  did  they  injure 
the  beetles  (Fig.  5B). 

Later  in  the  year,  when  the  activity  of  ants  and  beetles  was  much 
higher,  the  beetles  employed  a different  appeasement  technique.  As 
mentioned  before,  we  only  very  rarely  saw  the  discharge  of  tergal 
gland  secretions  by  the  beetles,  although  every  time  they  en- 
countered ants  they  flexed  their  abdomen  and  pointed  with  the 
abdominal  tip  toward  the  head  of  their  adversaries.  Usually  the  ants 
responded  by  antennating  the  tip  and  briefly  licking  it  (Fig.  5C). 
This  ordinarily  slowed  down  the  ants’  aggression  and  the  beetles 
used  the  ants’  distraction  to  escape.  Occasionally,  when  the  ants 
remained  very  persistent,  a white,  viscous  droplet  appeared  at  the 
abdominal  tip,  whereupon  the  ants  usually  very  eagerly  licked  it  up. 
This  appeasing  defensive  behavior  was  much  more  common  during 
the  interactions  between  Pella  and  Lasius  fuliginosus  than  the 
repellent  defense.  For  a series  of  simulation  experiments  we  cut  off 
the  last  3 segments  of  the  abdomen  of  freshly  killed  P.  laticollis, 
sealed  the  cut  with  wax,  pinned  the  segments  on  dissecting  needles 
and  presented  these  “dummies”  to  the  ants.  In  a total  of  60  tests 
(using  3 different  dummies)  the  ants  interrupted  their  run  in  47  cases 
(78%)  and  licked  the  abdominal  tip  briefly. 

Histological  investigations  revealed  that  the  abdominal  tips  of 
Pella  are  batteries  of  exocrine  glandular  structures,  all  of  which 
together  we  call  the  appeasement  gland  complex.  In  the  following 
section  we  give  a brief  description  of  the  glands  which  could  be 
involved  in  the  appeasement  behavior. 

The  most  comprehensive  study  of  the  glandular  morphology  of 
some  termitophilous  and  myrmecophilous  aleocharine  beetles  has 
been  published  by  Pasteels  (1968).  From  this  work  we  learned  that 
these  beetles  possess  a surprising  variety  of  exocrine  glandular 
structures  and  that  various  species  can  differ  considerably  in  their 
glandular  systems.  In  the  four  species  of  Pella  ( P . cognata,  P. 
funesta,  P.  humeralis,  P.  laticollis ) we  investigated,  we  did  not  find 
major  differences,  although  P.  humeralis  appeared  to  be  somewhat 


366 


Psyche 


[Vol.  88 


Fig.  8 Sagital  section  through  abdominal  tip  of  P.  humeralis.  GC:  glandular  cells 
under  10th  tergite  near  anus.  A:  anus. 


more  richly  endowed  with  hypodermal  glandular  cells,  especially  in 
the  area  of  the  paratergites. 

In  staphylinid  beetles  the  first  fully  developed  abdominal  seg- 
mental ring  (tergite  plus  sternite)  is  usually  considered  to  be  the  3rd 
abdominal  segment  (Blackwelder  1936).  All  Pella  species  have  a well 
developed  compound  tergal  gland  between  the  6th  and  7th  tergites 
(Fig.  6)  as  described  by  Jordan  (1913),  Pasteels  (1968)  and  Holl- 
dobler  (1970).  We  have  also  detected  glandular  cells  located  pri- 
marily in  the  7th  segment,  which  Pasteels  (1968)  calls  postpleural 
glands.  According  to  Pasteels  the  glandular  channels  associated 
with  these  cells  open  dorsolaterally  through  the  pleural  membrane 
between  the  7th  and  8th  segments.  Pasteels  could  clearly  see  these 
openings  in  several  species  (for  example  in  Gyrophanaena  affinis ), 
but  not  in  Pella  (Zyras)  humeralis.  In  a series  of  longitudinal,  trans- 
versal and  frontal  sections,  we  too  were  unable  to  detect  the  external 
openings  of  these  glandular  cells. 

At  the  anterior  edge  of  the  4th,  5th,  6th  and  7th  sternites  are 
found  clusters  of  glandular  cells  that  open  through  pores  in  the 
cuticle  (Fig.  7A).  They  are  especially  well  developed  in  the  7th 
sternite.  Pasteels  (1968)  assumes  that  the  secretions  of  these  glands 


1981] 


Holldobler,  Moglich,  & Maschwitz  — Pella 


367 


Fig.  9 Sagital  sections  through  abdominal  tip  of  Pella  laticollis  female.  GC: 
glandular  cell  clusters;  M:  membrane;  CH:  glandular  channels. 


368 


Psyche 


[Vol.  88 


serve  primarily  for  lubrication  to  alleviate  friction  between  the 
sternal  sclerites  when  the  beetles  flex  their  abdomen.  But  more  than 
any  other  of  the  abdominal  segments  the  last  3 tergites  (VII,  IX,  X) 
(Fig.  6)  and  the  8th  and  9th  sternites  are  richly  endowed  with 
glandular  epithelia,  the  individual  cells  of  which  open  through  pores 
in  the  cuticle  (Fig.  IB,  1C).  The  last  two  segments  can  be  telescoped 
with  especial  ease  into  the  preceding  segments,  and  during  the 
appeasement  process  the  beetles  often  move  them  slightly  back  and 
forth.  Furthermore,  there  are  clusters  of  glandular  cells  with  longer 
channels  under  the  10th  tergite  near  the  anus  (Fig.  8).  They  resemble 
the  type  of  cells  that  Holldobler  (1971)  located  in  the  same  position 
in  Atemeles  and  called  pygidial  glands.  We  have,  however,  aban- 
doned this  term,  because  it  is  very  confusing,  especially  in  the 
Aleocharinae,  where  the  last  visible  tergite  is  usually  not  the  8th 
tergite  (often  called  pygidium  in  the  Coleoptera)  but  the  10th  tergite. 

In  addition  to  these  hypodermal  glandular  structures,  females  and 
males  possess  special  exocrine  glandular  complexes  that  might  be 
involved  in  the  reproductive  processes  but  which  could  also  play  a 
role  in  the  myrmecophilous  behavior  of  the  beetles.  In  the  9th 
sternite  of  females  there  are  several  clusters  of  glandular  cells,  the 
channels  of  which  open  through  the  intersegmental  membrane  at 
the  tip  of  the  abdomen  and  near  the  oviduct  (Fig.  9).  Males  have 
similar  glands  in  the  9th  sternite  which  also  open  through  the 
intersegmental  membrane  near  the  posterior  part  of  the  genital 
chamber  (Fig.  10).  Furthermore,  males  possess  a very  large  gland- 
ular complex,  consisting  of  numerous  tightly  packed  glandular  cells 
each  connected  with  a long  channel  that  open  dorsally  in  bundles 
through  a membrane  at  the  genital  chamber  (Fig.  11).  We  assume 
that  the  secretions  of  this  gland  flow  into  the  genital  chamber. 
Females  do  not  have  this  gland,  but  the  spermathecal  gland  has  a 
very  smilar  appearance. 

Finally,  the  hindgut  might  also  be  involved  in  the  appeasement 
process.  On  several  occasions  we  observed  that  beetles,  upon 
presenting  their  abdominal  tip  to  the  ants,  released  a droplet  at  the 
anus  that  was  licked  up  by  the  ants. 

Discussion: 

Some  of  the  most  advanced  myrmecophilic  relationships  are 
found  in  the  aleocharine  beetles  Lomechusa  and  Atemeles.  We 


1981] 


Holldobler,  Moglich,  & Maschwitz  — Pella 


369 


Fig.  10  A)  Sagital  sections  through  the  abdominal  tip  of  P.  humeralis  male.  T: 
9th  and  10th  tergite;  S:  9th  sternite;  AE:  aedeagus.  B)  Close-up  of  sagital  section 
through  10th  tergite  and  9th  sternite.  GC:  glandular  cell  clusters. 


370 


Psyche 


[Vol.  88 


Fig.  1 1 A)  Glandular  complex  in  males  of  Pella  humeralis,  located  dorsally  of  the 
genital  chamber.  B)  Opening  of  the  bundles  of  glandular  channels  through  mem- 
brane; presumably  into  the  genital  chamber. 


1981] 


Holldobler,  Moglich,  & Maschwitz  — Pella 


371 


know  from  observations  by  Wasmann,  made  more  than  60  years 
ago,  that  these  beetles  are  both  fed  and  reared  by  their  host  ants. 
Both  chemical  and  mechanical  interspecific  communication  is  in- 
volved in  these  unusual  relationships.  These  aleocharines  have 
broken  the  communication  code  of  their  host  ants  and  are  thereby 
able  to  live  as  parasites  within  the  social  system  of  the  ant  colony 
(Holldobler  1967,  1970,  1971,  1972). 

Species  of  the  genus  Pella  are  less  advanced  in  their  myrmeco- 
philic  relationships.  Rather  than  occupy  the  brood  chambers  of  the 
ant  nest,  they  live  as  scavengers  and  predators  in  the  peripheral 
zones  around  the  nest,  at  the  garbage  dumps,  and  on  the  trunk 
routes.  Some  of  the  behavioral  features  of  Pella , however,  seem  to 
be  very  similar  to  those  of  Atemeles  and  Lomechusa.  In  fact,  these 
behavioral  patterns  might  be  preadaptations  for  the  evolution  of  a 
highly  advanced  myrmecophilic  relationship  in  the  aleocharine 
beetles.  In  particular,  the  appeasement  behavior  appears  to  be  an 
important  prerequisite  for  living  closely  with  ants.  This  “gentle” 
defense  technique  does  not  cause  excitement  in  the  ants,  as  a 
repellent  defense  would  do. 

Indeed,  our  observations  indicate  that  Pella  only  rarely  employ 
their  strongly  smelling  tergal  gland  secretions  when  they  are  near  the 
host  ant  colony.  This  defense  system  might  be  used  more  during  the 
migration  phase,  when  the  beetles  can  be  attacked  by  individual 
foraging  ants.  Similar  results  were  previously  obtained  with  Ate- 
meles (Holldobler  1970)  and  Lomechusa  (Holldobler  unpublished). 
In  the  presence  of  their  host  ants  these  species  use  the  appeasement 
defense  almost  exclusively. 

The  appeasement  behavior  also  plays  an  important  role  during 
the  adoption  of  Atemeles  by  their  host  ants.  When  encountering  a 
worker  of  the  host  species  near  the  ant’s  nest,  the  beetle  first  offers 
the  appeasement  gland  complex  (at  the  abdominal  tip)  to  the  ant. 
This  apparently  suppresses  aggressive  behavior  in  the  ant;  only  then 
does  the  beetle  lower  its  abdomen  to  permit  the  ant  access  to  the 
adoption  glands,  which  are  located  in  the  paratergites.  The  glandu- 
lar openings  are  surrounded  by  bristles.  These  are  grasped  and  used 
by  the  ant  to  carry  the  beetle  into  the  brood  chamber.  While  being 
carried,  Atemeles  adopts  the  same  posture  as  that  used  by  Pella 
during  the  “death  feigning”  behavior.  As  we  have  noted,  the  initially 
aggressive  ants  respond  by  either  ignoring  the  beetles  or  else  picking 


372 


Psyche 


[Vol.  88 


them  up  and  carrying  them  around  until  they  eventually  discard 
them,  usually  unharmed,  at  the  garbage  dump.  It  is  conceivable  that 
the  carrying  posture  of  Atemeles  has  evolved  from  a defensive 
“death  feigning”  behavior  employed  by  less  advanced  ancestral 
species. 

Finally,  Pella  beetles  do  not  have  adoption  glands  associated  with 
trichrome  bristles.  It  is  most  likely,  however,  that  the  small  clusters 
of  glandular  cells  in  the  paratergites  (for  example  in  P.  humeralis ) 
represent  morphological  precursors  of  the  massively  developed 
adoption  glands  in  Atemeles  and  Lomechusa. 

Acknowledgements 

We  would  like  to  thank  Hiltrud  Engel  for  technical  assistance, 
Dr.  V.  Puthz  for  identifying  the  beetles,  Ruiko  Pierce  for  translating 
the  work  of  Yasumatsu  from  Japanese  into  English,  and  A1  Newton 
and  Margaret  Thayer  for  helping  in  disentangling  the  segmental 
morphology  of  Pella.  This  work  was  supported  by  grants  from  the 
Deutsche  Forschungsgemeinschaft  and  National  Science  Founda- 
tion. 


References 

Bernardi,  R.,  C.  Cardani,  D.  Ghiringhella,  A.  Selva,  A.  Baggini,  M.  Pavan 
1967.  On  the  components  of  secretion  of  mandibular  glands  of  the  ant  Lasius 
(Dendrolasius)  fuliginosus.  Tetrahedron  Lett.  40,  3893-3896. 

Blum,  M.  S.,  R.  M.  Crewe,  J.  M.  Pasteels 

1971.  Defensive  secretions  of  Lomechusa  strumosa  a Myrmecophilous  Beetle. 
Ann.  Entomol.  Soc.  America  64,  975-975. 

Blum,  M.  S.,  H.  R.  Herman 

1978.  Venoms  and  Venom  Apparatuses  of  the  Formicidae:  Myrmeciinae, 
Ponerinae,  Dorylinae,  Pseudomyrmecinae,  Myrmicinae  and  Formi- 
cinae;  in  Handbook  of  Experimental  Pharmacology  48,  pp.  801-869. 
Springer  Verlag,  Berlin,  Heidelberg,  New  York. 

Brand,  J.  M.,  M.  S.  Blum,  H.  M.  Fales  and  J.  M.  Pasteels 

1973.  The  Chemistry  of  the  Defensive  Secretion  of  the  Beetle  Drusilla 
canaliculata.  J.  Insect  Physiol.  19,  369-382. 

Dumpert,  K. 

1972.  Alarmstoffrezeptoren  auf  der  Antenne  von  Lasius  fuliginosus  (Latr.) 
(Hymenoptera,  Formicidae).  Z.  Vergl.  Physiol.  76,  403-425. 

Gobwald,  K.,  Kloft,  W. 

1958.  Radioaktive  Isotope  zur  Erforschung  des  Staatenlebens  der  Insekten. 
Umschau,  58,  743-745. 


1981] 


Holldobler,  Moglich,  & Maschwitz  — Pella 


373 


Holldobler,  B. 

1967.  Zur  Physiologie  der  Gast-Wirt-Beziehungen  (Myrmecophilie)  bei  Amei- 
sen.  I.  Das  Gastverhaltnis  der  Atemeles-  und  Lomechusa  Larven  (Col. 
Staphylinidae)  zu  Formica  (Hym.  Formicidae).  Z.  Vergl.  Physiol.  56, 
1-21. 

1970.  Zur  Physiologie  der  Gast-Wirt-Beziehungen  (Myrmecophilie)  bei  Amei- 
sen.  II.  Das  Gastverhaltnis  des  imaginalen  Atemeles  pubicollis  Bris. 
(Col.  Staphylinidae)  zu  Mvrmica  und  Formica  (Hym.  Formicidae).  Z. 
Vergl.  Physiol.  66,  176-189. 

1971.  Communication  between  ants  and  their  guests.  Sci.  American,  March 
1971.  86-91. 

1972.  Verhaltensphysiologische  Adaptationen  an  okologische  Nischen  in 
Ameisennestern.  Verhandlungsbericht  der  Dtsch.  Zool.  Ges.  65.  Jah- 
resvers.  137-143. 

Jordan,  K.  H.  C. 

1913.  Zur  Morphologie  und  Biologie  der  myrmecophilen  Gattungen  Lome- 
chusa und  Atemeles  und  einiger  verwandter  Formen.  Z.  Wiss.  Zool.  107, 
346-386. 

Kistner,  D.  H. 

1971.  Studies  of  Japanese  Myrmecophiles  Part  I.  The  Genera  Pella  and 
Falagria  (Coleoptera,  Staphylinidae)  in  Entomological  Essays  to  Com- 
memorate the  Retirement  of  Professor  K.  Yasumatsu,  pp.  141-165. 
Hokurynkan  Publ.  Co.  Ltd.  Tokyo. 

1979.  Social  and  Evolutionary  Significance  of  Social  Insect  Symbionts.  In 
Social  Insects  (vol  I)  (ed.  H.  R Hermann)  pp.  340-413,  Academic  Press, 
Inc.  New  York,  San  Francisco,  London. 

Kistner,  D.  H.,  M.  S.  Blum 

1971.  Alarm  Pheromone  of  Lasius  (Dendrolasius)  spathepus  (Hymenoptera: 
Formicidae)  and  Its  Possible  Mimicry  by  Two  Species  of  Pella  (Cole- 
optera: Staphylinidae).  Ann.  Entomol.  Soc.  America  64,  589-594. 

Kloft,  W. 

1959.  Direktes  und  indirektes  Verfahren  zur  Messung  der  Beta-Strahlenab- 
sorption  von  kleinen  Gewebeschichten  an  Insekten.  Glas-Intr.  Techn.  3, 
79-82. 

Koblick,  T.  A.,  D.  H.  Kistner 

1965.  A revision  of  the  species  of  the  genus  Myrmechusa  from  tropical  Africa 
with  notes  on  their  behavior  and  their  relationship  to  Pygostenini 
(Coleoptera,  Staphylinidae).  Ann.  Ent.  Soc.  Amer.  58,  28-44. 

Kolbe,  W. 

1971.  Untersuchungen  iiber  die  Bindung  von  Zyras  humeralis  (Coleoptera, 
Staphylinidae)  an  Waldameisen.  Entomol.  Blatter  67,  129-136. 

Kolbe,  W.,  M.  G.  Proske 

1973.  Iso-Valeriansaure  im  Abwehrsekret  von  Zyras  humeralis  (Col.  Staphy- 
linidae). Entomol.  Blatter  69,  57-60. 

Maschwitz,  U.,  B.  Holldobler 

1970.  Der  Kartonnestbau  bei  Lasius  fuliginosus  Latr.  (Hym.  Formicidae).  Z. 
Vergl.  Physiol.  66,  176-189. 


374 


Psyche 


[Vol.  88 


Moore,  B.  P. 

1968.  Studies  on  the  chemical  composition  and  function  of  the  cephalic  gland 
secretion  in  Australian  termites.  J.  Insect  Physiol.  14,  33-39. 

Pasteels,  J.  M. 

1968.  Le  systeme  glandulaire  tegumentaire  des  Aleocharinae  (Coleoptera, 
Staphylinidae)  et  son  evolution  chez  les  especes  termitophiles  du  genre 
Termitella.  Arch.  Biol.  (Liege)  79,  381-469. 

Patrizi,  S. 

1948.  Contribuzioni  alia  conoscenze  delle  formiche  e dei  mirmicofile  dell’ 
Africa  orientale.  V.  Note  etologiche  su  Myrmechusa  Wasmann  (Cole- 
optera, Staphylinidae).  Bull.  1st.  Entomol.  Univ.  Bologno  17,  168-173. 

Quilico,  A.,  F.  Piozzi,  M.  Pavan 

1957.  The  structure  of  dendrolasin.  Tetrahedron  1,  177-185. 

Rathmayer,  W. 

1962.  Methylmethacrylat  als  Einbettungsmedium  fur  Insekten.  Experientia 
(Basel)  18,  47-48. 

SCHILDKNECHT,  H.,  H.  KRAMER 

1962.  Zum  Nachweis  von  Hyrdochinonen  neben  Chinonen  in  den  Abwehr- 
blasen  von  Arthropoden.  Z.  Naturforsch.  176,  701-702. 

Wasmann,  E. 

1886.  Uber  die  Lebensweise  einiger  Ameisengaste.  Dtsch.  Entomol.  Zeitschr. 
30,  49-66. 

1920.  De  Gastpflege  der  Ameisen.  Gebrtider  Borntraeger,  Berlin. 

1925.  Die  Ameisenmimikry.  Gebriider  Borntraeger,  Berlin. 

1930.  Zur  Biologie  von  Myrmedonia  (Zyras).  Entomol.  Berichte  8,  150-151. 

Wilson,  E.  O. 

1971.  The  Insect  Societies.  The  Belknap  Press  of  Harvard  University  Press, 
Cambridge,  Mass. 

Yasumatsu,  K. 

1937.  Lasius  fuliginosus  (Latreille)  var.  spathepus  (Wheeler  and  its  synecht- 
trans  Zyras  comes  Sharp  and  Zyras  cognathus  Markel  var . japonicus 
Sharp.  Nippon  no  Kochu  1,  47-51. 


BEHAVIORAL  ORIGIN  OF  TREMULATION, 

AND  POSSIBLE  STRIDULATION, 

IN  GREEN  LACEWINGS 
(NEUROPTERA:  CHRYSOPIDAE)1 

By  Peter  Duelli2  and  James  B.  Johnson3 

University  of  California,  Berkeley, 

Division  of  Biological  Control 
1050  San  Pablo  Ave. 

Albany,  CA  94706  USA 

Introduction 

Abdominal  vibration  or  “jerking”  in  connection  with  courtship 
behavior  has  been  described  for  several  green  lacewing  species  (e.g. 
Smith  1922;  Toschi  1965;  Tauber  1969;  Sheldon  and  MacLeod 
1974)  and  explored  in  detail  by  Henry  (1979,  1980a,  b,  c).  In 
Chrysoperla  carnea  (Stephens)  isolated  individuals  produce  long, 
patterned  sequences  of  discrete  short  bursts  of  rhythmic  vibration  of 
the  abdomen  in  the  vertical  plane.  The  wings  may  also  vibrate. 
Sexually  receptive  pairs  establish  duets  of  reciprocal  abdominal 
jerking.  Actual  drumming  of  the  abdomen  on  the  substrate  does  not 
occur.  It  had  been  assumed  that  abdominal  vibration  produces 
high-frequency  sounds  by  stridulation  (Adams  1962,  Riek  1967, 
Eichele  and  Villiger  1974,  Henry  1979)  and  acoustical  communica- 
tion was  discussed  in  connection  with  the  tympanal  ultrasound 
receptor  organ  described  by  Miller  (1970,  1971).  Courtship  and 
copulation  take  place  on  the  vegetation,  usually  on  the  underside  of 
leaves.  Henry  (1980a,  c)  in  his  work  with  Chrysoperla  spp.  demon- 
strated that  communication  is  performed  via  low-frequency  sub- 
strate vibration  and  not  by  airborne  sound.  Males  were  able  to 
establish  duets  with  females  within  a range  of  15  cm.  According  to 
Henry  (1980a,  b,  c),  differences  in  the  vibration  patterns  of  Chry- 


1 Published  with  the  approval  of  the  Director  of  the  Idaho  Agricultural  Experiment 
Station  as  Research  Paper  No.  81613. 

2Present  address:  Zoologisches  Institut,  Universitat  Basel,  Rheinsprung  9,  4051  Basel 
Switzerland. 

3Present  address:  Department  of  Entomology,  University  of  Idaho,  Moscow,  Idaho 
83843  U.S.A. 

Manuscript  received  by  editor  December  15,  1981. 


375 


376 


Psyche 


[Vol.  88 


soperla  rufilabris  (Burmeister),  C.  downesi  (Banks)  and  C.  carnea 
suggest  that  “acoustical”  communication  may  help  to  reproduc- 
tively  isolate  sympatric  lacewing  species. 

Since  the  vibration  produced  by  abdominal  jerking  in  lacewings 
seems  to  be  propagated  in  a transverse  wave  (perpendicular  to  the 
plane  of  the  substrate)  we  prefer  to  call  this  type  of  communication 
“tremulation”,  following  Busnel  et  al.  (1956),  Henry  (1980c)  and 
Morris  (1980).  On  the  other  hand,  sound  in  the  form  of  longitudinal 
waves,  is  produced  by  stridulation  and  percussion.  Possible  stridula- 
tory  structures  in  lacewings  were  first  described  for  the  chrysopid 
Meleoma  schwarzi  (Banks)  by  Adams  (1962)  and  later  for  other 
Neuroptera  (Riek  1967).  However,  to  date,  there  is  no  reported 
record  of  any  sound  produced  by  these  organs  (Henry  1980c).  In  M. 
schwarzi,  sound  may  be  produced  when  the  second  abdominal 
sternite,  with  its  regular  striae  of  microtrichia,  is  rubbed  against  the 
femora  by  abdominal  vibration  (Adams  1962).  C.  carnea  and  some 
other  species  of  Chrysopidae  may  stridulate  using  microtrichia  on 
the  venter  of  the  anal  lobe  of  the  forewings  and  dorsolaterad  on  the 
metanotum  (Riek  1967;  Henry  1979).  Alternatively,  these  paired 
areas  of  microtrichia  may  function  to  hold  the  wings  in  place  when 
at  rest  (Henry  1980c).  Thus,  tremulation  and  possible  stridulation 
are  both  produced  by  vibrating  the  wings  and  abdomen. 

Methods  and  Materials 

Observations  of  free  flight  and  mating  behavior  were  made  on  the 
following  species:  C.  carnea,  Eremochrysa punctinervis  McLachlan, 
E.  tibialis  Banks,  Mallada  basalis  (Walker),  Meleoma  hageni  Banks 
and  Nodita  n.  sp.  The  only  specialized  technique  required  for  this 
study  was  the  use  of  a strobe  light  to  illuminate  lacewings  on  a flight 
mill  (Duelli  1980).  By  varying  the  frequency  of  the  strobe  flashes,  it 
was  possible  to  determine  the  rate  of  the  wing  beats  and  other  body 
movements,  as  the  highest  flash  frequency  at  which  the  motion 
appeared  to  be  “frozen”  and  each  structure  was  seen  in  only  one 
position.  A multiple  of  this  frequency  again  produces  a frozen 
image,  but  the  body  is  seen  in  multiple  positions.  C.  carnea  and  M. 
basalis  were  examined  in  this  manner. 

Results 

Among  the  species  observed,  there  appeared  to  be  great  variation 
in  the  patterns  and  intensities  of  vibration  of  the  wings  and 
abdomen  during  courtship,  but  this  was  not  quantified.  The  beha- 


1981] 


Duelli  & Johnson  — Green  Lacewings 


377 


vior  was  strongly  developed  in  species  of  the  genera  Meleoma  and 
Eremochrysa,  but  was  even  more  conspicuous  in  the  Indo-Pacific 
lacewing  M.  basalis,  as  observed  on  the  island  of  American  Samoa. 
In  this  species,  the  males  flapped  their  wings  so  vigorously  that  they 
hit  the  substrate  and  produced  sounds  easily  perceptible  to  the 
human  ear.  Heavily  developed  pterostigmata  in  the  hind  wings  of 
the  male  may  enhance  substrate  vibration  and  protect  the  wings 
from  damage  (Fig.  1).  During  courtship,  the  males  moved  forward 


Figures  1-5.  Fig.  1.  Forewing  and  hind  wing  of  female  (left)  and  male  (right) 
Mallada  basalis.  The  arrow  indicates  the  heavily  developed  pterostigma  in  the 
hindwing  of  the  male.  Fig.  2.  Stationary  flight  of  a tethered  Chrysoperla  carnea 
male.  Strobe  flashes  (60  Hz,  exposure  0.25  sec)  show  the  extent  of  the  abdominal 
motion.  Fig.  3.  Chrysoperla  carnea  male  mounted  horizontally  on  a flight  mill. 
Strobe  flashes  (54  Hz)  show  the  exact  antiphase  of  abdominal  and  wing  vibration. 
Fig.  4.  Chrysoperla  carnea  male  mounted  on  a flight  mill  in  “natural”  flight  position 
as  shown  in  figure  2.  Any  forced  deviation  from  the  “natural”  body  angle  leads  to  an 
increased  amplitude  of  the  abdominal  vibration  (See  figure  3 for  comparison). 
Fig.  5.  Same  specimen  and  same  position  as  in  Fig.  3.  400  Hz  strobe  flashes  show 
the  full  flow  of  the  movements  of  wings  and  abdomen. 


378 


Psyche 


[Vol.  88 


and  backward  in  front  of  the  female,  and  sometimes  even  sideways. 
Especially  vigorous  males  were  seen  to  perform  small  jumps, 
reminiscent  of  take-off  behavior. 

A chrysopid  usually  flies  with  its  head  higher  than  its  abdomen.  If 
the  insect  is  mounted  on  a flight  mill  and  illuminated  with  a strobe 
light,  the  abdomen  can  be  seen  moving  up  and  down  in  the  same 
way  as  described  for  the  courtship  behavior  (Fig.  2).  When  mounted 
horizontally,  the  abdominal  movements  were  exaggerated  (Fig.  3). 
In  both  orientations  the  strobe  flashes  revealed  that  the  frequencies 
of  the  wing  beat  and  of  the  abdominal  vibration  were  the  same. 
With  each  down-stroke  of  the  wings  the  abdomen  was  lifted  (Figs.  3 
and  4).  The  flow  of  the  movements  can  be  seen  in  Fig.  5. 

Observations  made  during  this  study  indicate  that  the  wing  beat 
frequency  was  positively  correlated  with  temperature  and,  in  gen- 
eral, negatively  correlated  with  wing  length.  At  23°  C,  a wing  beat 
frequency  of  27  Hz  (strokes /sec)  was  recorded  for  C.  carnea  and  38 
Hz  for  the  smaller  M.  basalis.  Miller  (1975)  reported  similar  results, 
25  Hz  at  21-24°C  in  tethered  flying  C.  carnea. 

Discussion 

The  frequencies  of  abdominal  vibration  during  courtship  have 
been  reported  for  three  species  of  Chrysoperla.  For  C.  rufilabrus  the 
rates  were  14-18  Hz  (Henry  1980a).  No  temperature  data  were 
given.  In  C.  carnea  the  frequencies  varied  from  30  to  100  Hz  at 
24-28° C (Henry  1980c),  while  the  courtship  behavior  of  C.  downesi 
included  volleys  of  abdominal  vibration  with  a frequency  of  60-80 
Hz,  with  a mean  of  approximately  73  Hz,  at  24-29°  C (Henry 
1980b). 

The  greater  variability  in  the  frequencies  of  abdominal  vibration 
during  courtship,  relative  to  flight,  is  probably  related  to  two 
factors.  First,  there  is  no  minimum  rate  of  wing  beats  necessary  to 
maintain  flight.  Second,  the  maximum  possible  rate  is  increased, 
since  the  wings  merely  vibrate  rather  than  making  full  strokes. 
These  would  open  a wide  range  of  frequencies  for  chrysopids  to  use 
in  tremulation.  If  character  displacement  occurred,  as  hypothesized 
by  Henry  (1980b),  this  would  tend  to  expand  the  range  of  frequen- 
cies actually  used  by  chrysopids. 

Tremulation  has  also  been  reported  in  the  courtship  behavior  of 
other  groups  of  insects.  Plecoptera  communicate  via  the  substrate 
by  drumming  with  their  abdomens  (Rupprecht  1968).  Similar 


1981] 


Duelli  & Johnson  — Green  Lacewings 


379 


drumming  and/or  abdominal  vibration  is  known  from  certain 
Psocoptera  (Pearman  1928),  Orthoptera  (refs,  in  Rupprecht  1968 
and  Morris  1980),  Megaloptera  (Rupprecht  1975)  and  Mecoptera 
(Rupprecht  1974).  Wing  fluttering  is  also  involved  in  courtship  of 
Panorpa  spp.  (Mecoptera)  (Rupprecht  1974)  and  three  genera  of 
Coniopterygidae  (Johnson  and  Morrison  1979). 

The  function  of  the  abdominal  motion  in  flight  is  unknown.  In 
the  Diptera,  the  halteres  (modified  second  pair  of  wings)  have  been 
shown  to  act  as  specialized  organs  to  maintain  flight  stability 
(Pringle  1948).  They  vibrate  in  a vertical  or  nearly  vertical  plane 
and,  as  gyroscopic  indicators,  reveal  any  change  in  the  spatial 
orientation  of  the  thorax  via  sensors  at  their  bases.  The  halteres 
vibrate  with  the  same  frequency  as  the  wings,  but  in  antiphase.  Since 
the  same  is  true  for  the  abdominal  movements  in  lacewings,  it  is 
tempting  to  regard  their  abdominal  vibration  as  an  analogous 
gyroscopic  mechanism  to  stabilize  the  orientation  of  the  thorax 
during  the  slow  hovering  flight,  thus  keeping  the  insect  in  an  upright 
position  with  regard  to  the  horizontal  plane.  This  possibility  is 
supported  by  the  similar  orientation  and  abdominal  movements  of 
flying  Plecoptera,  Megaloptera  ( Sialis  spp.  and  Neohermes  sp.)  and 
Mecoptera  ( Panorpa  spp.)  as  observed  in  the  field. 

Indirect  morphological  evidence  also  supports  this  possibility. 
Whereas  most  other  nocturnal  insects  have  large  ocelli,  chrysopids 
and  most  other  Neuroptera  lack  ocelli.  An  important  function  of  the 
ocelli  in  locusts  and  other  insects  is  to  recognize  relative  changes  in 
the  height  of  the  horizon  (Taylor  1981)  and  thus  to  stabilize  the 
flight  position. 

Based  on  the  similarities  between  abdominal  vibration  during 
flight  and  courtship  behavior,  we  suggest  that  tremulation  behavior 
in  lacewings  and  perhaps  other  slow-flying  insects  may  have  evolved 
from  a particular  “pre-adapted”  feature  in  the  take-off  and  flight 
behavior,  where  its  main  function  might  be  flight  stabilization. 


Acknowledgments 

We  wish  to  thank  Dr.  P.  A.  Adams  for  verifying  the  identifica- 
tions of  the  species  studied,  Dr.  K.  S.  Hagen  for  his  helpful 
discussions  and  Dr.  F.  M.  Carpenter  for  his  suggestions  regarding 
the  manuscript. 


380 


Psyche 


[Vol.  88 


Literature  Cited 

Adams,  P.  A.  1962.  A stridulatory  structure  in  Chrysopidae  (Neuroptera).  Pan 
Pac.  Entomol.  38(3):  178-180. 

Busnel,  R.  G.,  B.  Dumortier  and  M.  C.  Busnel.  1956.  Recherche  sur  le 
comportement  acoustique  des  ephippigeres  (Orthoptera:  Tettigoniidae).  Bull. 
Biol.  Fr.  Belg.  3:  219-286. 

Duelli,  P.  1980.  Preovipository  migration  flights  in  the  green  lacewing,  Chryso- 
pa carnea.  Behav.  Ecol.  Sociobiol.  7:  239-246. 

Eichele,  G.  and  W.  Villiger.  Untersuchungen  an  den  Stridulationsorganen  der 
Florfliege  Chrysopa  carnea  Steph.  (Neuroptera:  Chrysopidae).  Int.  J.  Insect 
Morphol.  Embryol.  3(1):  41-46. 

Henry,  C.  S.  1979.  Acoustical  communication  during  courtship  and  mating  in 
the  green  lacewing,  Chrysopa  carnea.  Ann.  Entomol.  Soc.  Amer.  72(1):  68-79. 

Henry,  C.  S.  1980a.  Acoustical  communication  in  Chrysopa  rufilabrus  (Neurop- 
tera: Chrysopidae),  a green  lacewing  with  two  distinct  calls.  Proc.  Entomol.  Soc. 
Wash.  82(1):  1-8. 

Henry,  C.  S.  1980b.  The  courtship  call  of  Chrysopa  downesi  Banks  (Neuroptera: 
Chrysopidae):  Its  evolutionary  significance.  Psyche  86:  291-297. 

Henry,  C.  S.  1980c.  The  importance  of  low-frequency,  substrate-borne  sounds  in 
lacewing  communication  (Neuroptera:  Chrysopidae).  Ann.  Entomol.  Soc.  Amer. 
73(6)  617-621. 

Johnson,  V.  and  W.  P.  Morrison.  1979.  Mating  behavior  of  three  species  of 
Coniopterygidae  (Neuroptera).  Psyche  86:  395-398. 

Miller,  L.  A.  1970.  Structure  of  the  green  lacewing  tympanal  organ  ( Chrysopa 
carnea,  Neuroptera).  J.  Morphol.  131:  359-382. 

Miller,  L.  A.  1971.  Physiological  responses  of  green  lacewings  (Neuroptera: 
Chrysopidae)  to  ultrasound.  J.  Insect  Physiol.  17:  491-506. 

Morris,  G.  K.  1980.  Calling  display  and  mating  behavior  of  Copiphora  rhinoc- 
eros Pictet  (Orthoptera:  Tettigoniidae).  Anim.  Behav.  28:  42-51. 

Pearman,  J.  V.  1928.  On  sound  production  in  the  Psocoptera  and  on  a presumed 
stridulatory  organ.  Entomol.  Monthly  Mag.  64:  8-11. 

Pringle,  J.  W.  S.  1948.  The  gyroscopic  mechanism  of  the  halteres  in  Diptera. 
Phil.  Trans.  Roy.  Soc.  B.  233:  347-384. 

Riek,  E.  F.  1967.  Structures  of  unknown,  possible  stridulatory,  function  on  the 
wings  and  body  of  Neuroptera;  with  an  appendix  on  other  endopterygote 
orders.  Austr.  J.  Zool.  15:  337-348. 

Rupprecht,  R.  1968.  Das  Trommeln  der  Plecopteren.  Z.  Vergl.  Physiol.  59: 
38-71. 

Rupprecht,  R.  1974.  Vibrationssignale  bei  der  Paarung  von  Panorpa.  Experien- 
tia  30:  340-341. 

Rupprecht,  R.  1975.  Die  Kommunikation  von  Sialis  (Megaloptera)  durch  Vi- 
brationssignale. J.  Insect  Physiol.  21:  305-320. 

Sheldon,  J.  K.  and  E.  G.  MacLeod.  1974.  Studies  on  the  Biology  of  the  Chry- 
sopidae. IV.  A field  and  laboratory  study  of  the  seasonal  cycle  of  Chrysopa 
carnea  Steph.  in  Central  Illinois.  Trans.  Amer.  Entomol.  Soc.  100:  437-512. 

Smith,  R.  C.  1922.  The  biology  of  the  Chrysopidae.  Mem.  Cornell  Univ.  Agr. 
Exp.  Sta.  58:  1291-1372. 


1981] 


Duelli  & Johnson  — Green  Lacewings 


381 


Tauber,  C.  A.  1969.  Taxonomy  and  biology  of  the  lacewing  genus  Meleoma 
(Neuroptera:  Chrysopidae).  Univ.  Calif.  Publ.  Ent.  58:  1-62. 

Taylor,  C.  P.  1981.  Contribution  of  compound  eyes  and  ocelli  to  the  steering  of 
locusts  in  flight.  I.  Behavioral  analysis.  J.  Exp.  Biol.  93:  1-18. 

Toschi,  C.  A.  1965.  The  taxonomy,  life  histories,  and  mating  behavior  of  the 
green  lacewings  of  Strawberry  Canyon.  Hilgardia  36:  391-431. 


ARTHROPODS  ATTRACTED  TO  LUMINOUS  FUNGI 


By  John  Sivinski 

Department  of  Entomology  and  Nematology 
University  of  Florida 
Gainesville,  Florida  32611 

Some  fungi  emit  light.  Luminescence  may  be  present  in  mycelia 
[e.g.  a number  of  Mycena  species  (Wassink  1978)]  or  in  both 
mycelia  and  fruiting  bodies  [e.g.  North  American  populations  of 
Panellus  (=  Panus ) stypticus,  Buller  1924].  Lights  have  been 
described  as  blue,  white,  or  green  depending  on  the  species  (Buller 
1924,  Wassink  1978).  Emission  intensities  vary  considerably.  In  the 
forests  of  Borneo  Mycena  (=  Poromycena ) manipularis  are  visible 
at  ca.  40  meters  (Zahl  1971).  An  Australian  species  1 “pours  forth  its 
emerald  green  light”  with  sufficient  intensity  to  read  by  (Lauterer 
1900  in  Buller  1924).  North  American  forms,  such  as  examined  here, 
tend  to  be  dimmer.  The  eye  often  requires  several  minutes  of  dark 
adaptation  before  their  glows  become  visible. 

The  receiver(s)  toward  which  fungi  direct  their  luminous  signals 
are  unknown.  Lights  have  been  supposed  to  lure  spore  dispersing 
insects  (Ewart  1906),  but  such  an  argument  fails  to  account  for 
mycelial  lights  (Ramsbottom  1953).  There  has  apparently  been  no 
conjecture  on  the  benefits  mycelia  accrue  by  glowing.  The  different 
environments  of  mycelia  and  fruiting  bodies  make  it  questionable 
whether  their  lights  are  directed  at  identical  receivers  or  even  serve 
similar  functions. 

Until  this  time  any  proposed  reactions  of  animals  to  fungal  lights 
have  been  speculative.  I here  present  evidence  that  certain  arthro- 
pods are  more  likely  to  be  captured  in  traps  baited  with  light- 
emitting  mycelia  and  fruiting  bodies  than  in  controls  containing 
fungus-free  substrate  or  dead  and  dark  specimens  of  luminous 
species.  Several  possible  interactions  between  fungi  and  attracted 
arthropods  are  discussed. 


■Described  as  Panus  incandescens,  a name  of  doubtful  taxonomic  value  (see 
Wassink  1978). 

Manuscript  received  by  editor  September  29,  1981. 


383 


384 


Psyche 


[Vol.  88 


Methods 

Test  tubes  (10  X 75  mm)  were  covered  with  Tack  Trap®,  a sticky 
trapping  compound,  and  capped  with  a cork.  Luminous  twigs, 
conifer  needles  and  leaf  fragments,  covered  with  mycelia  of  a 
Mycena  sp.,  were  put  into  31  such  tubes.  An  identical  number  of 
control  tubes  contained  similar  but  nonluminous  forest  litter.  Tubes 
with  glowing  fungi  were  placed  as  closely  as  possible  to  the  original 
position  of  their  contents  (note  that  mycelia  are  most  abundant  deep 
in  litter,  but  traps  were  placed  on  the  litter  surface).  Controls  were 
set  ca.  80  mm  to  the  side.  Glass  screw  top  vials  (14  X 40  mm)  were 
also  coated  with  Tack  Trap®.  From  3-6  fruiting  bodies  of  the 
luminous  mushroom  Dictyopanus  pusillus  were  put  into  72  such 
vials.  An  identical  number  of  controls  contained  3-6  D.  pusillus, 
killed  and  rendered  nonluminous  by  bathing  in  alcohol.  Luminous 
and  control  vials  were  alternately  placed,  ca.  80  mm  apart,  on  and 
by  rotting  logs  on  which  D.  pusillus  had  been  found.  Traps  were  put 
out  at  night,  gathered  the  following  morning,  and  arthropods  stuck 
on  their  surfaces  removed. 

All  specimens  were  captured  during  August  in  Alachua  County, 
Florida. 


Results 

More  arthropods  were  captured  on  traps  baited  with  glowing 
fungal  mycelia  ( Mycena  sp.)  and  luminous  fruiting  bodies  ( D . 
pusillus ) than  their  respective  controls  (x2  = 10.14,  p<  .001;  \ = 
6.41,  p<  .01,  see  Table  1).  Taxa  significantly  more  abundant  on 
luminous  traps  in  the  summed  samples  are  Collembola  (x2  — 12.81, 
p < .001),  and  Diptera  = 5.54,  p < .025).  It  is  of  interest  that 
Collembola  are  not  attracted  to  the  bioluminescence  of  a sedentary 
luminous  predator,  larvae  of  the  fungus  gnat  Orfelia  fultoni  (Sivin- 
ski  1982).  Predators,  i.e.  spiders,  ants,  earwigs  occur  in  a luminous: 
dark  ratio  that  borders  on  significance  (x2  = 3.76,  p < .10).  Groups 
captured  in  statistically  indistinguishable  numbers  on  luminous  and 
control  traps  are  Isopods  (x2  — 0.78,  p > .25)  and  Amphipods  (x2  = 
0.59,  p > .25).  An  unusual  set  of  captures  is  the  5 crickets, 
Eunemobius  carolinus,  taken  only  with  luminous  mycelia. 


1981] 


Sivinski  — Arthropods  and  Luminous  Fungi 


385 


Table  1.  The  numbers  of  Arthropods  captured  on  traps  containing  luminous 
mycelia  ( Mycena  sp.),  luminous  fruiting  bodies  ( Dictyopanus  pusillus),  and  their 
respective  controls. 


Mycena 

sp. 

Control 

D. 

pusillus 

Control 

Summed 

Fungi 

Summed 

Control 

Collembola 

22 

8 

31 

14 

53 

22 

Isotomidae/ 

Entomobryidae 

21 

8 

12 

7 

32 

11 

Sminthuridae 

1 

0 

19 

7 

20 

7 

Diptera 

8 

2 

11 

5 

19 

7 

Phoridae 

2 

1 

7 

2 

8 

3 

Sphaeroceridae 

0 

0 

1 

0 

1 

0 

Cecidomyiidae 

5 

0 

2 

3 

7 

3 

Ceratopogonidae 

1 

0 

0 

0 

1 

0 

Psychodidae 

0 

1 

0 

0 

0 

1 

Mycetophilidae 

0 

0 

1 

0 

1 

0 

Predators 

12 

4 

17 

12 

29 

16 

Araneida 

3 

1 

7 

4 

10 

5 

Formicidae 

9 

1 

9 

8 

18 

9 

Carabidae 

0 

1 

0 

0 

0 

1 

Dermaptera 

0 

0 

1 

0 

1 

0 

Hymenoptera 

3 

1 

1 

2 

4 

3 

Isopods 

32 

29 

37 

30 

69 

59 

Amphipods 

0 

1 

9 

7 

9 

8 

Acari 

0 

1 

1 

0 

1 

1 

Orthoptera 

8 

1 

2 

4 

10 

5 

Gryllidae 

5 

0 

0 

0 

5 

0 

Blattellidae 

3 

1 

2 

4 

5 

5 

Cicadellidae 

1 

0 

1 

0 

2 

0 

Thysanoptera 

0 

0 

1 

1 

1 

1 

Unidentified 

0 

2 

2 

3 

2 

5 

All  Arthropods 

86 

49 

113 

78 

199 

127 

386 


Psyche 


[Vol.  88 


Discussion 

Attraction  of  insects  to  fungal  lights  does  not  demonstrate  that 
luring  arthropods  is  the  function  of  the  bioluminescence.  With  this 
caveat  in  mind,  note  that  an  acceleration  in  the  rate  of  certain 
fungus /insect  interactions  even  as  an  effect  of  a bioluminescent 
signal  is  apt  to  influence  the  evolution  of  luminous  fungi.  In 
particular,  the  argument  that  fungal  lights  are  functionless,  and  by 
implication  harmless  by-products  of  metabolism,  loses  force  (see 
also  Lloyd  1977).  Bearing  a light  near  arthropods  is  unlikely  to  be 
selectively  neutral  (for  counterviews,  see  Buller  1924;  Prosser  and 
Brown  1961). 

Some  possible  functions  of  fungal  glows  become  more  plausible 
with,  or  fail  to  find  support  in,  the  presented  data.  Both  are 
discussed  below.2 

Attraction  of  spore  dispersers:  Stinkhorn  fungi  (Phallales)  use 
odor,  and  perhaps  color,  to  attract  spore  dispersing  insects.  Diptera, 
in  particular,  consume  a sweet  malodorous  spore-containing  mu- 
cous smeared  on  the  fungal  surface.  Spores  develop  after  being 
discharged  in  the  insect  feces  (discussed  in  Ramsbottom  1953).  An 
early  conjecture  on  the  function  of  fruiting  body  luminescence  was 
that  lights,  like  odor  and  color  in  stinkhorns,  lure  spore  dispersers 
(Ewart  1906;  see  also  Lloyd  1974,  1977). 3 

A large  proportion  of  the  animals  attracted  to  luminous  fungi  are 
potential  consumers  of  its  spores.  Many  Collembola  feed  on  fungal 
spores,  mycelia,  and  fruiting  bodies.  Some  members  of  captured 
Diptera  families  breed  in  fungi.  The  phorid  Megaselia  halterata,  for 
instance,  is  a pest  of  cultivated  mushrooms  (Oldroyd  1964).  Whether 
spores  of  D.  pusillus  pass  unharmed  through  the  insect  gut  is 


2The  following  functions  concern  heterospecific  receivers;  however,  biolumines- 
cence is  often  intimately  associated  with  mating  (see  Lloyd  1977).  Sexual  congress  in 
relevant  Basidiomycetes  consists  of  exchange  of  nuclei  between  haploid  mycelia.  Is  it 
possible  that  glows  might  direct  the  growth  of  photo-sensitive  hyphae  at  this  stage 
and  so  serve  as  mating  signals?  Such  an  explanation  fails  to  account  for  luminosity  in 
diploid  mycelia  or  the  fruiting  body. 

3Insects  may  evolve  an  affinity  for  fungal  lights  due  to  “rewards,”  in  food,  shelter, 
etc.,  the  fungus  provides.  An  alternative  is  that  attraction  is  due  to  fungal 
exploitation  of  arthropod  “phototropisms.”  The  function  of  “phototropisms”  are 
often  obscure.  Some  are  apparently  effects  of  orientation  systems  based  on  the 
relative  position  of  celestial  objects  (see  Lloyd  1977). 


1981] 


Sivinski  — Arthropods  and  Luminous  Fungi 


387 


unknown.  Nor  is  it  known  if  attracted  flies,  such  as  phorids  and 
cecidomyiids,  would  be  useful  agents  of  dispersal.  Vagile  adults  may 
not  feed  on  fungal  materials.  Protein  consumption  by  cecidomyiids 
is  particularly  rare  (see  Sivinski  and  Stowe  1981).  Spores  may  be 
moved,  however,  by  attachment  to  the  surface  of  a passing  insect. 

The  topography  and  timing  of  luminous  displays  are  often 
suggestive  of  guiding  dispersers.  In  Mycena  pruinosa-viscida  and 
M.  rorida  from  the  Far  Eastern  tropics  only  the  spores  emit  light 
(Haneda  1955).  Most  fruiting  body  lights  are  restricted  to,  or 
brighter  in,  the  spore  bearing  hymenium  (Wassink  1978)  and 
Panellus  stypticus  glows  most  strongly  at  the  time  of  spore  matura- 
tion (Buller  1924).  Conscription  of  dispersal  agents  is  less  likely  to 
account  for  light-emitting  mycelia,  unless  mycelial  cells  pass  safely 
through  the  gut  or  can  be  carried  to  new  locations  on  an  arthropod’s 
exoskeleton. 

Attraction  of  carnivores:  Predaceous  arthropods  were  found  on 
glowing  traps  in  numbers  that  border  on  significance,  and  fungus/ 
predator  interactions  can  be  imagined  as  important  in  the  evolution 
of  bioluminescence.  Luminous  fungi  might  concentrate  carnivores 
about  them  by  exploiting  their  “phototropisms.”  If  predators  arrive 
at  rates  effectively  greater  than  lured  fungivores,  the  resulting 
predator:prey  ratio  may  favor  the  fungus  (an  argument  similar  to 
but  more  evolutionarily  feasible  than  the  “burglar  alarm”  theory  of 
Dinoflagellate  luminescence;  Burkenroad  1943;  see  Buck  1978). 
Such  an  advantageous  ratio  is  not  obvious  in  my  sample.  Alterna- 
tively, carnivores  could  seek  out  luminous  fungi  as  locales  of  high 
prey  density.  Glowing  mushrooms  might  be  mistaken  for  lumines- 
cent animal  prey. 

Attraction  of fungivores:  If  luminous  mycelia  are  unpalatable,  or 
otherwise  difficult  to  ingest,  then  fungivores  attracted  to  lights 
might  consume  adjacent  competitors. 

Attraction  of  fertilizers:  Lloyd  (1974)  suggests  that  arthropods 
lured  by  luminescent  fungus  might  excrete  beneficial  materials  and 
so  aid  growth.  Any  nutritional  gain  must  be  balanced  by  the 
metabolic  expense  of  the  signal. 

Repulsion  of  negatively  phototropic  fungivores:  Bioluminescence 
might  repel  an  organism’s  negatively  phototropic  enemies  or  com- 
petitors (Nicol  1962;  see  also  Sivinski  1981  and  citations).  Repulsion 
is  particularly  plausible  in  explaining  luminous  mycelia,  some  of 


388 


Psyche 


[Vol.  88 


which  occur  buried  in  litter,  inside  rotting  logs,  or  on  roots  deep 
underground  where  the  opacity  of  the  environment  precludes 
attraction  as  a function  of  light. 

Among  surface  dwelling  arthropods,  there  is  no  indication  of  a 
light-avoiding  taxon.  This  does  not  preclude  repulsion.  A rare,  but 
dangerous,  enemy  could  keep  fungal  lights  burning  but  escape 
inclusion  in  the  present  sample,  especially  since  mycelia  baited  traps 
were  not  placed  in  the  area  of  greatest  mycelial  abundance,  deep  in 
the  leaf  litter.  The  intended  receiver  may  not  be  an  arthropod  or 
even  macroscopic.  Protozoa  sometimes  respond  to  lights.  A glow 
could  repel  certain  pathogens  and  keep  the  fungus  free  of  particular 
diseases. 

Light  as  a warning  signal:  Lights  emitted  by  unpalatable  fungi 
might  serve  as  warning  signals  directed  towards  nocturnal  fungi- 
vores  (a  similar  function  has  been  hypothesized  for  ancestral 
flowers,  Hinton  1973).  Of  North  American  fungi  with  luminous 
fruiting  bodies,  one,  P.  stypticus,  is  a bitter  tasting  purgative,  while 
another,  Omphalotus  olearius,  is  a toxic  hallucinogen  (Miller  1979; 
the  palatability  of  D.  pusillus  is  unknown).  Pleurotus  japonicus,  a 
luminescent  Japanese  species,  is  deadly  poisonous  (Buller  1924). 
However,  the  luminous  fruiting  bodies  of  Malaysian  Mycena 
manipularis  are  quickly  attacked  by  fungus  gnats  (Corner  1954; 
gnats  could  be  specialists,  immune  to  toxins).  Again  there  is  no 
evidence  of  arthropods  avoiding  fungal  lights.  My  traps,  of  course, 
would  fail  to  quantify  the  discouragement  of  deer  or  other  large 
fungivores. 

Like  aposematic  insects,  luminous  mushrooms  often  occur  in 
clumps  (kin  groups?)  (see  illustrations  in  Buller  1924,  Harvey  1957; 
also  descriptions  in  Wassink  1978).  Aggregations  might  intensify 
warning  signals  (Cott  1957)  and  be  instrumental  in  the  evolution  of 
conspicuousness  (Fisher  1930,  for  arguments  concerning  the  kin 
selection  of  aposematism).  Several  tropical  light  emitters,  however, 
apparently  occur  singly  (see  Wassink  1978). 

White  fungi  can  reflect  enough  celestial  light  to  be  surprisingly 
obvious  at  night  (noticed  at  twilight  by  Lloyd  1977).  An  assumption 
of  similar  receivers  for  the  bright  white  and  luminous  signals  of 
fruiting  bodies  allows  the  nocturnal  aposematic  signal  hypothesis  to 
be  tested  with  a larger  sample.  Mushrooms  that  appear  to  me  to  be 
uniformly  bright  white  include  6 toxic  species,  13  edible  and  5 whose 


1981] 


Sivinski  — Arthropods  and  Luminous  Fungi 


389 


palatability  is  unknown  (color  and  palatabilities  from  photos  and 
text  of  Miller  1979).  This  distribution  does  not  support  the  aposem- 
atism  argument  (in  comparison  with  a random  sample  of  41  non- 
poisonous  and  9 poisonous  species  x2  = 0.80  p > .25). 

Summary 

Arthropods,  principally  Collembola  and  Diptera,  are  attracted  to 
the  lights  of  luminous  fungal  mycelia  ( Mycena  sp .)  and  fruiting 
bodies  ( Dictyopanus  pusillus ).  Such  attraction  does  not  prove  that 
bioluminescence  has  evolved  to  lure  insects  but  does  affect  the 
plausibility  of  hypotheses  concerning  the  function  of  fungal  glows. 
The  possibilities  of  lights  being  used  to  lure  spore  dispersers,  attract 
consumers  of  fungivores  and  competing  fungi,  repel  negatively 
phototropic  fungivores,  and  serve  as  warning  signals,  are  discussed. 

Acknowledgments 

Comments  by  J.  E.  Lloyd,  T.  J.  Walker,  T.  Forrest,  S.  Wing,  and 
P.  Sivinski  improved  the  paper.  B.  Hollien  professionally  prepared 
the  manuscript.  Dr.  J.  W.  Kimbrough  identified  D.  pusillus  and  J. 
Sivinski  helped  gather  specimens.  Florida  Agricultural  Journal 
Series  No.  3280. 


Literature  Cited 


Buck,  J. 

1978.  Functions  and  evolutions  of  bioluminescence.  Pages  419-460  in  P.  J. 
Herring  (ed.).  Bioluminescence  in  action.  Academic  Press,  New  York. 
Buller,  A.  H. 

1924.  The  bioluminescence  of  Panus  stypticus.  Researches  on  fungi.  Vol.  3.  pp. 
357-431. 

Burkenroad,  M.  D. 

1943.  A possible  function  of  bioluminescence.  J.  Mar.  Res.  5:  161  164. 
Corner,  E.  J.  H. 

1974.  Further  descriptions  of  luminous  agarics.  Trans.  Br.  Mycol.  Soc.  37: 
256-271. 

Cott,  H.  B. 

1957.  Adaptive  coloration  in  animals.  Methuen  and  Co.,  Inc.,  London. 
Ewart,  A.  J. 

1906.  Note  on  the  phosphorescence  of  Agaricus  ( Pleurotus ) candenscens.  Bull. 
Viet.  Nat.  23:  174. 

Fisher,  R.  A. 

1930.  The  genetical  theory  of  natural  selection.  Oxford  Univ.  Press,  London. 


390 


Psyche 


[Vol.  88 


Haneda,  Y. 

1955.  Luminous  organisms  of  Japan  and  the  Far  East.  Pages  335-386  in  F.  H. 
Johnson  (ed.).  The  luminescence  of  biological  systems.  Am.  Assoc.  Adv. 
Sci.,  Wash.,  D.  C. 

Harvey,  E.  N. 

1952.  Bioluminescence.  Academic  Press,  New  York. 

Hinton,  H.  E. 

1973.  Natural  deception.  Pages  96-159  in  R.  L.  Gregory  and  E.  H.  H. 
Gombrich  (eds.).  Illusion  in  nature  and  art.  Charles  Scribner’s  Sons, 
New  York. 

Lloyd,  J.  E. 

1974.  Bioluminescent  communication  between  fungi  and  insects.  Fla.  Ento- 
mol.  57:  90 

1977.  Bioluminescence  and  communication.  Pages  164-183  in  T.  A.  Sebeok 
(ed.).  How  animals  communicate.  Ind.  Univ.  Press,  Bloomington,  Ind. 

Miller,  O.  K.,  Jr. 

1979.  Mushrooms  of  North  America.  E.  P.  Dutton,  New  York,  N.  Y. 
Nicol,  J.  A.  C. 

1962.  Animal  luminescence.  Adv.  Comp.  Physiol.  Biochem.  1:  217-273. 
Oldroyd,  H. 

1964.  The  natural  history  of  flies.  W.  W.  Norton  and  Co.,  Inc.,  New  York. 
Prosser,  C.  L.,  and  F.  A.  Brown,  Jr. 

1961.  Comparative  animal  physiology.  W.  B.  Saunders  Co.,  Philadelphia,  Pa. 
Ramsbottom,  J. 

1953.  Mushrooms  and  toadstools.  Collins,  London. 

SlVINSKI,  J. 

1981.  The  nature  and  possible  functions  of  bioluminescence  in  Coleoptera 
larvae.  Coleopts.  Bull,  (in  press) 

1982.  Prey  attraction  by  luminous  larvae  of  the  fungus  gnat  Orfelia  fultoni. 
(submitted) 

Sivinski,  J.,  and  M.  Stowe. 

1981.  A kleptoparasitic  cecidomyiid  and  other  flies  associated  with  spiders. 
Psyche.  87:  337-348. 

Wassink,  E.  C. 

1978.  Luminescence  in  fungi.  Pages  1 71-197  in  P.  J.  Herring  (ed.).  Biolumines- 
cence in  action.  Academic  Press,  New  York. 

Zahl,  P.  A. 

1971.  The  secrets  of  nature’s  night  lights.  Natl.  Geogr.  140:  45-70. 


PSYCHE 

INDEX  TO  VOLUME  88,  1981 


INDEX  TO  AUTHORS 

Annette,  Aiello.  Life  History  of  Antaeotricha  sp.  (Lepidoptera:  Oecophoridae: 
Stenomatinae)  in  Panama.  163 

Carlin,  Norman  F.  Polymorphism  and  Division  of  Labor  in  the  Dacetine  Ant  Orec- 
tognathus  versicolor  (Hymenoptera:  Formicidae).  231 
Carpenter,  Frank  M.  Dedication:  Robert  E.  Silberglied.  197 
Dahlstrom,  Tina.  See  Topoff,  Howard. 

Droual,  Robert  and  Howard  Topoff.  The  Emigration  Behavior  of  Two  Species  of  the 
Genus  Pheidole  (Hymenoptera:  Formicidea).  135 
Duelli,  Peter  and  James  B.  Johnson.  Behavioral  Origin  of  Tremulation,  and  Possible 
Stridulation,  in  Green  Lacewings  (Neuroptera:  Chrysopidae).  375 
Edwards,  G.  B.  Sound  Production  by  Courting  Males  of  Phidippus  mystaceus  (Ara- 
neae:  Salticidae).  199 

Ginsberg,  Howard  S.  Historical  Development  of  Bee  Foraging  Patterns  in  Central 
New  York  State.  337 

Hathaway,  Mary.  Polistes  gallicus  in  Massachusetts  (Hymenoptera:  Vespidae).  163 
Haverty,  Michael  I.  See  Howard,  Ralph  W. 

Holldobler,  Bert.  Trail  Communication  of  the  Dacetine  Ant  Orectognathus  versi- 
color (Hymenoptera:  Formicidae)  245 

Holldobler,  B.,  M.  Moglich,  and  U.  Maschwitz.  Myrmecophile  Relationship  of  Pella 
(Coleoptera:  Staphylinidae)  to  Lasius  fuliginosus  (Hymenoptera:  Formicidae). 
347 

Howard,  Ralph  W.,  Eldon  J.  Mallette,  Michael  /.  Haverty,  and  Richard  V.  Smythe. 
Laboratory  Evaluation  of  Within-Species,  Between-Species,  and  Parthenogenetic 
Reproduction  in  Reticulitermes  flavipes  and  Reticulitermes  virginicus.  75 
Johnson,  James  B.  See  Duelli,  Peter. 

Kearns,  R.  S.  and  R.  T.  Yamamoto.  Maternal  Behavior  and  Alarm  Response  in  the 
Eggplant  Lace  Bug,  Gargaphia  solani  (Heidemann)  (Tingitidae:  Heteroptera). 
215 

Levings,  Sally  C.  and  James  F.  A.  Traniello.  Territoriality,  Nest  Dispersion,  and 
Community  Structure  in  Ants.  265 

MacKay,  William  P.  A Comparison  of  the  Nest  Phenologies  of  Three  Species  of  Po- 
gonomyrmex  Harvester  Ants  (Hymenoptera:  Formicidae).  25 
Mallette,  Eldon  J.  See  Howard,  Ralph  W. 

Maschwitz,  U.  See  Holldobler,  B. 

Moglich,  M.  See  Holldobler,  B. 


391 


Newton,  Alfred  F.  Jr.  New  Name  for  the  Extinct  Genus  Mesagyrtes  Ponomarenko 
(Coleoptera:  Silphidae  S.L.).  335 

Nimmo,  Andrew  P.  Francis  Walker  Types  of,  and  New  Synonymies  for,  North  Amer- 
ican Hydropsyche  species  (Trichoptera:  Hydropsychidae).  259 
Parker,  F.  D.  See  Tepedino,  V.  J. 

Pujdak,  Susan.  See  Topoff  Howard. 

Richman,  David  B.  and  Willard  H.  Whitcomb.  The  Ontogeny  of  Lyssomanes  viridis 
(Walckenaer)  (Araneae:  Salticidae).  127 
Rothstein,  Aaron.  See  Topoff,  Howard. 

Sivinski,  John.  Arthropods  Attracted  to  Luminous  Fungi.  383 
Smythe,  Richard  V.  See  Howard,  Ralph  W. 

Steiner,  A.  L.  Anti-predator  Strategies.  II.  Grasshoppers  (Orthoptera,  Acrididae) 
Attacked  by  Prionyx  parkeri  and  Some  Tachyspex  Wasps  (Hymenoptera,  Sphe- 
cinae  and  Larrinae):  A Descriptive  Study.  1 
Stuart,  Robin  J.  Abdominal  Trophallaxis  in  the  Slave-Making  Ant,  Harpagoxenus 
americanus  (Hymenoptera:  Formicidae).  331 
Tepedino,  V.  J.  and  F.  D.  Parker.  The  Effect  of  Flower  Occupancy  on  the  Foraging 
of  Flower-Visiting  Insects.  321 
Topoff,  Howard.  See  Droual,  Robert. 

Topoff,  Howard,  Aaron  Rothstein,  Susan  Pujdak,  and  Tina  Dahlstrom.  Statary  Be- 
havior in  Nomadic  Colonies  of  Army  Ants:  The  Effect  of  Overfeeding.  151 
Traniello,  James  F.  A.  See  Levings,  Sally  C. 

Ward,  Philip  S.  Ecology  and  Life  History  of  the  Rhytidoponera  impressa  Group 
(Hymenoptera:  Formicidae).  I.  Habitats,  Nest  Sites,  and  Foraging  Behavior. 
89 

Ward,  Philip  S.  Ecology  and  Life  History  of  the  Rhytidoponera  Group  (Hymenop- 
tera: Formicidae).  II.  Colony  Origin,  Seasonal  Cycles  and  Reproduction.  109 
Whitcomb,  Willard  H.  See  Richman,  David  B. 

Yamamoto,  R.  T.  See  Kearns,  R.  S. 


392 


INDEX  TO  SUBJECTS 


All  new  genera,  new  species  and  new  names  are  printed  in  capital  type. 


A comparison  of  the  nest  phenologies  of 
three  species  of  Pogonomyrmex  har- 
vester ants,  25 

Abdominal  trophallaxis  in  Harpagoxe- 
nus,  33 1 

Alarm  response  in  Gargaphia,  215 
Antaeotricha,  life  history,  163 
Anti-predator  strategies.  II.  Grasshop- 
pers attacked  by  Prionyx  parkeri  and 
some  Tachysphex  wasps:  a descriptive 
study,  1 

Ants,  territoriality,  nest  dispersion,  and 
community  structure,  265 
Apis  mellifera,  337 
Army  ants,  151 

Arthopods  attracted  to  luminous  fungi, 
383 

Bee  foraging  patterns,  337 
Behavioral  origin  of  tremulation  in  green 
lacewings,  375 
Chrysopidae,  375 
Cicadas,  population  ecology,  175 
Dedication:  Robert  E.  Silberglied,  197 
Ecology  and  life  history  of  the  Rhytido- 
ponera  impressa  group  (Hymenoptera: 
Formicidae).  I.  Habitats,  nest  sites, 
and  foraging  behavior,  89 
Ecology  and  life  history  of  the  Rhytido- 
ponera  impressa  group.  II.  Colony, 
seasonal  cycles,  and  reproduction, 
109 

Effect  of  flower  occupancy  on  the  forag- 
ing of  flower-visiting  insects,  321 
Eggplant  lace  bug,  215 
Emigration  behavior  of  two  species  of 
Pheidole,  135 
Flower-visiting  insects,  321 
Francis  Walker  types  of  Hydropsyche, 
259 

Gargaphia  solani,  215 
Grasshoppers,  1 
Green  lacewings,  375 
Harpagoxenus  americanus,  33 1 


Harvester  ants,  25 

Historical  development  of  bee  foraging 
patterns,  337 

Hydropsyche,  new  synonymies,  259 
Hydropsyche  alternans,  261 
Hydropsyche  confusa,  259 
Hydropsyche  reciproca,  262 
Laboratory  evaluation  of  within-species, 
between-species,  and  parthenogenetic 
reproduction  in  Reticulitermes  flavipes 
and  Reticulitermes  virginicus,  75 
Lasius  fuliginosus,  347 
Life  history  of  Antaeotricha  sp.  in  Pan- 
ama, 163 

Luminous  fungi,  383 

Lyssomanes  viridis,  1 27 

Maternal  behavior  in  Gargaphia,  215 

Mesagyrtes,  335 

Mesecanus,  335 

Myrmecophilic  relationships  of  Pella  to 
Lasius,  347 

Neivamyrmex  nigrescens,  151 
New  name  for  extinct  genus  Mesagyrtes, 
335 

New  Synonymies  for  Hydropsyche,  259 
Notes  on  the  population  of  ecology  of 
cicadas  in  the  Cuesta  Angel  forest  ra- 
vine of  Northeastern  Costa  Rica,  175 
Ontogeny  of  Lyssomanes  viridis,  127 
Orectognathus  versicolor,  231,  245 
Parthenogenetic  reproduction  in  Reticu- 
litermes, 75 
Pella,  347 
Pheidole,  135 
Phidippus  mystaceus,  1 
Pogonomyrmex,  25 
Polistes  gallicus  in  Massachusetts  (Hy- 
menoptera: Vespidae),  169 
Polymorphism  and  division  of  labor  in 
Orectognathus,  23 1 
Prionyx,  1 
Reticulitermes,  75 
Rhytidoponera  impressa,  89,  109 


393 


Silberglied,  Robert  E.,  dedication  to, 
197 

Slave-making  ant,  331 

Sound  production  by  males  of  Phidip- 
pus,  199 

Statary  Behavior  in  nomadic  colonies  of 
army  ants:  the  effect  of  overfeeding, 
151 

Stridulation  in  green  lacewings,  375 

Tachysphex,  1 

The  emigration  behavior  of  two  species 
of  the  genus  Pheidole  (Formicidae: 
Myrmicinae),  135 


The  ontogeny  of  Lyssomanes  viridis 
(Walckenaer)  (Araneae:  Salticidae)  on 
Magnolia  grandiflora,  127 
Territoriality,  nest  dispersion,  and  com- 
munity structure  in  ants,  265 
Trail  communication  in  Orectognathus, 
231 

Tremulation  in  green  lacewings,  375 
Trophallaxis  in  Harpagoxenus,  331 
Walker  types  of  Hydropsyche,  259 
Wasps,  1 


394 


CAMBRIDGE  ENTOMOLOGICAL  CLUB 


A regular  meeting  of  the  Club  is  held  on  the  second  Tuesday 
of  each  month  October  through  May  at  7:30  p.m.  in  Room  154, 
Biological  Laboratories,  Divinity  Avenue,  Cambridge.  Entomolo- 
gists visiting  the  vicinity  are  cordially  invited  to  attend. 


BACK  VOLUMES  OF  PSYCHE 

Requests  for  information  about  back  volumes  of  Psyche  should 
be  sent  directly  to  the  editor. 


F.  M.  Carpenter 
Editorial  Office,  Psyche 
16  Divinity  Avenue 
Cambridge,  Mass.  02138 


FOR  SALE 

Reprints  of  articles  by  W.  M.  Wheeler 

The  Cambridge  Entomological  Club  has  for  sale  numerous  reprints 
of  Dr.  Wheeler’s  articles  that  were  filed  in  his  office  at  Harvard 
University  at  the  time  of  his  death  in  1937.  Included  are  about 
12,700  individual  reprints  of  250  publications.  The  cost  of  the 
reprints  has  been  set  at  5c  a page,  including  postage;  for  orders 
under  $5  there  will  be  an  additional  handling  charge  of  50c.  A list  of 
the  reprints  is  available  for  $1.00  from  the  W.  M.  Wheeler  Reprint 
Committee,  Cambridge  Entomological  Club,  16  Divinity  Avenue, 
Cambridge,  Mass.  02138.  Checks  should  be  made  payable  to  the 
Cambridge  Entomological  Club. 


c/> 


O 

UIISNI^NVINOSHIMS  S3  ! H VB  8 ll^ll  B R AR  I ES  SMITHSONIAN^  INST  i TUT  1 01 
r-  v 2 r-  2 r~ 


m v -V??''  z:  m 

ARIES  SMITHSONIAN  INSTITUTION^ NOlinillSN I NVIN0SH1IWS  S3l8Vyai 

2 \ C/>  2 CO 


I 
C/) 

o 
z , 

> s ' w > *'  2 

2 co  **v  z oo 

UIJ.SNI_NVINOSHJ.INS  S3UJVU9I1  LIBRARIES  SMITHSONIAN  INSTITUTIOI 

Z \ £ 2 ^ 


X/  — 

AS£i/  o o Xi'VosvV^x  _ o 

^ Z „j  Z 

ARIES  SMITHSONIAN  INSTITUTION  NOIlfUllSNI  NVINOSHIIWS  S3  I 8V88I 
2:  t“  2 r-  2 

^5ov7x  0 . m O Z . o 


in 


m 

nilJLSNl  NVINOSHimS  S3  I avyairLIBRAR  I ES  ^SMITHSON  IAN  INSTITUTE 
z ^ £ z to  2 .v 

< E - .vW  vv 

f i 5 

I AR  I ES^SMITHSONIAn"  INSTITUTION^NOIJ-nillSNI  NVINOSHJIWS^SS  IBVaa 

00  5 ___ < n — 00 

ItJ  7k  CJ  CJ 


nj.ij.sm  nvinqshjjns  saiavaan  libraries  Smithsonian  institute 

r <,  2 r-  z ^ 

ca  x S ^ ® 

Wr'^y  m ^ m N^/vostoS?'  ■ '/.  /'  ''  ^ 

!AR  I ES</>SMITHS0NIAN~INSTITUTI0NtnN0linilJSNI~NVIN0SHlHMS  S3  I BVdai 

^2  ^ . 2 00  2 GO 

- 1 ! jftv  1 


1 

in 

y 1"  2 

OO  '***  Z W 

MwmnOLi  1 iiaic*  oi  1 viwvi  a n i » r>  r%  a n i c o puitupamiam  IMCTITI  IT!  D 


niiism  nvinoshiiws  saiavaan  libraries  Smithsonian  institutio 


SMITHSONIAN  INSTITUTION  NOlinillSN! 


nvinoshiiws  S3  lava  a i 


LIBRARIES  SMITHSONIAN  INSTITUTIO 

rn  ~~ 


TAR  I ES  SMITHSONIAN  INSTITUTION  NOIXfUliSNI  NVINOSH1IWS  S3iaVaai 


RAR  I ES^SMITHSONIAN^  INSTITUTION  ^NOlinillSN  I 


in  m 


.niusNi  NViNosHims  S3 1 y vd a n libraries 


< 

z: 

o 

CO 


S 

to 


NVINOSHIIWS  S3ldVd8l 
co