Vous êtes sur la page 1sur 127

MIDLANDS

 STATE  UNIVERSITY  

DEPARTMENT  OF  APPLIED  BIOSCIENCES  AND  BIOTECHNOLOGY  

LEVEL  1.2  

HABB  103:  PRINCIPLES  OF  ECOLOGY  


 

Lecturer:  J.  Bare  

Module  description  

The  module  includes  four  one-­‐hour  lectures  and  two  three-­‐hour  practicals  per  week,  but  
attendance  may  be  required  at  additional  meetings  for  extra  lectures  or  tests.  The  duration  
of  the  modules  will  be  twelve  weeks.  

Major  Topics  are:  What  is  Ecology?  Ecology  and  levels  of  organization  in  biological  systems:  
individuals,  populations,  communities,  ecosystems,  and  the  biosphere;  distribution  and  
abundance  of  organisms  in  space  and  time:  causes  and  consequences;  populations  growth;  
estimation  of  population  size  and  density;  concepts  of  habitat  and  niche;  competition;  intra-­‐  
and  inter-­‐specific  competition;  relations  among  species:  commensalisms,  mutualism,  
parasitism  and  predation;  population  regulation.  

Grading  Criteria  

In  this  module  the  final  examination  will  constitute  70%  of  the  final  mark.  The  remaining  
30%  will  come  from  continuous  assessment  marks  as  follows:  20%  from  continuous  
assessment  of  practicals  and  10%  from  continuous  assessment  of  tests  and  assignments.  

For  your  theory  continuous  assessment,  you  will  write  two  in-­‐class  tests  (one  midway  
through  the  module  and  the  other  one  towards  the  end  of  the  module)  and  two  
assignments.    

References  

STILING,  P.  D  Ecology  and  Applications  (2nd  Edition)  Prentice-­‐Hall.  

BEGON,  M.,  HARPER,  J.  H.  and  TOWNSEND,  C.  R.  (1990).  Ecology:  Individuals  ,  Populations  
and  Communities  Blackwell  Scientific  Publications,  Oxford.  

BEGON,  M  and  MORTIMER,  M.  (1986).  Population  Ecology:  A  Unified  study  of  Animals  and  
Plants  Blackwell  Scientific  Publications,  Oxford.  
VARLEY,  G.  C.,    GRADWELL,  G.  R.  AND  HASSEL,  M.  P.  (1975).  Insect  Population  Ecology  
Blackwell  Scientific  Publications,  Oxford.  

OSBORNE,  P.  L.  (2000).  Tropical  Ecosystems  and  Ecological  Concepts.  Cambridge  University  
Press.  

BARBOUR,  M.  G.,  BULK,  J.  H  and  PITTS,  W.  D.  (1987).  Terrestrial  Plant  Ecology  (2nd  Edition)  
Benjamin  Cummings,  Menlo  Park,  California.  

CAMPBELL,  N.  A.    (1993).  Biology    (3rd  edition).  Benjamin  Cummings,  Menlo  Park,  California.  

CHAPMAN,  J.  L.  and  REISS,  M.  J.  (1992).  Ecology:  Principles  and  Applications.  Cambridge  
University  Press.  

COLLINVAUX,  P.  A.  (1973).  Introduction  to  Ecology.  John  Wiley  and  Sons,  New  York.  

ENRICH,  P.  R.  and  ROUGHGARDEN,  J.  (1987).  The  Science  of  Ecology,  MacMillan,  New  York.  

KIKAWA,  J.  and  ANDERSON,  D.  J.  (1986).  Community  ecology:  Pattern  and  Process  
Blackwell  Scientific  Publications,  Oxford  

KNOX,  B.,  LADGES,  P.  and    EVANS,  B.  (1994).  Biology  McGraw-­‐Hill  Book  Company,  Sydney,  
New  York,  San  Francisco,  Auckland,  Bogota,  Caracas,  Lisbon,  London  ,  Madrid,  Mexico  City,  
Milan,  Montreal,  New  Dehli,  San  Juan,  Singapore,  Tokyo,  Toronto.  

KREBS,  C.  J.  (1978).  Ecology  (2nd  Edition).  Harper  and  Row,  New  York.  

MACKENZIE,  A.,  BALL,  A.  S.  and  VIRDEE,  S.  R.  (1999).  Instant  Notes  in  Ecology.  Viva  Books  
Private  Limited,  New  Dehli,  Mumbai,  Chenai.  

MCINTOSH,  R.  P.  (1985).  The  Background  of  Ecology:  Concept  and  Ecology,  Cambridge  
University  Press,  Cambridge.  

ODUM,  E.  P.  (1971).  Fundamentals  of  Ecology  (3rd  Edition)  W.  B.  Saunders,  Philadelphia  

ODUM,  E.  P.  (1983).  Basic  Ecology  and  our  Endangered  Life  Support  System.  Sinauer  
Associates  Sunderland,  Massachusettes  

   
Lecture  1:  INTRODUCTION  
Module:  Principles  of  Ecology  
Dept.  of  Applied  Biosciences  and  Biotechnology,  Midlands  State  University  

1.1.   What  is  Ecology?  


1.2.   Ecology  and  levels  of  organisation  
1.3.   Relations  of  ecology  to  other  fields  of  biology  
1.4.   Doing  ecology  

___________________________________________________________________________  

         
INTRODUCTION  
What  is  Ecology?  
The  term  “ecology”  was  coined  by  the  German  zoologist  Ernest  Haeckel  in  1866  to  describe  
a  then  emerging  speciality  area  of  biology.  

Ernest  Haeckel  derived  the  term  ecology  from  the  Greek  word  Oikos  which  originally  
referred    (in  Greek)  to  a  family  household  and  its  daily  operations  and  maintenance;  the  
same  root  (Oikos)  also  gave  rise  to  the  word  “economy”  which  referred  to  organisms  linked  
together  in  conflict  as  well  as  in  mutual  aid:  

Haeckel  elaborated  on  this  in  1869:  “By  ecology,  we  mean  the  body  of  knowledge  
concerning  the  economy  of  nature,  the  investigation  of  the  total  relations  of  the  organism  
(i.e.  animal,  plant,  protistan,  fungus,  bacterium,  etc.)  both  to  its  organic  and  inorganic  
environment;  including  above  all,  its  friendly  and  inimical  relations  with  those  animals  and  
plants  (or  any  other  organisms)  with  which  it  comes  directly  or  indirectly  into  contact.  In  a  
word,  ecology  is  the  study  of  all  the  complex  interrelationships  referred  to  by  Darwin  as  the  
conditions  of  the  struggle  for  existence”.  

From  then  on  ecology  has  been  various  defined:  

1.   Ecology    =  the  study  of  the  interaction  of  organisms  with  their  environments,  both  
abiotic  (physical-­‐chemical)  and  abiotic  (other  organisms).  

The  problem  of  this  definition  is  that  it  is  very  broad  and  inclusive,  and  it  leaves  very  
little  that  is  NOT  ecology,  and  there  is  need  to  narrow  it  down  somewhat.    

2.   Ecology=    scientific  natural  history  (Charles  Elton,  1927,  Animal  Ecology).  

Where:  

Natural  history    =    the  study  of  organisms  in  nature  with  respect  to  all  aspects  of  
their  biology,  from  mating  and  reproduction  to  predator  
thwarting,  diet  and  finding  food,  etc.    

Natural  history  gave  rise  not  only  to  ecology  but  also  to  systematics,  ethology  
(behaviour),  comparative  biology  and  evolutionary  biology.  

Advantage:     points  out  the  origin  of  many  ecological  problems  in  the  ancient  field  
of  natural  history.  

Problem:     uncomfortably  vague.  

3.   Ecology    =    the  study  of  structure  and  function  of  nature  (Eugene  Odum,  1963).  
Advantage:     emphasises  the  form  and  function  idea  that  permeates  biology.  

Problem:     still  not  completely  clear.    

4.   Ecology  =    the  scientific  study  of  the  distribution  and  abundance  of  organisms  (H.  G.  
Andrewartha,  1961,  Introduction  to  the  study  of  Animal  populations).    

Advantage:     clear  and  restrictive  definition.  

Problem:     this  definition  is  static  and  it  leaves  out  the  idea  of  relationships.  

5.   Ecology    =  the  scientific  study  of  the  interactions  that  determine  the  abundance  and  
distribution  of  organisms  (Charles  J.  Krebs,  1972,  Ecology).  

Where:  

Distribution  =    the  geographical  extent  of  a  population  or  any  other  ecological  
unit.  

Abundance=  the  number  or  quantity  of  organisms  in  a  given  area  (density).  

This  definition  brings  out  the  idea  that  ecologists  are  interested  in  knowing  where  
organisms  are  found,  how  many  occur  there,  and  why?  

For  the  purposes  of  this  module  we  are  going  to  use  definition  number  5.  

Ecology  and  levels  of  organisation  in  biological  systems  


We  have  to  recognise  9  different  levels  of  organisation  within  biological  systems:  

Molecule  

Organelle  

Cell  

Individual  

Population  

Community  

Ecosystem  

Biome  

Biosphere  

However,  ecology  is  that  branch  of  biology  that  deals  with  levels  of  organisation  from  the  
individual  level  and  above.  That  is:  
Individual  

Population  

Community  

Ecosystem  

Biome  

Biosphere  

Individual  organism=  this  is  the  basic  biotic  unit  that  has  to  be  studied  in  relation  to  the  
environment.  It  is  easy  to  define  individual  in  some  species  (e.g.  in  humans),  but  more  
difficult  in  colonial  (asexual  reproducing)  species,  especially  if  polyps  maintain  physiological  
contact.)  

Population=  a  group  of  individuals  of  the  same  species  inhabiting  the  same  area  at  the  same  
time.  

Community=  an  association  of  interacting  populations,  usually  defined  by  the  nature  of  
their  interactions  or  the  place  in  which  they  live  (e.g.  herbivore  community,  plankton  
community).  

Ecosystem=  a  self-­‐sufficient  habitat  where  living  organisms  and  the  non-­‐living  environment  
interact  to  exchange  energy  and  matter  in  a  continuing  cycle  (e.g.  forest  ecosystem,  ocean  
ecosystem,  etc).  

Biome=  large  areas  of  approximately  uniform  habitat,  consisting  of  distinctive  combinations  
of  plant  and  animal  species.  

Biosphere=  the  part  of  the  planet  containing  living  organisms,  the  living  world.  

Reductionism=  a  philosophy  which  states  that  the  higher  levels  of  organisation  of  complex  
systems  can  be  fully  explained  through  knowledge  of  the  smallest  components.    

In  other  words,  reductionism  is  the  idea  that  breaking  a  system  down  into  increasingly  
smaller  parts  will  enable  one  to  understand  it  completely.  

However,  biological  systems  are  extremely  complex,  especially  at  the  higher  levels  of  
organisation  (populations,  communities  and  ecosystems),  and  are  subject  to  stochastic  
(chance)  events  that  make  such  reductionism  impossible.  

Emergence=  the  occurrence  of  characters  at  the  higher  levels  of  organisation  which  could  
not  have  been  predicted  from  a  knowledge  of  the  lower  level  components.  

Ecology  of  the  individual  -­‐  deals  with  how  organisms  are  affected  by  (and  how  they  affect)  
their  environment.    
In  other  words,  we  are  not  concerned  so  much  with  the  interaction  between  individuals  and  
their  environment,  but  rather  with  the  numbers  of  individuals  and  the  process  leading  to  
changes  in  the  number  of  individuals.  

Autecology=study  of  the  individual  in  relation  to  the  environmental  conditions  (e.g.  
streamlined  bodies  of  animals  from  flowing  rivers;  floating  seeds  in  island-­‐dwelling  palm  
trees).    

In  other  words,  Autecology  is  the  ecology  of  individual  organisms  or  species.  

Ecology  of  populations  (Population  Ecology)  -­‐  deals  with  the  presence  or  absence  of  
particular  species,  with  their  abundance  or  rarity,  and  with  the  trends  or  fluctuations  in  their  
numbers.  

There  are  two  approaches  at  the  population  level:  

(i)-­‐Deals  first  with  the  attributes  of  individual  organisms,  and  then  considers  the  way  in  
which  these  combine  to  determine  the  characteristics  of  the  population.  This  is  a  
reductionist  approach.  

(ii)-­‐deals  directly  with  the  characteristics  of  populations,  and  tries  to  relate  these  to  aspects  
of  the  environment.  This  is  a  systems  approach  which  allows  for  emergent  properties.    

*historically  tied  to  animal  studies.    

Community  ecology  –  often  considered  together  with  ecosystem  ecology.  

It  deals  with  the  structure  of  communities  (e.g.  how  many  species,  which  species)  and  with  
the  functions  of  communities/ecosystems  (i.e.  pathways  followed  by  energy,  materials,  
nutrients  and  other  chemicals  through  them).  

There  are  two  approaches  at  the  community  level:  

(i)-­‐study  the  component  populations  to  gain  an  understanding  of  these  patterns  and  
processes.  This  is  a  reductionist  approach.  

(ii)-­‐look  directly  at  the  properties  of  populations/ecosystems  themselves  (e.g.  species  
diversity,  rate  of  biomass  production).  This  is  a  systems  approach  which  allows  for  emergent  
properties.    

*historically  tied  to  plant  studies.    

Ecosystem  ecology  –    as  for  community  ecology.  


Synecology  =   studies  groups  of  organisms  in  relation  to  their  environment,  includes  
population,  community,  and  ecosystem  ecology,  i.  e.  the  ecology  of  plant  and  
animal  communities.  

Relations  of  Ecology  to  other  fields  of  biology  


Ecology  tends  to  overlap  with:  

(i)Physiology  

(ii)Behaviour  (Ethology)  

(iii)Genetics  

(iv)Evolution  

Ecology  and  Physiology  

How  do  physiological  mechanisms  act  to  affect  survival  and  reproduction  at  the  population  
level?  

Physiology  (a  morass  of  within-­‐organism  characteristics  which  specify  how  an  individual’s  
life  processes  respond  to  the  environment)  is  very  much  ignored  at  the  population  and  
community  levels  but  provides  a  means  of  coupling  organism’s  response  to  the  
environment.    

–  effects  of  abiotic  factors  (e.g.  temperature,  light,  pressure,  pH,  O2  concentration,  etc.)  on  
organisms,  act  through  physiological  changes.  

–  effects  of  biotic  factors  (e.g.  food,  allelopathic  chemicals,  etc.)  on  organisms,  act  through  
physiological  changes.  

This  can  lead  to  a  subject  called  physiological  ecology  or  environmental  physiology.  

Ecology  and  Behaviour  

How  do  behavioural  mechanisms  act  to  influence  survival  and  reproduction  at  the  
population  level?  

Behavioural  responses  of  organisms  to  the  environment  (e.g.  temperature  preferences,  prey  
location  and  selection,  mate  choice,  sociality,  decisions  to  forage  or  to  look  for  mates,  etc.)  
can  determine  survival  and  reproduction  at  the  population  level.    

–  decisions  on  how  to  spend  your  time  and  energy,  e.g.  to  foraging  for  food  or  looking  also  
affect    survival  and  reproduction  at  the  population  level.  

This  can  lead  to  a  subject  called  behavioural  ecology.  


Ecology  and  Genetics  

–  population  level,  population  genetics  

–  most  characteristics  of  organisms  that  affect  their  interaction  with  the  environment  have  
a  genetic  component  (morphology,  physiology  and  behaviour).    

–  Individuals  in  a  population  differ  genetically  and  this  can  result  in  differences  in  their  
survival  and  reproduction  under  a  given  set  of  environmental  conditions.  Changes  in  the  
genetic  make-­‐up  of  a  population  over  time  results  in  EVOLUTION.  

Ecology  and  Evolution  

Ecological  interactions  can  result  in  changes  in  populations  over  time  (i.e.  evolution).  

This  can  lead  to  a  subject  called  evolutionary  biology  or  evolutionary  ecology.  

According  to  Evelyn  G.  Hutchinson,  there  are  two  causes  for  observable  patterns  or  
relationships,  e.g.  for  the  observable  geographical  range  of  blue  Jays  in  North  America:  

Proximal  (ecological)  cause  may  involve  limits  of  the  bird’s  tolerance  of  the  physical  
environment  (e.g.  temperature,  scarcity  of  food,  abundance  of  parasites  or  predators,  etc.).    

In  other  words,  the  proximal  explanation  looks  at  functions  causing  what  is  happening  here  
and  now.  

Ultimate  (evolutionary)  cause  addresses  how  the  bird  came  to  have  these  properties  in  the  
first  place;  this  is  usually  a  result  of  past  evolutionary  changes.    

In  other  words,  the  ultimate  explanation  of  the  present  distribution  and  abundance  of  this  
bird  lies  in  the  ecological  experiences  of  its  ancestors,  i.e.  the  ultimate  explanation  looks  at  
how  the  current  situation  was  influenced  by  past  events  including  evolution.  

Adaptation  =  a  genetically  determined  characteristic  that  enhances  the  ability  of  an  
individual  to  cope  with  its  environment;  or  an  evolutionary  process  by  which  organisms  
become  better  suited  to  their  environment.  

Doing  Ecology  
What  do  ecologists  do?  

Discover  and  describe  patterns  in  nature  (WHAT  AND  WHERE).  

Explain  observed  patterns  in  nature  (HOW  AND  WHY).  

(a)Proximate  explanation  =     functions  causing  what  is  happening  here  and  now.  
(b)Ultimate  explanation  =     how  the  current  situation  was  influenced  by  past  
events  including  evolution.  

Predict  or  Control  patterns  (APPLIED  ECOLOGY  =  use  of  ecological  principles  and  knowledge  
for  human  benefit)  

There  are  therefore  three  basic  approaches  to  ecological  problems:  

Descriptive  approach  

This  is  where  the  investigator  asks  “what”  questions,  e.g.  what’s  there?  The  investigator  can  
make  an  inventory  and  describes  what  he/she  finds,  e.g.  they  can  describe  the  major  
vegetation  groups  (Tundra,  temperate,  deciduous  forests,  grasslands,  etc.).    This  is  mainly  
natural  history.  

–  Darwin  admonished,  “one  might  as  well  go  into  a  gravel  pit,  count  the  pebbles,  and  
describe  the  colours”.  

Functional  approach  

In  this  approach  the  investigator  asks  “how”  questions,  e.g.  how  does  the  system  operate?    

This  approach  is  oriented  towards  dynamics  and  relationships  studies  and  tries  to  identify  
the  proximate  causes  of  the  dynamic  responses  of  populations  and  communities  to  
immediate  factors  in  the  environment.  

Evolutionary  approach  

The  investigator  asks  “why”  questions,  e.g.  why  does  natural  selection  favour  this  particular  
ecological  solution?  

–  studies  “ultimate”  causes  of  certain  patterns  and  considers  organisms  and  their  
relationships  as  historical  products  of  evolution.  

Methods  in  Ecology  

The  principal  method  used  in  ecology  and  other  sciences  is  the  hypothetico-­‐deductive  
approach.    

You  start  with  a  series  of  observations  which  appear  to  show  a  constant  pattern  or  
relationship,  e.g.  you  note  that  in  Chimanimani,  pine  seedlings  do  not  grow  under  a  canopy  
of  mature  deciduous  trees.  

Based  on  these  observations  you  generate  a  theory,  e.g.  pine  seedlings  cannot  tolerate  low  
light  levels  found  under  such  a  canopy.  
Using  your  theory  as  a  guide  you  can  generate  hypotheses  which  predict  the  outcome  of  
various  experiments  if  your  theory  is  correct,  e.g.:  

●   Over  a  range  of  sites  there  should  be  a  positive  correlation  between  light  
intensity  and  pine  seedling  density.  

●   Pine  seedlings  should  show  inability  to  survive  at  low  light  levels  in  controlled  
laboratory  experiments.  

If  you  artificially  increase  light  intensity  under  a  forest  canopy  in  the  field,  you  should  be  
able  to  see  an  increase  in  pine  seedling  abundance.  

If  your  experiments  all  produced  the  results  you  predicted,  you  can  conclude  that  you  have  
successfully  explained  the  observed  pattern;  if  not,  you  must  change  your  theory,  generate  
new  hypotheses  and  run  new  experiments  to  test  them.  

Deduction  (logic)  =   the  process  of  reasoning  in  which  a  conclusion  follows  necessarily  
from  the  stated  premises;  inference  by  reasoning  from  the  general  to  
the  specific.  

Kinds  of  evidence  used  by  ecologists  


Hypotheses  are  tested  by  collecting  data  in  two  ways:  

Observation  and  Monitoring  of  natural  systems,  e.g.  measuring  the  density  of  Blue  
Jays  nests  in  areas  with  and  without  dogwoods  to  test  the  hypothesis  that  dogwoods  
influence  Blue  Jays  nest  density.  

Advantage  =     very  close  to  natural  conditions  –  results  may  be  real.  

Disadvantage  =    causes  of  results  may  be  difficult  to  identify  because  of  
confounding  factors  (e.g.  areas  studied  may  differ  more  than  
in  just  dogwood  density).  

Experiments  
Types  of  experiments  in  Ecology:  

Field  Experiments  –  manipulation  of  natural  systems  to  test  hypotheses,  e.g.  remove  
dogwoods  (independent  variable)  from  one  area  but  not  from  a  control  area,  then  compare  
Blue  Jay  nest  densities  (dependent  variable)  between  the  two  areas.      

An  effort  must  be  made  to  try  and  vary  only  the  independent  variable  of  interest  and  keep  
all  other  independent  variables  the  same  in  all  treated  areas.  

Advantage  =     can  isolate  and  better  identify  causes.  


Disadvantage  =    treatment  may  have  unintended  effects  (e.g.  dogwood  
removal  may  stimulate  production  of  other  food  items  that  
the  Blue  Jays  may  like).  

*Therefore,  there  is  great  realism  (natural  system)  but  low  confidence  (complex  system  with  
many  uncontrolled  variables).  

Laboratory  experiments  –  manipulation  of  artificial  systems  to  test  hypotheses.  

Advantage  =     highly  controlled  system.  

Disadvantage  =   even  more  unrealistic  than  field  experiments;  we  should  seek  
simplicity,  but  distrust  it.  

*Therefore,  there  is  low  realism  (unnaturally  simple  system)  but  high  confidence  (most  
other  variables  tightly  uncontrolled).  

Theoretical/mathematical  Models  –  an  abstract  representation  of  ecological  interactions.  

Advantage  =     less  costly,  can  answer  many  questions  because  it  is  flexible.  

Disadvantage  =   difficult  to  account  for  complexity.  

*Models  are  only  as  good  as  the  data  and  relationships  used  to  construct  them.  Used  
primarily  when  field  experiments  are  impossible  (too  large  or  too  long,  e.g.  global  warming).    

   
Lecture  2:  POPULATION  STRUCTURES  
Module:  Principles  of  Ecology  
Dept.  of  Applied  Biosciences  and  Biotechnology,  Midlands  State  University  

2.1.   Distributions  of  Populations  


2.2.   Dispersion  –  spacing  between  individuals  
2.3.   Temporal  changes  in  population  size  

___________________________________________________________________________  
POPULATION  STRUCTURES  
The  term  population  is  defined  differently  in  various  sciences:  

-­‐ In  human  demography  a  population  is  a  set  of  humans  in  a  given  area.  
-­‐ In  genetics  a  population  is  a  group  of  inter-­‐breeding  individuals  of  the  same  species,  
which  is  isolated  from  other  groups.  
-­‐ In  ecology  a  population  is  a  group  of  individuals  of  the  same  species  inhabiting  the  
same  area  at  the  same  time.  

A  population  can  be  a  discrete  “natural”  unit,  e.g.  the  Blue  gill  sunfish  of  Conesus  Lake  
in  the  USA,  or  a  population  maybe  arbitrarily  chosen  by  the  investigator,  e.g.  the  
elephant  population  of  Hwange  National  Park.  

Characteristics  of  Populations  

(i) Distribution  -­‐  geographical  range,  or  areal  extent  


(ii) Dispersion  –  spacing  between  individuals  
(iii) Abundance  –  density  in  a  given  area  or  total  number  in  the  population  
(iv) Age  structure  –  relative  or  absolute  numbers  in  different  age  classes  
(v) Genetic  Composition  -­‐    

 We  might  also  want  to  know  the  following  about  the  dynamics  of  the  population  (i.e.  its  
changes  over  time):  

(i) Dispersal  and  migration  –  movement  through  space  


(ii) Population  dynamics  –  change  in  number  (or  density)  over  time  due  to  births,  
deaths,  immigration  or  emigration.  
(iii) Changing  age  structure  (demographics)  
(iv) Changing  genetic  composition  –  evolution  

Distribution  of  populations  

-­‐ Distribution  is  the  geographical  extent  of  a  population  (or  any  other  ecological  unit  
such  as  a  species  or  community).    
-­‐ Biogeography  is  the  scientific  study  of  the  distribution  of  organisms.  

With  respect  to  distribution  we  can  ask  questions  such  as:  

“Why  are  individuals  of  a  given  species  present  in  some  species  but  absent  from  others?”  

“What  determines  (or  limits)  the  geographical  range  of  a  population  or  species?”  

The  distribution  of  a  population  is  determined  by  a  number  of  factors:  
1.   Dispersal  –  if  a  species  is  not  found  in  a  particular  area  we  can  ask  the  question;  is  
the  species  absent  because  its  “propagules”  (seeds,  spores,  larvae,  adults,  etc.)  did  
not  or  could  not  reach  the  area?  –  due  to  some  geographical  barrier  and/or  limited  
dispersal  ability.  

The  question  can  be  answered  by  transplant  experiments.  These  involve  the  
experimenter  moving  some  individuals  to  an  area  where  they  are  not  found  to  see  if  
they  can  survive,  grow  and  reproduce  there,  e.g.  the  sugar  maple  was  transplanted  
to  Europe  and  succeeded  (a  good  habitat),  further  north,  it  failed  (because  it  was  too  
cold),  further  south,  it  failed  (because  it  was  too  hot),  and  further  west,  it  also  failed  
(because  it  was  too  dry).  Such  experiments  can  help  to  determine  the  potential  
range  (ecological  range,  i.e.  all  sites  within  a  species’  fundamental  niche)  of  a  
species.    

2.   Behaviour  (Habitat  selection)  –  in  some  instances  an  organism  or  species  can  survive  
in  an  area  if  introduced  and  restrained  there,  but  does  not  occur  there  naturally  
(even  if  its  propagules  reach  the  area).  This  may  be  due  to  the  fact  that  the  organism  
might  be  limited  by  an  inflexible  behavioural  habitat  preference  –  i.e.  individuals  do  
not  see  any  suitable  habitat,  and  they  continue  to  move  on.  This  behaviour  may  have  
evolved  under  circumstances,  where  it  was  advantageous  (i.e.  had  adaptive  value)  in  
the  old  habitat,  but  is  maladaptive  here.  

3.   Interactions  with  other  species  –  parasites,  diseases,  prey,  competitors,  predators,  


etc.    

A  species  may  suffer  severely  in  increased  mortality  and/or  reduced  fecundity  due  to  
intense  parasitism,  disease  outbreaks,  food  shortages,  competition  or  predation.  

In  order  to  find  out  whether  the  absence  of  a  particular  species  in  a  particular  area  is  
due  to  interactions  with  other  species,  an  investigator  could  perform  transplant  
experiments  that  also  exclude  suspected  interacting  species.  

4.   Physical  and  chemical  factors  –  abiotic  factors  (conditions,  abiotic  resources)  might  
be  outside  the  range  tolerated  by  the  species.    

For  example,  temperatures  may  be  too  high  or  too  low  for  the  species  in  question.  

Two  important  considerations  to  make  regarding  geographical  limitation:  


(1)  Geographical  barriers  –     areas  of  inhospitable  habitat,  in  which  a  species  cannot  
survive,  separating  areas  of  hospitable  habitat  (some  of  
which  are  not  occupied),  e.g.  oceans  for  terrestrial  
species,  land  for  aquatic  species,  deep  water  for  
shallow  water  species,  mountains  for  low-­‐altitude  
species,  valleys  for  high-­‐altitude  species.  

(2)  Dispersal  ability  –     some  species  are  extremely  mobile  (e.g.  birds,  such  as  
the  arctic  tern),  while  others  are  stationary,  at  least  as  
adults  (e.g.  trees).    

The  offspring/dormant  stages  (e.g.  seeds,  spores,  larvae,  etc.)  are  often  the  stage  
that  disperses.  

The  spread  of  a  species  accidentally  introduced  by  man  into  a  region  where  they  
never  previously  occurred  (i.e.  “exotic”  or  “introduced”  species)  illustrates  the  
dispersal  ability  of  a  species,  e.g.:  

(i)   Gypsy  moth  –     introduced  to  America  by  a  French  astronomer  working    
at  Harvad  University  near  Boston  in  1850.  
(ii)   African  honey  bee  –   an  aggressive  subspecies  introduced  to  Brazil  in  1956    
to  improve  honey  productivity.  
(iii)   Colorado  beetle  –     rare  case  of  spread  from  the  “New  World”  to  the  “Old    
World”.  

Dispersion  –  Spacing  of  individuals  (spatial  distribution  of  


organisms)  
Dispersion  =  spatial  pattern  of  distribution  of  individuals  within  populations.  

The  spatial  distribution  of  all  living  organisms  depends  on  a  variety  of  factors  such  as  soil  
type,  moisture  conditions,  temperature  variations,  the  presence  of  neighbours  or  
competitors  and  so  on.  

Distributions  may  change  with  time  as  they  are  affected  by  such  things  as  the  movements,  
births  and  deaths  of  organisms.  Furthermore,  different  life  stages  of  the  same  species  may  
have  different  distributional  patterns.  

The  individuals  of  a  population  can  follow  three  basic  types  of  spatial  distribution:  

(1)  Random  distribution,  in  which  there  is  an  equal  chance  of  an  individual  of  occupying  
any  point  in  an  area  irrespective  of  the  position  of  other  individuals.  A  striking  
feature  of  random  distribution  is  the  lack  of  any  system,  e.g.  some  individuals  occur  
in  groups  and  some  are  equally  spaced,  some  individuals  are  close  together  and  
others  are  wide  apart.  

Mechanism:  chance  events  (e.g.  passive  seed  dispersal  by  wind);  no  interaction  
between/among  individuals.  

(2)   Regular  (or  uniform,  even,  spaced,  or  over  dispersion)  distribution,  which  occurs  
when  distances  between  individuals  are  approximately  equal  –  more  equally  spaced  
than  is  expected  by  chance.  

Mechanism:  either  individuals  avoid  each  other  (e.g.  when  animals  are  relatively  
crowded  they  move  away  from  each  other  and  tend  to  be  the  same  distances  apart)  
or  competition  eliminates  individuals  which  are  too  close  together.  

(3)   Clumped  (aggregated,  contagious,  or  under  dispersion)  distribution,  in  which  
individuals  are  closer  together  than  is  expected  by  chance.  Random  or  regular  
distributions  seldom  occur  in  nature  and  most  populations  occur  in  definite  clumps  
or  patches.  

Mechanism:  this  comes  about  because  most  environmental  factors  either  are  
unevenly  distributed,  i.e.  individuals  are  attracted  to  (or  survive  better  in)  a  given  
area  (i.e.  the  habitat  is  “patchy”),    

individuals  are  attracted  to  each  other  (gregarious  –  for  protection  against  
predation),  

or  

offspring  fail  to  disperse  far  from  parents.  

Temporal  Changes  in  Population  Size  


Very  few,  if  any,  populations  grow  exponentially  for  any  length  of  time  in  nature.  We  can  
only  see  this  in  the  case  of  population  recovery  after  a  disturbance  (storm,  drought,  volcanic  
eruption,  etc.)  or  introduction  of  “exotic”  species.  

The  size  (or  density)  of  a  population  can  vary  temporally  (i.e.  over  time).  For  example  the  
number  of  breeding  pairs  of  the  Grey  Heron  was  seen  in  the  Thames  Valley  to  fluctuate  
during  the  period  1928  –  1970  (fig  1).  

 
 

Fig.  1:     Population  fluctuations  of  the  Grey  Heron  in  the  Thames  individuals  are  attracted  
to  Valley  

The  decrease  in  population  size  occurred  in  the  cold  winters.  The  Grey  Heron  feeds  on  fish  
and  in  cold  winters,  with  the  water  bodies  capped  with  ice,  food  supply  was  limited  and  this  
lowered  the  population  size.    

Populations  can  remain  stable  over  a  period  of  time,  they  can  fluctuate  cyclically  or  they  can  
continue  to  decline  until  they  become  extinct.  

Population  fluctuations  are  caused  by  four  processes  (called  demographic  processes):  

1.   Mortality  –  death  of  individuals  (decreases  population  sizes).  

2.   Natality  (or  fecundity)  –  production  of  new  individuals  (increases  population  sizes).  

3.   Emigration  –  movement  of  individuals  out  of  the  population  (by  migration,  dispersal,  
etc.)  (decreases  population  sizes).  

4.   Immigration  –  movement  of  individuals  into  a  population  (by  migration,  dispersal,  


etc.)  (increases  population  sizes).  

We  can  summarise  how  these  four  processes  affect  population  size  over  time  in  the  
following  equation:  

Nnow  =  Nthen  +  B  –  D  +  I  –  E  
Where:  

Nnow  =  number  (or  density)  at  present  time  

Nthen    =  number  (or  density)  at  some  time  in  the  past  

B  =  number  of  births  over  that  time  

D  =  number  of  deaths  over  that  time  

I  =  number  of  immigrations  over  that  time  


E  =  number  of  emigrations  over  that  time  

Similarly,  we  can  predict  the  future  population  size  (or  density)  as  

Nfuture  =  Nnow  +  B  –  D  +  I  –  E  
or  

Nt+1  =  Nt  +  B  –  D  +  I  -­‐  E  


Here  we  will  consider  only  the  first  two  of  these  processes  –  birth  and  death.  

Natality  (or  fecundity)  –  production  of  new  individuals  by  birth,  hatching,  germination  or  
fission.  

Mortality  –  loss  of  individuals  due  to  their  death  –  can  be  due  to  interactions  with  other  
organisms  (e.g.  diseases,  parasitism,  predation,  starvation,  etc.)  or  loss  of  homeostasis  in  
the  face  of  environmental  challenge  (change  in  conditions  =  physiological  death).  

“Control”  of  population  numbers  can  result  from  the  density  dependence  of  mortality  
and/or  fecundity.  

A  processes  is  said  to  be  density  dependent  if  its  and  outcome  depend  on  the  number  of  
individuals  in  the  population.  

Density-­‐dependence  reflects  intra-­‐specific  competition  in  which  individuals  of  the  same  
species  compete  amongst  each  other  for  food,  space,  light  or  other  resources.  

The  tendency  for  N  to  increase  when  it  is  low,  and  to  decrease  when  it  is  high  results  in  
logistic  growth.  Possible  patterns:  

-­‐ Both  birth  and  death  rates  are  density-­‐dependent  


-­‐ Only  birth  rate  is  density-­‐dependent  
-­‐ Only  death  rate  is  density-­‐dependent  

Stability  and  fluctuations  of  Populations  


 
 
 

Population  fluctuations  (up  and  down  movements  on  the  graph)  are  due  to  environmental  
changes,  predators,  parasites,  genotype,  mutations,  etc.  

Equilibrium  is  a  state  of  a  system  which  does  not  change.  

Any  equilibrium  may  be  stable  or  unstable.  For  example,  the  equilibrium  of  a  pencil  standing  
on  its  tip  is  unstable;  that  of  a  picture  frame  on  the  wall  is  (usually)  stable.  

An  equilibrium  is  considered  stable  if  the  system  returns  to  it  after  small  disturbances.  If  the  
system  moves  away  from  the  equilibrium  after  small  disturbances,  then  the  equilibrium  is  
unstable.  

Types  of  population  fluctuations  


Three  kinds  of  changes  in  population  size  may  occur  according  to  the  discrete-­‐time  logistic  
model  (Ricker’s  model).  The  discrete-­‐time  logistic  model  

Nt+1  Nt  R  [(K  -­‐  N)/K]  can  be  written  as  Nt+1  =  Ntexp[r0(1-­‐Nt/K)]  

This  version  of  the  discrete-­‐time  logistic  model  (Ricker’s  model)  can  be  used  to  generate  
simulations  of  the  three  types  of  population  dynamics,  which  are  as  follows:  

1. Equilibral  
This  is  where  we  can  get:  

(a)   Monotonous  increase  in  numbers  =sigmoidal,  described  by  the  logistic  equation,  
(b)  Damping  oscillations.  

 
Fig.  2.1:  Monotonous  increase  in  numbers  under  the  discrete-­‐time  logistic  model  

       
Fig.  2.2:  Damping  oscillations  under  the  discrete-­‐time  logistic  model  

In  figs  2.1  and  2.2  the  model  has  stable  equilibrium,  only  the  patterns  of  approaching  
the  equilibrium  are  different.  

In  fig.  2.2  (Damped  oscillations)  N  increases  when  it  is  below  the  carrying  capacity,  K  
and  then  decreases  after  overshooting  K,  and  then  stabilises  when  N=K.  In  this  case  
the  variations  in  population  size  are  marked  at  first  and  then  decrease  with  time.  The  
population  is  going  towards  stability,  e.g.  when  pests  are  disturbed  by  pesticides  and  
then  develop  pesticide  resistance.  In  other  words,  the  magnitude  of  the  oscillations  
decreases  over  time  until  eventually  N  stabilises  at  K.  

Ricker’s  model  is  stable  if  0  <  r  <  2.  The  question  is  what  happens  to  model  
populations  if  stability  is  lost?  Non-­‐equilibrium  dynamics  may  be  of  two  types:  limit  
cycle  –  when  the  trajectory  repeats  itself,  and  chaotic  when  the  trajectory  does  not  
repeat  itself.  

2. Cyclic  (or  periodic  or  limit  cycle)  


Populations  that  cycle  can  do  so  with  various  periods.  A  population  cycling  with  a  
period  2  repeats  itself  every  two-­‐time  steps  (Fig.  2.3).  
 
 
 
 
 
 
 
 
Fig.2.3:  Limit  cycle  with  period=2  under  the  discrete  time  logistic  model  
 
A  population  cycling  with  a  period  4  repeats  itself  every  four-­‐time  steps  (Fig.  2.4).  
 
 
 
 
 
 
 
 
 
 
 
 
Fig.2.4:  Limit  cycle  with  period  =  4  under  the  discrete  time  logistic  model  

 
3. Chaotic  (fluctuations,  but  without  a  regular  period).  Chaotic  populations  never  
exactly  repeat  themselves  (fig.2.5).    

 
 
 
 
 
 
Fig.2.5:  Limit  cycle  with  period  =  4  under  the  discrete  time  logistic  model  

Natural  populations  appear  to  show  all  of  these  different  kinds  of  dynamics.  

   
Lecture  3:  POPULATION  GROWTH  
Module:  Principles  of  Ecology  
Dept.  of  Applied  Biosciences  and  Biotechnology,  Midlands  State  University  

3.1.   Population  Growth  


3.2.   Population  Growth  forms  
3.3.   Isolated  populations,  unlimited  environments  and  the  exponential  model  
3.4.   Isolated  populations,  in  limited  environments  and  the  logistic  model  

___________________________________________________________________________  

   
POPULATION  GROWTH  
Self-­‐reproduction  is  the  main  feature  of  living  organisms.  This  is  what  distinguishes  them  
from  non-­‐living  things.  

Population  growth  is  essentially  a  multiplicative  process  (i.e.  populations  grow  by  
multiplication  not  by  addition)  and  is  continuous  where  there  is  a  complete  overlap  of  
generations.  

Populations  grow  by  birth  and  immigration  and  decline  by  death  and  emigration.  If  birth  +  
immigration  exceed  death  +  emigration,  the  population  grows  and  vice  versa.  

The  rate  at  which  individuals  are  added  to  or  removed  from  the  population  by  birth  and  
death  is  a  function  of  the  size  of  that  population.  

Population  Growth  Forms  


Population  have  characteristic  patterns  of  increase  which  are  called  population  growth  
forms.  These  are  the  Exponential  growth  form  for  isolated  populations  in  unlimited  
environments  and  the  logistic  growth  form  for  isolated  populations  in  limited  
environments.  

The  shapes  of  the  characteristic  curves  of  these  two  population  growth  forms  are  the  J-­‐
shaped  curve  for  exponential  growth  and  S-­‐shaped  or  sigmoid  curve  for  logistic  growth:  

 
Exponential  and  logistic  growth  models  help  to  solve  different  kinds  of  problems  in  ecology,  
e.g.:  

(i) How  long  will  it  take  for  a  population  to  grow  to  a  specific  size?  
(ii) What  will  be  the  population  size  after  n  years  (or  generations)?  
(iii) How  long  can  a  population  survive  at  non-­‐favourable  conditions?  

Isolated  Populations  in  Unlimited  Environments  (Density-­‐


independent  Population  Growth)  
We  will  start  by  considering  an  isolated  population  (i.e.  one  not  interacting  with  other  
populations  through  predation,  parasitism,  competition,  etc.)  in  an  unlimited  environment  
(where  growth  is  unlimited  because  resources  are  not  limiting).  

We  can  consider  two  possible  situations:  

Discrete  Generations  (and  the  Discrete-­‐Time  Exponential  model)  


Here  the  species  has  a  discrete  (short)  breeding  season  and  a  life  span  of  less  than  a  year,  
i.e.  generations  do  not  overlap,  e.g.  annual  plants  and  “univoltine”  insects.  

Changes  in  population  size  in  this  case  can  be  described  by  the  Discrete-­‐Time  Exponential  
Model:  

Nt+1  =  RoNt  
Where:  

t  is  time  measured  in  generations  

Ro  is  the  net  reproduction  rate  

Nt  is  the  population  size  (females)  at  generation  t  

Nt+1  is  the  population  size  (females)  at  generation  t+1  


For  monovoltine  organisms,  R  is  the  average  number  of  offspring  per  one  parent.  For  
example,  in  univoltine  insects  with  a  sex  ratio  of  1:1,  R  =  Fecundity/2.  

If  all  the  offspring  survive  to  breed  the  following  season  the  population  size  after  one  time  
step  will  be  given  by:  

Nt+1  =  RoNt  
Ro  here  can  be  called  the  multiplication  constant.  
Here  we  are  taking  discrete  time  steps  which  can  be  described  by  a  difference  equation:  

Nt+1  /Nt  =  Ro  

The  discrete-­‐time  exponential  model,  Nt+1  =  RoNt    is  a  difference  equation  and  it  gives  the  
number  of  individuals  that  will  be  in  the  population  after  a  single  time  step  (e.g.  after  one  
generation).  

If  we  assume  Ro  is  constant  over  time,  we  can  calculate  the  size  of  future  generations  as  in  
the  table  below:  

Starting  with  an  initial  population  of  10  individuals  (N0  =  10)  and  R0  of  1.5:  

GENERATION   POPULATION  SIZE  (Nt)  


0   10  
1   15=(1.5)(10)  
2   22.5=(1.5)(15)  
3   33.75=(1.5)(22.5)  
 

If  we  plot  this  sort  of  data  we  get  population  growth  curves  with  discrete  time  steps  but  
they  will  be  J  –  shaped.  

 
Note:   Different  R  values  give  different  curves.  Higher  R  means  the  population  grows  faster.  
Different  N0  values  give  curves  of  the  same  shape.  

*By  analogy,  if  you  put  $10  in  the  bank  at  5%  interest  you  w ill  get  a  lot  less  m oney  in  
20  years  than  if  you  put  % 100  in  the  bank,  although  in  both  cases  your  balance  
grows  in  the  same  way.  

The  table  above  also  shows  that  the  size  of  the  increment  in  the  population  in  one  time  step  
depends  on  the  size  of  the  population  at  that  time.  

Populations  therefore  grow  by  multiplication  (proportional  or  geometric  increase)  rather  
than  by  addition  (absolute  or  arithmetic  increase).  

Overlapping  Generations  (and  the  Continuous-­‐Time  Exponential  


model)  
While  the  discrete-­‐time  exponential  model  is  used  to  describe  the  growth  of  isolated  
populations  with  discrete  generations  in  unlimited  environments,  the  Continuous-­‐Time  
Exponential  model  describes  the  growth  of  isolated  populations,  with  overlapping  
generations,  living  in  unlimited  environments.  

Here  we  species  that  have  prolonged  or  continuous  breeding  (i.e.  no  specific  breeding  
season)  and  overlapping  generations  and  there  is  need  for  a  continuous  model  (no  discrete  
time  steps)  which  produces  a  smooth  curve.  

This  kind  of  population  growth  is  best  described  by  the  continuous-­‐time  exponential  
equation:  

dN/dt  =  rN    =  (b-­‐d)N  

Here  dN/dt  is  the  rate  at  which  the  population  is  growing  at  each  instant.    

In  the  continuous-­‐time  exponential  model,  we  have  modelled  an  instantaneous  rate  of  
growth,  not  the  number  that  will  be  there  after  a  single  time  step  (or  next  year).  

The  continuous-­‐time  exponential  model  means  that  changes  in  the  population  size  instantly  
feedback  on  the  population’s  growth  rate.  If  we  add  a  single  individual  right  now,  the  
population’s  growth  rate  will  instantly  change  from  rN  to  r(N+1).  

The  discrete-­‐time  exponential  model  means  that  the  number,  say  of  insects  hatching  in  the  
spring  of  next  year  depends  on  the  number  of  insects  hatching  in  the  spring  of  this  year.  

There  is  no  instantaneous  adjustment  of  the  population  growth  rate.  
In  the  continuous-­‐time  exponential  model,  rather  than  take  discrete  (relatively  large)  time  
steps  as  in  the  discrete-­‐time  exponential  model,  we  take  infinitesimally  small  time  steps  
and  population  growth  can  be  described  by  a  differential  equation:  

dN/dt  =  rN    =  (b-­‐d)N  

Where:    

N  =  population  size  

t  =  time  

r  =intrinsic  rate  of  increase  (or  per  capita  rate  of  population  growth,  etc.)  

b  =  instantaneous  birth  rate  

d  =  instantaneous  death  rate  

Plotting  the  output  of  this  differential  equation  yields  an  exponential  growth  curve:  

 
 

 
The  integral  form  of  this  equation  is    

Nt/No  =  ert  
We  can  rearrange  this  to  give  N  at  any  given  time  as:    

Nt  =  Noert  
There  are  three  possible  model  outcomes:  

(i) Population  exponentially  declines  (r<  0)  


(ii) Population  exponentially  increases  (r>  0)  
(iii) Population  does  not  change  (r  =  0)  

Both  the  continuous-­‐time  and  discrete-­‐time  models  of  density-­‐independent  growth  result  in  
exponential  growth.  When  a  population  is  increasing  exponential,  growth  continues  at  a  
constant  rate  without  limit.  

The  continuous-­‐time  and  discrete-­‐time  exponential  models  sometimes  do  have  similar  
model  outcomes,  but  not  always:  

→   Population  exponentially  declines  if:  

R  <  0   (in  the  discrete-­‐time  model)  

r  <  0     (in  the  continuous-­‐time  model)  

→   Population  exponentially  declines  if  

R  =  1   (in  the  discrete-­‐time  model)  

r  =  0   (in  the  continuous-­‐time  model)  

Parameter  r  is  called:    

-­‐   Malthusian  parameter,  

-­‐   Intrinsic  rate  of  increase,    


-­‐   Instantaneous  rate  of  natural  increase,  or  

-­‐   Population  growth  rate.  

Assumptions  of  the  Continuous-­‐time  Exponential  Model  


(i) Continuous  reproduction  (e.g.  no  seasonality),  
(ii) All  organisms  are  identical  (e.g.  no  age  structure),  and  
(iii) Environment  is  constant  in  space  and  time  (e.g.  resources  are  unlimited).  

However,  the  exponential  model  is  robust,  it  gives  reasonable  precision  even  if  these  
conditions  do  not  hold.    

Organisms  may  differ  in  age,  survival,  and  mortality,  but  the  population  consist  of  a  large  
number  of  organisms  and  their  birth  and  death  rates  tend  to  average.  

Note  that  in  the  exponential  growth  form,  population  density  increases  rapidly  in  
exponential  fashion  and  then  stops  abruptly  as  environmental  resistance  or  another  limiting  
factor  becomes  effective  more  or  less  suddenly.  

 
Applications  of  the  Exponential  Models  
These  models  accurately  describe  the  initial  rate  of  growth  of  some  populations  and  
therefore  find  application  in:  

-­‐ Microbiology  (e.g.  the  growth  of  bacteria  in  petri  dishes)  
-­‐ Conservation  biology  (restoration  of  disturbed  populations)  
-­‐ Fishery  (prediction  of  fish  dynamics)  
-­‐ Plant  and  insect  quarantine  (species  invading  new  habitats  often  grow  exponentially)  

Isolated  Populations  in  limited  Environments  (Density-­‐dependent  


Population  Growth)  
The  Logistic  Model  

We  very  seldom  see  the  kind  of  explosive  population  growth  described  by  the  exponential  
model.  Most  environments  are  limited  in  terms  of  resources,  and  growth  rate  depends  on  
the  population’s  current  size  (N).  

Even  if  an  organisms  is  released  into  a  good  habitat  with  a  large  quantity  of  available  
resources,  as  its  population  size  grows,  the  amount  available  for  each  individual  becomes  
limiting.  The  growth  of  the  population  slows  (by  increasing  mortality  or  decreasing  fecundity  
or  both)  as  density  increases.  In  this  case  population  growth  is  density-­‐dependent  and  is  
best  described  by  the  logistic  model  which  produces  an  S-­‐shaped  growth  curve.  
 
The  population  begins  to  grow  exponentially,  but  as  population  size  increases,  the  rate  of  
growth  declines.  Population  size  then  slowly  approaches  a  maximum  value,  K,  the  carrying  
capacity  of  the  population.  K  is  the  maximum  number  (or  density)  the  habitat  can  sustain,  
and  is  the  upper  asymptote  of  the  sigmoid  curve.  

Each  individual  in  our  population  I  competing  with  other  members  of  the  population  for  
limiting  resources,  i.e.  there  is  intra-­‐specific  competition.    

The  logistic  model  was  developed  by  Belgian  Mathematician  Pierre-­‐Verhulst  (1838)  who  
suggested  that  the  rate  of  population  increase  may  depend  on  population  density.  

Density-­‐dependent  population  growth  can  best  be  described  by  the  continuous-­‐time  
logistic  equation    

 
for  populations  with  overlapping  generations,  and  the  discrete-­‐time  logistic  equation    

Nt+1  =R0Nt[(K  –  N)/K]    


for  populations  with  discrete  generations.  In  both  models,  when  the  population  is  very  
small,  nearly  0  in  size  (K  -­‐  N)/K  is  nearly  1,  so  the  population  undergoes  density-­‐independent  
growth.  Similarly,  when  the  population  is  at  carrying  capacity,  N  =  K  and  so  (K  -­‐  N)/K  =  0  and  
the  population  does  not  grow  at  all.  
Thus,  the  population  has  density-­‐independent  growth  at  low  density,  and  density-­‐
dependent  growth  at  high  density.  

Discrete  Generations  (and  the  Discrete-­‐Time  logistic  model)  


Here  our  species  with  a  discrete  (short)  breeding  season  and  a  life  span  of  less  than  a  year  
(i.e.  generations  do  not  overlap)  is  now  living  in  a  limited  environment  in  which  resources  
will  become  limiting  when  a  certain  population  size  is  reached.  

The  discrete-­‐time  logistic  equation    

Nt+1  =R0Nt[(K  –  N)/K]    


is  identical  to  the  discrete-­‐time  exponential  equation  for  population  growth  in  unlimited  
environments  except  that  R0Nt  is  now  multiplied  by  a  new  term  (K  –  N)/K,  which  can  be  
viewed  as  a  proportion  of  available  “space”  (or  resources)  left  for  any  given  N.    

(K  –  N)  is  the  amount  of  available  space  (total  amount  of  space  minus  amount  already  
occupied),  while  K  is  the  total  amount  of  space.  

Parameter  K  can  also  be  interpreted  as  the  amount  of  resources  expressed  in  the  number  
of  organisms  that  can  be  supported  by  these  resources.  

The  discrete-­‐time  logistic  equation    

Nt+1  =R0Nt[(K  –  N)/K]    


is  also  a  difference  equation  and  it  gives  the  number  of  individuals  that  will  be  in  the  
population  after  a  single  time  step.  In  other  words,  the  discrete-­‐time  logistic  model  treats  
time  as  occurring  in  clear  steps,  lets  say,  of  a  year  and  describes  the  number  of  organisms  at  
the  next  time  step.      

Plotting  the  output  of  this  difference  equation  yields  an  S-­‐shaped  or  sigmoid  curve:  
 
 

Overlapping  Generations  (and  the  Continuous-­‐Time  logistic  model)  


Here  we  have  species  that  have  prolonged  or  continuous  breeding  and  overlapping  
generations  living  in  limited  environments.  

Because  generations  overlap,  population  growth  is  best  described  by  a  differential  equation  
which  produces  a  smooth  curve:  

dN/dt  =  rN[(K  –  N)/K)]      

This  is  a  rate  equation,  and  dN/dt  is  the  rate  at  which  the  population  is  growing  at  each  
instant.  We  have  also  modelled  an  instantaneous  rate  of  growth,  not  the  number  that  will  
be  there  after  a  single  time  step  (or  next  year).  

This  equation  is  identical  to  the  continuous-­‐time  exponential  equation  for  population  
growth  in  unlimited  environments  except  that  rN  is  now  multiplied  by  a  new  term  (K  –  
N)/K,  which  takes  into  account  the  effect  of  intra-­‐specific  competition  between  the  
individuals.    

Just  like  in  the  continuous-­‐time  exponential  model,  changes  in  the  population  size  instantly  
feedback  on  the  population’s  growth  rate.  If  we  add  a  single  individual  right  now,  the  
population’s  growth  rate  will  instantly  change  from  rN  to  r(N+1),  orif  you  added  a  lot  of  
individuals  to  the  population,  so  that  there  were  more  than  K,  the  population  would  reduce  
to  about  K  very  quickly.  

Thus,  in  the  continuous-­‐time  logistic  model  the  population’s  growth  rate  adjusts  itself  
instantaneously,  so  that  there  is  a  gradual  slowing  of  growth  as  N  increases.  Consequently,  
the  population  can  never  overshoot  its  carrying  capacity.  
Under  the  discrete-­‐time  logistic  model,  it  is  possible  for  the  population  to  overshoot  its  
carrying  capacity.  There  is  no  instantaneous  adjustment  of  the  population  growth  rate.  The  
discrete-­‐time  logistic  model  tells  us  something  about  what  happens  when  the  effects  of  
density-­‐dependence  are  not  instantaneous,  but  lag  behind  the  population’s  growth  in  time.  

In  both  models  of  logistic  growth,  population  growth  rate  declines  with  population  size,  N,  
and  reaches  0  when  N  =  K.  if  the  population  size  exceeds  K,  then  the  population  growth  rate  
becomes  negative  and  population  size  declines.  

The  logistic  equations,  therefore  describe  changes  in  rates  of  growth  of  populations.  Growth  
is  slow  at  first,  becomes  faster  and  faster  until  a  maximum  is  reached  and  then  falls  away  
finally  to  zero  as  the  system  reaches  equilibrium  (where  birth  rate  =  death  rate).  

The  logistic  model  has  two  equilibria:  N  =  0  and  N  =  K.  

The  first  equilibrium  is  unstable  because  any  small  deviation  from  this  equilibrium  will  lead  
to  population  growth.  For  example,  immigration  of  a  small  number  of  organisms  results  in  
growth  of  a  population.  The  population  never  returns  to  its  equilibrium.  Instead  population  
numbers  increase  until  they  reach  the  stable  equilibrium  N  =  K.  N  =  K  is  stable  because  after  
a  small  disturbance  the  population  returns  to  its  equilibrium.    

The  dynamics  of  the  discrete-­‐time  logistic  model  are  similar  to  those  of  the  continuous-­‐time  
logistic  model  if  the  population  growth  rate  is  small  (0  <  R  <  0.5).  However,  if  the  population  
growth  rate  is  high,  then  the  model  may  exhibit  more  complex  dynamics  such  as  damping  
oscillations,  cycles  or  chaos  because  of  a  time  delay  in  feedback  mechanisms.  

There  are  no  intermediate  steps  between  time  t  and  time  t+1.  Thus,  over  compensation  
may  occur  if  the  population  grows  or  declines  too  fast  passing  the  equilibrium  point  (K).  

Assumptions  of  the  Continuous-­‐Time  logistic  model  


(i) The  feedback  mechanism  is  instantaneous.  
(ii) There  is  complete  overlap  of  generations  
(iii) Constant  carrying  capacity  
(iv) All  individuals  equal  in  reproductive  potential.  

The  provisors  of  the  logistic  growth  are  met  in  the  growth  of  yeast  and  floating  pond  weed  
in  culture.  

In  both  cases  the  sigmoid  growth  pattern  is  a  close  approximation  to  the  logistic  function.  

The  feedback  mechanisms  are  the  production  of  alcohol,  which  limits  the  growth  of  the  
thin-­‐walled  daughter  yeast  cells  after  budding,  and  mutual  shading  of  the  pond  weed  
fronds,  which  limits  photosynthesis.  
Both  mechanisms  regulate  the  populations  below  levels  where  nutrients  in  the  culture  are  
exhausted.  However,  not  all  populations  show  the  logistic  growth.  In  many  cases,  the  
population  will  briefly  overshoot  its  carrying  capacity,  then  falls  back  down  below  K,  etc,  
resulting  in  oscillations.  

This  overshoot  of  K  and  resultant  oscillations  is  typically  due  to  a  time  lag  between  the  
acquisition  of  resources  and  the  production  of  offspring,  i.e.  the  number  of  offspring  
produced  when  the  population  reaches  K  may  actually  have  been  set  some  time  when  N  
was  still  below  K,  etc.  

Applications  of  the  logistic  Model  


-­‐   Microbiology  (growth  of  yeast),  

-­‐   Biological  control  (introduction  of  the  natural  enemy),  

-­‐   In  managing  game  reserves,  

-­‐   Cattle  ranging  and  

-­‐   In  the  study  of  almost  all  populations  with  a  strong  interaction  among  individuals.  

   
Lecture  4:  ESTIMATION  OF  POPULATION  SIZE  AND  DENSITY  
Module:  Principles  of  Ecology  
Dept.  of  Applied  Biosciences  and  Biotechnology,  Midlands  State  University  

4.1.   Censuring  a  whole  population  


4.2.   Quadrat  Sampling  
4.3.   Simple  Random  or  Systematic  Sampling  
4.4.   Stratified  Sampling  
4.5.   Capture-­‐Mark-­‐Recapture  Method  

___________________________________________________________________________  

   
Estimation  of  Population  Size  and  Density  
Recall  we  define  ecology  as  the  scientific  study  of  interactions  that  determine  the  
abundance  and  distribution  of  organisms  –  here  we  concentrate  on  the  abundance  of  
populations.  

Two  different  measures  of  abundance  can  be  considered:  

(i) Total  number  of  individual  in  the  entire  population  –  seldom  measurable  
(ii) Density  =  the  number  of  individuals  in  a  population  per  unit  area  (e.g.  
number/m2  or  number/ha)  or  volume  (e.g.  number/m3  or  number/litter).  

Censuring  a  whole  population  (total  counts)  


In  cases  where  individuals  are  readily  detected  (often  by  sight),  population  size  can  be  
determined  by  direct  counts  of  all  individuals  in  an  area,  especially  if  their  numbers  are  
small  and  the  area  is  well  bounded  and  not  too  large.  

Example:     trees  in  a  small  isolated  forest.  

Migrating  populations  can  be  censured  using  aerial  photography.  This  method  is  often  used  
when  the  population  has  seasonal  migration.      

When  numbers  are  too  large  or  effort  to  obtain  total  counts  would  be  too  great  (  especially  
when  organisms  are  cryptic),  the  experimenter  can  count  only  a  small  but  representative  
portion  of  the  population  and  use  this  sample  to  estimate  the  population  size.  

Quadrat  Sampling  
-­‐   good  for  sessile  organisms  (e.g.  plants,  sessile  animals  like  adult  barnacles  and  
mussels).  

-­‐   the  site  is  marked  off  in  grid,  and  a  few  quadrats  are  randomly  selected  for  intensive  
study.  

-­‐   the  results  are  used  to  estimate  the  density  of  the  entire  population  (i.e.  mean)  as  
well  as  reliability  of  the  sampling  estimate  (i.e.  variance).  

Simple  random  or  Systematic  Sampling  


-­‐ Two  possible  objectives  for  sampling:  
(i) to  estimate  average  population  density,  and  
(ii) to  make  a  map  of  population  density.  

Traditionally,  random  sampling  was  preferred  to  systematic  sampling  because  random  
sampling  helped  to  avoid  subjective  selection  of  sampling  locations.  However,  systematic  
sampling  has  no  elements  of  subjectivity  if  the  sample  location  is  selected  prior  to  
examining  the  area.  For  example,  there  are  no  subjective  decisions  if  we  sample  every  tenth  
potato  plant  and  count  the  number  Colorado  potato  beetles  on  each  plant.  

Moreover,  systematic  sampling  has  an  advantage  over  random  sampling  if  the  number  of  
samples  is  large  because  of  more  uniform  coverage  of  the  entire  area.  It  is  especially  
important  for  making  population  maps.  

Random  sampling  can  be  used  if  the  objective  is  to  estimate  the  mean  population  density  
and  the  number  of  samples  is  not  large.  

Traditional  statistical  methods  include  estimation  of  the  mean  population  density  (M),  
standard  deviation  (S.D),  and  standard  error  (S.E),  which  is  the  standard  deviation  of  the  
sample  mean.  

S.D  =    S.D     or       S_     =    Sx    


    √N          X               √N  
The  equation  for  standard  deviation  is  derived  assuming  that  all  samples  are  independent.  
This  is  a  very  strong  assumption  which  is  unrealistic  in  many  situations.  Samples  separated  
by  a  small  distance  are  often  positively  correlated.  Before  using  standard  statistics  it  is  
important  to  test  if  the  samples  are  correlated.  This  is  however,  beyond  the  scope  of  this  
module.  

Stratified  sampling  
Stratified  sampling  is  used  if  the  sampled  area  (or  volume)  is  heterogeneous.  In  stratified  
sampling,  the  area  is  subdivided  into  two  or  more  portions  which  are  sampled  separately.  
For  example,  pine  sawflies  prefer  to  spin  their  cocoons  close  to  the  tree,  thus  the  area  
adjacent  to  trees  (within  1m  radius)  can  be  sampled  separately  from  the  rest  of  the  area.  
The  mean  population  density,  M,  is  estimated  as  a  weighted  mean  of  the  mean  densities  Mi,  
in  each  stratum,  i,  with  weights,  wi  equal  to  the  area  covered  by  statum  i:  

M  =  ∑MiWi  
The  standard  error,  S.E,  of  the  mean  is  equal  to:  

SE  =  √∑SEi2.Wi2  
Where:    

SEi  is  the  standard  error  for  the  mean  in  stratum  i.    
Capture-­‐Mark-­‐Recapture  Sampling  
This  method  is  good  for  mobile  organisms  (e.g.  most  animals).  

In  the  first  step  of  this  method,  a  sample  of  individuals  is  captured  from  an  animal  
population,  marked  in  some  way  (e.g.  by  painting,  tagging,  clipping  ears  or  toes,  etc.)  and  
released.  

The  marked  animals  are  given  sufficient  time  to  mix  back  into  the  population.  

A  second  sample  is  then  taken  from  the  population.  By  assuming  that  the  ratio  of  the  
marked  animals  to  the  total  number  of  animals  in  the  recaptured  sample  is  the  same  as  the  
ratio  of  the  individuals  in  the  first  sample  to  the  total  number  of  individuals  in  the  whole  
population,  the  size  of  the  population  can  be  estimated.  

Suppose  the  population  is  of  size  N,  such  that  N  is  the  number  we  wish  to  estimate.  Suppose  
M  organisms  were  captured,  marked  and  released  back  into  the  population.    

After  some  time  which  should  be  sufficient  for  the  organisms  to  mix,  n  organisms  were  
captured,  and  m  of  these  were  marked.  The  proportion  of  the  recaptured  organisms  is  
assumed  to  be  the  same  as  the  proportion  of  the  marked  organisms:  

m/n  =  M/N  
Population  size  can  be  found  as:  

N  =  (nM)/m    (the  Lincoln  Index)  


Where:  

N  =  population  size  you  wish  to  estimate  

M  =  number  of  animals  captured,  marked,  and  released  

n  =  total  number  of  individuals  in  the  recapture  sample  

m  =  number  of  marked  individuals  in  the  recapture  sample  

For  example,  you  capture  and  mark  100  animals,  you  recapture  200  animals  and  find  that  50  
of  them  (i.e.  ¼)  were  marked,  your  estimate  of  the  population  size  will  be:  

N  =     nM     =     200  x  100  
                     m                  50        
 
=   4/1  x  100  

=   400  animals  

The  following  conditions  should  be  met:  

(i) No  immigration,  emigration,  births  or  deaths  between  the  release  and  capture  
times.  
(ii) The  probabilities  of  being  caught  are  equal  for  all  individuals  (including  the  
marked  ones),  i.e.  the  marks  should  not  affect  the  behaviour  of  the  marked  
animals  in  any  way.  
(iii) Marks  (or  tags)  are  not  lost  and  are  always  recognizable.  

   
Lecture  5:  CONCEPTS  OF  HABITAT  AND  NICHE  
Module:  Principles  of  Ecology  
Dept.  of  Applied  Biosciences  and  Biotechnology,  Midlands  State  University  

5.1.   The  Niche  Concept  


5.2.   Niche  Segregation  and  Competition  
5.3.   Overlapping  Niches  and  Competition  
5.4.   Character  displacement  and  Competition  

___________________________________________________________________________  

   
The  Niche  concept  
The  habitat  of  an  organism  is  the  place  where  it  leaves.  The  ecological  niche,  on  the  other  
hand  is  a  more  inclusive  term  that  includes  not  only  the  physical  space  occupied  by  an  
organism  but  also  its  functional  role  in  the  community  (e.g.  its  trophic  position)  and  its  
position  in  environmental  gradients  of  temperature,  pH,  soil  and  other  conditions  of  
existence.  

Organisms  of  any  given  species  can  survive,  grow,  reproduce  and  maintain  a  viable  
population  only  within  certain  temperature  limits.  Their  range  of  temperature  is  the  species’  
ecological  niche  in  one  dimension.  Organisms  of  the  species  in  question  may  also  be  able  to  
survive,  grow,  reproduce  and  maintain  a  viable  population  only  within  certain  limits  of  
relative  humidity.  Taking  temperature  and  relative  humidity  together,  the  niche  becomes  
two-­‐dimensional  and  can  be  visualised  as  an  area  and  so  on.  

The  niche  has  been  variously  defined  as  either  the  predominant  trophic  role  of  an  organism  
within  a  community  (by  Charles  Elton,  1927)  or  the  range  of  physical  environments  in  which  
a  species  can  live  (by  Joseph  Grinnell,  1917).  

The  first  definition  emphasised  the  “profession”  of  the  species  and  the  second  one  
emphasised  its  “address”.  

Evelyn  G.  Hutchinson  (1957)  defined  a  niche  as  a  multi-­‐dimensional  space  or  hyper  volume  
within  which  the  environment  permits  an  individual  or  species  to  survive  indefinitely.  The  
definition  is  more  close  to  Grinnell’s  definition.  It  became  popular  because  the  range  of  
tolerance  to  ecological  factors  can  be  easily  measured,  whereas  species’  “profession”  
cannot  be  measured  easily.  

Hutchinson  also  defined  the  Fundamental  Niche  of  a  species  as  the  extent  of  its  total  
environment  which  it  could  potentially  exploit  in  the  absence  of  other  species  and  its  
Realized  Niche  as  the  resources  which  it  actually  utilizes  when  other  species  are  present.  

Niche  segregation  and  competition  


Within  the  same  species  competition  is  often  reduced  when  different  stages  in  the  
organism’s  life  history  occupy  different  niches,  e.g.  the  tadpole  functions  as  a  herbivore  and  
the  adult  frog  as  a  carnivore  in  the  same  pond.  Niche  segregation  may  even  occur  between  
the  sexes,  e.g.  in  the  woodpecker  of  the  genus  Dendrocarpus  the  male  and  the  female  differ  
in  bill  size  and  foraging  behaviour.  

In  Anopheline  mosquitoes,  females  are  blood  suckers  whilst  males  feed  on  plant  juices.  This  
means  there  is  no  competition  for  resources  between  the  sexes  in  these  species.  
Overlapping  Niches  and  competition  
 

Fig  1:     Distribution  of  Platyhelminth  flat  worms  Planaria  gonocephala  and  P.  montengrina  
along  temperature  gradients  in  streams  where  they  occur  together  (sympatry)  and  
alone  (allopatry)  

The  above  example  illustrates  that,  if  two  competing  species  coexist  in  a  stable  
environment,  then  they  do  so  as  a  result  of  niche  differentiation,  i.e.  differentiation  of  their  
realized  niches.  If,  however,  there  is  no  differentiation,  one  of  the  competing  species  will  
dominate  or  exclude  the  other.  

Character  displacemet:  Sympatry  and  Allopatry  


Differences  in  closely  related  species  are  often  accentuated  in  sympatry  (when  species  occur  
in  the  same  area)  and  weakened  in  allopatry  (when  species  occur  in  different  geographical  
areas)  by  an  evolutionary  process  called  character  displacement.  

Character  displacement  has  two  adaptive  values:  

(i) It  enhances  niche  displacement  thus  reducing  competition,  and  


(ii) It  enhances  genetic  segregation  by  maintaining  species  distinctiveness  (i.e.  
preventing  hybridization)  and  thereby  maintaining  greater  species  diversity  in  
the  community.  

Specialists  and  Generalists  

A  species  which  exploits  a  narrow  niche  is  called  a  specialist  while  one  which  exploits  a  
broad  niche  is  called  a  generalist.  
 

Lecture  6:  COMPETITION  


Module:  Principles  of  Ecology  
Dept.  of  Applied  Biosciences  and  Biotechnology,  Midlands  State  University  

6.1.   Introduction  
6.2.   Intra-­‐specific  competition  
6.2.1.   Scramble  competition  
6.2.2.   Contest  competition  

___________________________________________________________________________  

   
 

COMPETITION  
Introduction  
Competition  is  the  use  or  defence  of  a  resource  by  one  individual  that  reduces  the  
availability  of  that  resource  to  other  individuals,  of  the  same  species  or  of  other  species.  

In  other  words,  competition  is  an  attempt  to  gain  exclusive  or  prior  access  to  a  reserve  
which  is  in  limited  supply  and  it  leads  to  a  reduction  in  the  survivorship,  growth  and/or  
reproduction  of  the  competing  individuals.  

Resource  =     a  substance  or  object  required  by  an  organism  for  normal  maintenance,  
growth  and  reproduction.  What  constitutes  a  resource  is  very  much  species-­‐
specific,  e.g.:  

o Most  photoautotrophs  require  sunlight,  CO2,  H2O  and  a  number  of  inorganic  
nutrients  (including  a  source  of  nitrogen,  phosphorus,  potassium,  etc.);  some  
also  require  various  organic  molecules  (e.g.  certain  vitamins)  as  well.  
 
o Most  heterotrophs,  on  the  other  hand,  require  organic  molecules  as  an  
energy  source  and  various  inorganic  substances  to  maintain  the  composition  
of  their  body  fluids  (e.g.  NaCl),  but  may  also  require  specific  organic  
molecules  they  cannot  synthesize  themselves  (e.g.  various  vitamins  and  
essential  amino  acids).  

“Space”  can  also  be  viewed  as  a  resource,  e.g.  nesting  and  roosting  sites  in  birds  and  
attachment  sites  in  barnacles  and  mussels.  

If  the  supply  of  a  resource  is  scarce  relative  to  its  demand  by  an  organism,  it  is  referred  to  as  
a  limiting  resource  for  that  organism.  

Competition  is  among  the  most  important  factors  in  population  dynamics  of  many  species.  
It  often  determines  the  upper  limit  of  fluctuations  of  population  numbers.  

Intra-­‐specific  competition  
Intra-­‐specific  competition  is  competition  between  individuals  of  the  same  species.  

Main  features  of  intra-­‐specific  competition  


Intra-­‐specific  competition  leads  to  decreased  rates  of  resource  intake  per  individual,  
perhaps  to  decreased  rates  of  individual  growth,  or  to  decreases  in  amounts  of  stored  
reserves.  They  may,  in  turn,  lead  to  decreases  in  survivorship  and/or  decreases  in  fecundity.  
The  first  common  feature  of  intra-­‐specific  competition  is  that  its  ultimate  effect  is  a  
decreased  contribution  of  the  individual  to  the  next  generation.  

The  second  common  feature  of  intra-­‐specific  competition  is  that  the  resources  for  which  
individuals  compete  must  be  in  limited  supply.  

The  third  common  feature  of  intra-­‐specific  competition  is  that  the  competing  individuals  
are,  in  essence,  equivalent  –  but  in  practice  very  much  less  so.  The  fact  that  they  have  been  
classified  as  the  same  species  implies  that  they  have  many  fundamental  features  in  
common,  and  they  may  be  expected  to  use  similar  resources  and  react  in  much  the  same  
way  to  conditions.  However,  it  should  be  noted  that  the  competing  individuals  are  not  
reciprocal.  

There  are  many  occasions  when  intra-­‐specific  competition  is  very  much  one-­‐sided:  a  strong,  
early  seedling  will  probably  shade  a  stunted,  late  one;  and  tall  genotypes  of  maize,  for  
instance,  will  usually  shade  and  suppress  short  genotypes  of  the  same  species.  Thus,  
competing  individuals  of  the  same  species  are  not  entirely  equivalent.    

This  lack  of  exact  equivalence  means  that  the  ultimate  effect  of  competition  is  far  from  
being  the  same  on  different  individuals.  Weak  competitors  may  make  only  a  small  
contribution  to  the  next  generation  or  no  contribution  at  all.  

Strong  competitors  may  have  their  contribution  only  negligibly  affected.  Indeed,  a  strong  
competitor  may  actually  make  a  larger  proportional  contribution  when  there  is  intense  
competition  than  when  there  is  no  competition  at  all.  Thus  competition  can  increase  fitness  
in  a  species.    

The  fourth  and  final  common  feature  of  intra-­‐specific  competition  is  that  the  effect  of  
competition  on  any  individual  is  greater,  the  more  competitors  there  are.  The  effects  of  
intra-­‐specific  competition  are  therefore  said  to  be  density-­‐dependent.  

Anything  that  has  a  relationship  with  the  density  of  the  population  is  said  to  be  density  
dependent.  

For  example,  the  continuous-­‐time  logistic  equation:  

dN/dt  =  rN[(K  -­‐  N)/N]  

shows  that  the  rate  of  population  increase  depends  on  population  density.  At  first  
population  growth  rate  increases  with  population  density  and  then  decreases  with  density  
as  intra-­‐specific  competition  creeps  in.  

we  can  talk  of  the  density  dependence  of  a  number  of  parameters,  e.g.:  

-­‐   the  dependence  of  mortality  on  density  


-­‐   the  dependence  of  birth  rate  on  density  

-­‐   the  dependence  of  growth  rate  on  density  

With  density  dependent  birth/death  rate,  there  is  a  tendency  of  birth/death  rare  to  increase  
with  density.  

With  density  dependent  growth  rate,  there  is  a  tendency  of  growth  rate  to  decrease  with  
density.  

Mortality,  however,  can  be  density  independent.  Mortalities  caused  by  changes  in  weather  
conditions  (e.g.  frost,  heat  waves,  etc.);  volcanic  eruptions,  earthquakes,  etc.,  are  density  
independent  (in  contrast  with,  lets  say,  mortalities  caused  by  competition  or  food)  and  were  
.  

Mortalities  caused  by  predation  are  also  classified  as  density  independent  because  they  are  
not  determined  by  the  density  of  the  prey  population.  Mortality  of  insects  due  to  pesticide  
application  is  also  density-­‐independent.  

Scramble  and  contest  


-­‐ Intra-­‐specific  competition  can  be  conveniently  divided  into  scramble  and  contest  
competition  

Scramble  Competition  
Scramble  competition  occurs  where  there  is  exactly  equal  partitioning  of  resources  with  no  
clear  losers  or  winners.  All  individuals  within  a  species  are  affected  to  the  same  extent.  One  
individual  is  unlikely  to  gain  the  whole  resource  and  under  high  population  levels  this  form  
of  interaction  may  result  in  insufficient  of  the  resource  for  all.  

Interactions  between  individuals  are  usually  indirect,  i.e.,  competing  individuals  do  not  
interact  with  each  other  directly.  Instead,  individuals  respond  to  the  levels  of  a  resource  
which  has  been  depressed  by  the  presence  and  activity  of  other  individuals.  

Thus,  grasshoppers  competing  for  food  are  not  directly  affected  by  other  grasshoppers,  but  
by  the  reduction  in  food  level  and  the  increased  difficult  of  finding  good  food  that  has  been  
left  by  the  others.  

In  such  cases,  competition  may  be  described  as  exploitation  in  that  each  individual  is  
affected  by  the  amount  of  resources  that  remains  after  it  has  been  exploited  by  the  others.  

Main  features  of  scramble  competition  


(i) Interactions  between  individuals  are  usually  indirect.  
(ii) Competition  is  usually  for  food,  nesting  sites  or  mates.  
(iii) Resources  must  be  limiting  
(iv) Scramble  involves  rapid,  efficient  location  and  prior  use  of  resources  
(v) The  resources  are  rapidly  depleted.  
(vi) Competitors  are  approximately  equivalent.  Each  individual  does  less  well  than  if  
it  had  exclusive  access  to  the  resource.  
(vii) There  is  a  threshold  density  above  which  the  population  rapidly  declines  

Fig.1:  Plot  of  mortality  against  population  density  

Fig.  1  shows  a  generalised  relationship  between  population  density  and  percentage  


mortality  resulting  from  scramble  competition.  

Contest  Competition  
Contest  Competition  occurs  when  there  is  unequal  partitioning  of  resources  so  that  there  
are  clear  winners  and  clear  losers.  

Under  the  conditions  of  contest  competition,  the  competitors  interact  in  such  a  way  that  
some  individual  retreat  or  are  eliminated  so  that  the  entire  resource  can  be  utilised  by  the  
victor.  

When  individuals  interact  directly  with  each  other,  one  individual  will  actually  prevent  
another  from  occupying  a  portion  of  the  habitat  and  so  from  exploiting  the  resources  in  it.  
In  such  cases  competition  may  be  described  as  interference.  

This  is  seen,  for  instance,  in  motile  animals  that  defend  territories.  The  result  is  often  that  
the  territory  itself  often  becomes  a  resource.    For  example,  Lion  prides  in  Savuti  national  
Park  in  Botswana  mark  their  territories  with  urine  and  fiercely  guard  them  against  invasion  
by  other  lion  prides.  In  most  cases  fierce  fighting  occurs  and  may  result  in  death  if  the  loser  
cannot  escape.  

In  many  other  vertebrate  populations,  however,  the  contest  may  become  ritualised  so  that  
the  dispute  can  usually  be  settled  without  substantial  damage  to  the  contestants.  A  classical  
example  of  this  form  of  interaction  is  the  study  of  woodland  owl  populations  by  Southern  
(1970).  The  woodland  was  divided  up  into  a  number  of  territories  through  intense  vocal  
conflict  between  the  owls.  

Only  pairs  of  birds  holding  a  territory  were  able  to  breed.  The  young  birds  were  expelled  
from  the  woodland  and  formed  a  reserve  population  of  animals  which  only  breed  following  
the  death  or  successful  challenge  of  the  resident  pair.    

Main  features  of  contest  competition  

(i) Interactions  between  individuals  are  direct.  


(ii) Interactions  are  usually  for  space  incorporating  critical  resources.  
(iii) Resources  potentially  or  occasionally  limiting.  
(iv) Contest  involves  active  exclusion  of  competitors  (i.e.,  interference).  
(v) Competitors  are  not  equivalent,  there  being  clear  winners  and  losers.  
(vi) There  is  a  threshold  density  above  which  a  fixed  number  of  individuals  survive.  

Fig.2:   Plot  of  k-­‐value  against  log  population  density  in  (a)  scramble  
competition,  and  (b)  contest  competition.  
Fig.  2(a)  shows  that  at  the  threshold  density  (T),  mortality  caused  by  scramble  competition  
rises  sharply  from  0%  to  100%.  In  contrast,  the  k-­‐value  for  contest  competition  (Fig  2  (b))  
rises  with  a  slope  =  1,  i.e.,  a  constant  number  of  individuals  survive.  However,  real  slopes  are  
less  than  this  because  of  phenotypic  and  genotypic  variations  between  organisms  (see  plots  
with  broken  lines  in  figs  2(a)  and  2(b)).  

Figs.  2(a)  and  2(b)  show  that  ain  both  scramble  and  contest  competition  there  is  no  
competition  at  all  at  low  densities:  all  individuals  have  as  much  resource  as  they  need,  and  
all  individuals  need  and  get  the  same  amount  (%  mortality  =  0).  Above  a  threshold  density  
of  T  individuals,  in  scramble  competition,  all  the  individuals  still  get  an  equal  share,  but  this  
is  now  less  than  what  they  require,  and  as  consequence  they  all  die  (as  is  found  in  dung  
insects,  and  insect  forest  defoliators).  That  is  the  slope  of  the  ideal  slope  in  fig  2  (a)  suddenly  
changes  from  0  to  infinite  as  the  threshold  T,  is  passed.  

In  contest  competition,  on  the  other  hand,  individuals  fall  into  two  classes  when  the  
threshold  is  exceeded.  T  individuals  still  get  an  equal  and  adequate  share  of  the  resource,  
and  survive  (these  are  the  winners);  all  other  individuals  (the  losers)  get  no  resource  at  all,  
and  therefore  die.  Thus,  in  Fig  2  (b)  the  ideal  slope  changes  at  threshold  from  0  to  1.  

Whilst  there  can  be  no  survivors  in  scramble  competition,  there  are  always  just  T  survivors  
in  contest  competition  irrespective  of  the  initial  density,  because  mortality  compensates  
exactly  for  the  excess  number  of  individuals.  

Scramble  and  contest  can  also  be  viewed  in  terms  of  fecundity.  Below  the  threshold  there  is  
no  competition,  all  individuals  produce  the  maximum  number  of  offspring.  Above  the  
threshold,  scramble  leads  to  production  of  offspring  whatsoever,  while  contest  leads  to  T  
individuals  producing  the  maximum  number  of  offspring  and  the  rest  producing  none  at  all.  

In  conclusion  we  need  to  note  that  intra-­‐specific  competition  results  in:  

(i) Evolution  of  efficiency  in  use  of  resources.  


(ii) Broaden  niche  breath.  

 
(iii) Regulation  of  size  through  contest  competition  and  the  exclusion  of  some  
individuals  from  access  to  critical  resources.  

   
Lecture  7:  INTERSPECIFIC  COMPETITION  
Module:  Principles  of  Ecology  
Dept.  of  Applied  Biosciences  and  Biotechnology,  Midlands  State  University  

7.1.   Inter-­‐specific  competition  


7.2.   Exploitative  competition  
7.3.   Interference  competition  
7.4.   Lotka  –  Volterra  Model  
7.5.   Competitive  Exclusion  Principle  
7.6.   Coexistence  

___________________________________________________________________________  

   
INTER-­‐SPECIFIC  COMPETITION  
Inter-­‐specific  competition  is  competition  between  individuals  of  different  species,  e.g.,  
between  lions  and  hyenas,  cattle  and  donkeys,  etc.  

The  essence  of  inter-­‐specific  competition  is  that  individuals  of  one  species  suffer  a  reduction  
in  fecundity,  survivorship  or  growth  as  a  result  of  resource  exploitation  or  interference  by  
individuals  of  another  species.  

Inter-­‐specific  completion  occurs  between  similar  species  with  similar  ecological  


requirements  and  usually  results,  if  the  populations  are  resource  limited  and  emigration  is  
prevented,  in  the  decline  and  ultimate  extinction  of  one  of  the  species.  

If  competing  species  are  ecologically  identical  (i.e.  use  the  same  resources),  then  inter-­‐
specific  competition  is  equivalent  to  intra-­‐specific  competition.  Each  individual  competes  
with  all  organisms  of  both  populations.  As  a  result,  the  population  growth  rate  of  each  
species  is  determined  by  the  sum  of  numbers  of  both  populations:  

dN1   =   r1N1(K1  –  (N1+N2))      


dt              K1      

   

dN1   =   r1N1(K1  –  N1-­‐N2))     equation  1  

dt                K1  
 

for  species  1,  and  

dN2   =   r2N2(K2  –  (N2+N1))      


dt              K2      

   

dN2   =   r2N2(K2  –  N2-­‐N1))     equation  2  

dt                K2  
for  species  2.  
The  species  with  a  higher  carrying  capacity  (K)  always  wins.  Higher  carrying  capacity  means  
the  species  can  endure  more  crowding  than  the  other  species  (e.g.,  due  to  more  effective  
search  for  resources).  

If  the  competing  species  are  sufficiently  different  then  intra-­‐specific  competition  is  stronger  
than  inter-­‐specific  competition.  Individuals  of  another  species  are  not  considered  as  “full”  
competitors.  As  a  result,  the  number  of  inter-­‐specific  competitors  is  multiplied  by  a  weight  
αi<1:  

dN1   =   r1N1(K1  –  N1-­‐α12N2))       equation  3  

dt                K1  
 

dN2   =   r2N2(K2  –  N2-­‐α21N1))       equation  4  

dt                K2  
 

Suppose  that,  together,  10  individuals  of  species  2  have  the  same  competitive  inhibitory  
effect  on  species  1  as  does  a  single  species  1  individual.  The  total  competitive  inhibitory  
effect  on  species  1  (both  intra-­‐  and  inter-­‐specific)  will  then  be  equivalent  to  (N1  +  N2/10))  
species  1  individuals.  

We  call  the  constant  –  1/10  in  this  case,  the  coefficient  of  competition  and  denote  it  by  α12  
since  it  measures  the  competitive  inhibitory  effect  on  species  1  of  species  2.  In  other  words,  
multiplying  N2  by  α12  converts  it  to  a  number  of  N1  –  equivalents.    Note  that  α12<1  means  
that  species  2  has  less  inhibitory  effect  on  species  1  than  species  1  has  on  itself,  α12>1  
means  that  species  2  has  a  greater  inhibitory  effect  on  species  1  than  species  1  has  on  itself.  

 α21  measures  the  competitive  inhibitory  effect  on  species  2  of  species  1.  

Species  coexistence  is  possible  if  intra-­‐specific  competition  is  stronger  than  inter-­‐specific  
competition.  This  occurs  if  competing  species  have  different  preferences  in  resource  
utilisation.  

Inter-­‐specific  competition  is  conveniently  divided  into  exploitative  and  interference  


competition.  

 
Exploitative  Competition  

 
In  Exploitative  Competition  individuals  interact  with  each  other  indirectly,  responding  to  a  
resource  level  which  has  been  depressed  by  the  activities  of  competitors.  

When  inter-­‐specific  competition  involves  resource  exploitation,  both  species  consume  the  
resource  reducing  its  availability.  This  depletion  may  move  either  or  both  species  from  their  
“saturated  range”  down  into  their  “limiting  range”  or  from  the  limiting  range  to  below  the  
“threshold”  (where  the  individual  dies  or  the  population  goes  extinct).  

Examples  of  exploitative  competition  include  the  following:  

-­‐   American  Bison  and  domestic  cattle  compete  for  grass  

-­‐   Goats  and  sheep  compete  for  grass  

-­‐   Tsessebe  and  Wildebeest  compete  for  grass  

-­‐   Various  fresh  water  algal  species  compete  for  inorganic  nitrogen  

Interference  Competition  
In  Interference  Competition  interaction  between  individuals  is  direct.  An  individual  of  one  
species  may  prevent  individuals  of  other  species  from  occupying  a  good  habitat,  from  using  
resources,  or  may  actually  kill  individuals  of  the  other  species,  i.e.  the  behaviour  of  one  
species  limits  others’  access  to  or  use  of  resources.  

Thus  interference  competition:  


-­‐   may  be  aggressive  (antagonistic)  behaviour,  e.g.,  scavengers  at  a  Zebra  kill  on  the  
Serengeti  plains  in  Tanzania  or  fierce  fighting  between  lions  and  hyenas  at  a  Zebra  
kill  in  Savuti  National  Park,  

-­‐   may  be  “unintended”  –  e.g.  turbulence  generated  by  filter-­‐feeding  water  fleas  
(Cladocera:  like  Daphnia)  can  disrupt  filter-­‐feeding  by  the  much  smaller  rotifers  (note  
Cladocerans  may  also  kill  the  smallest  rotifers),  or  

-­‐   may  be  due  to  chemicals  (e.g.  toxins)  –  like  “allelopathic”  chemicals  produced  by  
several  plant  species  that  inhibit  the  germination  of  other  (potentially  competing)  
species.    

Field  Experiments  
We  can  ask  various  questions  about  competition  in  the  field:  

-­‐   If  two  species  coexist  naturally  do  they  compete?  

*   To  answer  such  a  question  you  can  perform  a  removal  experiment  –  you  can  remove  
one  species  and  observe  the  effect  on  the  other  species  –  if  the  population  of  the  
other  species  increases,  it  means  they  may  have  been  competitors.  

-­‐   If  two  species  do  not  coexist,  why  not,  is  it  competition?  

*   You  can  perform  an  addition  experiment  and  determine  whether  the  addition  of  one  
species  affects  the  density  of  the  other  –  does  competition  determine/limit  
distribution.  

Examples  if  interference  competition  from  field  experiments  


1.   Bedstraws  –  Tansley  (1917)’s  experiment  with  two  species  of  bedstraws  (Galium  
spp)  in  England  provides  a  clear  example  of  both  interference  competition  and  the  
competitive  exclusion  principle.  

-­‐   Bedstraws  of  the  genus  Galium  are  small  plants  which  can  grow  on  soil  of  any  pH.  
Galium  hercynicum  (Gh),  however,  occurs  on  acidic  soils,  while  Gallium  pumilum  
(Gp)  occurs  on  calcareous,  basic  soils  in  England,  but  when  alone  either  species  can  
grow  on  soils  of  any  pH.  

-­‐   The  fundamental  niche  for  both  Gh  and  Gp  are  soils  of  all  pH’s.  

-­‐   When  grown  together,  Gh  wins  on  acidic  soils  and  Gp  wins  on  basic  soils.  

-­‐   The  realized  niche  for  Gh  is  made  up  of  acidic  soils  only,  while  the  basic  soils  
constitute  the  realised  niche  for  Gp.  
-­‐   Gh  drives  Gp  to  extinction  on  acidic  soils  while  Gp  drives  Gh  to  extinction  on  basic  
soils.  This  shows  that,  which  species  is  the  winner  depends  on  the  habitat  and  that  
unless  competing  species  partition  the  soil  pH  dimension,  one  will  be  driven  into  
extinction.  

-­‐   The  mechanism  of  exclusion  in  this  example  involves  overgrowth  and  shading  (i.e.  
shoot  competition).  

2.   Barnacles  –  Connell  (1961)’s  work  with  barnacles  in  Scotland  also  provides  another  
clear  example  of  both  interference  competition  and  the  competitive  exclusion  
principle.  

-­‐   Barnacles  are  marine  crustaceans  found  in  the  rocky  intertidal  zone.  The  adult  
barnacles  are  sessile  animals  attached  to  rocks.  When  covered  with  water  they  filter-­‐
feed  on  suspended  particles  such  as  zooplankton  and  detritus.  

-­‐   The  larvae,  which  are  planktonic,  disperse,  then  settle  and  metamorphose.  

 
 

 
-­‐   The  adults  of  two  species  of  barnacles,  Balanus  balanoides  and  Chthamalus  stellatus  
co-­‐occur  in  the  same  areas  in  the  field,  but  do  not  overlap  in  tidal  level  occupied.  

-­‐   Balanus  (B)  is  restricted  to  the  lower  levels  (near  the  low  tide  mark),  Chthamalus  
(Ch)  is  restricted  to  higher  levels  (near  the  high  tide  mark).  Desiccation  is  frequent  at  
the  high  tide  level  but  infrequent  at  the  low  tide  level.  

-­‐   The  larvae  of  both  species  settle  in  all  zones  but  the  subsequent  disappearance  of  Ch  
from  the  B  zone  suggests  either  that  B  excludes  Ch  from  the  zone  of  their  potential  
overlap  or  that  Ch  is  simply  unable  to  live  there.    

-­‐   Connell  carried  out  a  removal  experiment  in  order  to  distinguish  between  the  two  
alternatives.  He  removed  B  from  some  plots  and  compared  the  results  with  those  
from  the  controls  (unaltered).  

Results:     Ch  survives  at  all  levels  where  B  is  removed  –  so  Ch’s  fundamental  
niche  is  the  entire  intertidal  zone,  but  its  realized  niche  is  only  near  
the  high  tide  line.  

-­‐   In  regions  where  B  eliminates  Ch,  it  does  so  by  interference  
competition:  B  smoothers,  undercuts  or  crushes  Ch.  

-­‐   Connell  concluded  that  Ch’s  distribution  was  limited  by  competition.  

In  the  opposite  experiment,  Connell  removed  Ch  from  some  plots  and  noted  the  
effect  on  B.  

Results:     removal  of  Ch  had  no  effect  on  b  whose  distribution  was  still  limited  
to  the  lower  reaches.  

-­‐   Connell  concluded  that  the  distribution  of  B  is  restricted  by  a  physical  
factor  (probably  inability  to  resist  desiccation),  not  by  competition.  

-­‐   Although  the  competition  between  B  and  Ch  is  markedly  one-­‐sided,  
these  two  barnacles  partition  the  niche  on  the  tidal  level  dimension.  

Laboratory  Experiments  
A  laboratory  is  a  controlled  environment  –  you  can  force  two  species  to  compete  for  one  
resource  (e.g.  you  can  put  them  in  the  same  test  tube)  –  the  work  of  Gause.  

Example  of  interference  competition  from  laboratory  experiments  


G.  F.  Gause  was  a  Russian  ecologist  who  studied  competition  in  laboratory  experiments  
using  three  species  of  Paramecium  in  193  and1935.  
Paramecium  is  a  ciliated  protozoan  which  can  be  maintained  in  test  tubes  feeding  on  
bacteria  (and  yeast)  growing  on  oatmeal.  

(i) P.  aurelia  (Pa)     vs   P.  caudatum  (Pc):  

Result:     Pa  wins;  Pc  goes  into  extinction  –  complete  exclusion  

(ii) P.  bursaria  (Pb)   vs   P.  caudatum  (Pc):  

Result:     with  these  two  species  we  get  stable  coexistence  (no  competitive  
exclusion),  

But  careful  examination  of  the  tubes  indicates  that  the  two  species  do  not  
overlap  in  space:    

Pc  tended  to  live  and  feed  on  the  bacteria  suspended  in  the  medium,  whilst  
Pb  was  concentrated  on  the  yeast  cells  at  the  bottom  of  the  tubes.    

Pb  has  symbiotic  green  algae  and  algal  photosynthesis  provides  02  to  the  
paramecium  even  in  the  hypoxic  zone  of  the  test  tubes.  

So  here  the  species  partition  the  niche  even  though  they  are  confined  to  a  
single  test  tube.    

The  Lotka  -­‐  Volterra  Model  


While  the  logistic  model  is  a  population  growth  model  with  takes  into  account  intra-­‐specific  
competition,  the  Lotka  –  Volterra  model  is  a  competition  model  including  the  effect  of  both  
intra-­‐specific  and  inter-­‐specific  competition  on  population  dynamics.  

Remember:  

-­‐   differential  equations:    

dN/dt  =  slope  of  live  individuals  at  a  particular  instant  in  time  

-­‐ exponential  equation:  

dN/dt  =rN  –  no  intraspecific  competition  (unlimited  environment)  

-­‐ logistic  equation:  


dN/dt  =  rN[(K  -­‐  N)/K]  –  includes  intraspecific  competition.  

As  N  approaches  K,  dN/dt  declines  

If  N>K,  dN/dt  becomes  negative.  


Now  the  Lotka  –  Volterra  Model  is  like  the  logistic  model,  but  it  now  includes  the  density  of  
the  other  species.    

Considering  the  population  dynamics  of  species  1  in  the  presence  of  species  2,  the  logistic  
equation  becomes:  

dN/dt1  =  r1N1[(K1  –  N1  –  α12N2)/K1]  

so  in  the  final  term  (previously  (K1  –  N1)/N1  we  now  add  N2  multiplied  by  α12  to  N1.  

Where:   α12  is  the  competition  coefficient  and  measures  the  competitive  effect  
of  species  2  on  species  1  

When  α12  has  a  large  value  (i.e.  approaches  1.0)  it  means  species  2  is  a  strong  competitor  
and  exerts  a  large  inhibitor  effect  on  species  1.  

α12  defines  the  inhibitory  effect  of  species  2  on  species  1  in  terms  of  equivalent  numbers  of  
species  1  individuals.  

For  example:  

If  species  2  is  very  similar  to  species  1,  α12  approaches  1.0  such  that  an  individual  of  species  
2  has  nearly  as  much  effect  on  dN/dt1  as  an  individual  of  species  1  would,  i.e.  there  is  near  
complete  niche  overlap:  

(K1  –  N1  -­‐  α12  N2)/K1  approaches  (K1  –  N1  –N2)/K1  

If  species  2  is  quite  dissimilar  to  species  1,  there  is  only  slight  niche  overlap  -­‐  α12  approaches  
zero  (N2  has  little  effect  on  dN/dt1):    

(K1  –  N1  -­‐  α12  N2)/K1  approaches  (K1  –  N1)/K1  

Similarly,  the  population  dynamics  of  species  2  can  be  described  by  the  equation:  

dN/dt2  =  r2N2[(K2  –  N2  –  α21N1)/K2]  

Note:     some  authors  use  α  (instead  of  α12)  for  the  effects  of  species  2  on  species  1  and  β  
(instead  of  α21)  for  the  effects  of  species  1  on  species  2    or  even  W1  and  W2  for  α12  
and  α21,  respectively.  
 
Possible  outcomes  of  the  Lotka  -­‐  Volterra  Model  
We  can  employ  a  simple  graphical  model  to  predict  the  outcome  of  competition  between  
two  species  if  we  know  the  initial  densities  of  each  (i.e.  N1  and  N2)  and  the  respective  
competition  coefficients  (i.e.  α12  and  α21).  

We  can  start  by  graphing  “zero  net  growth  isocline”  (ZNGI)  for  species  1  on  a  plot  of  N1  vs  
N2.    This  line  shows  all  the  combinations  of  N1  and  N2    that  will  give  no  net  growth  (or  
decline)  in  the  population  size/density  of  species  1  over  time,  i.e.  where:    

dN/dt  =  0  

In  order  to  draw  the  N1  isocline  we  will  use  the  fact  that  on  it  dN/dt  =  0  

 i.e.      r1N1[(K1  –  N1  –  α12N2)/K1]  =  0  

This  is  true,  less  importantly,  when  r1  or  N1  are  zero,  and  more  importantly  when:    

K1  –  N1  –  α12N2  =  0       or     N1  =  K1–  α12N2  

Since  our  isocline  is  a  straight  line  we  can  draw  it  by  finding  two  points  on  it  and  joining  
them.  

Along  the  x-­‐axis  (i.e.  along  N1,  where  N2  =  0),  dN/dt  =  0  when  N1  =K1,  as  there  are  no  
individuals  of  species  2.  (N2  =  0;  N1  =K1)  is  point  A  in  fig  1.  

Along  the  y-­‐axis  (i.e.  along  N2,  where  N1  =  0),  dN/dt  =  0  when  N2  =K1/α12  where  there  are  
enough  individuals  of  species  2  to  completely  fill  up  all  the  spaces  for  species  1,  e.g.  if  K1  is  
100  and  each  species  2  individual  fills  up  half  of  a  species  1  individual  space,    α12=0.5  and  so  
you  need    

K1/  α12  =  100/0.5  =  200  N2  individuals  to  reach  species  1’s  ZNGI.  
(N1  =  0;  N2  =K1/α12  )  is  point  B  

 
 Fig.  1:  N1  isocline  

Anywhere  “below”  (to  the  left  of)  the  ZNGI  there  is  “available  space”  for  species  1  (i.e.  it  is  
below  its  carrying  capacity),  so  N1  will  increase  towards  the  line.  

Anywhere  “above”  (to  the  right  of)  the  ZNGI  there  are  more  individuals  of  species  1  than  
there  is  available  space  for  them  (i.e.  it  is  above  its  carrying  capacity),  so  N1  decreases  
towards  the  line.  

Fig.2:  (a)  N2  –  isocline     (b)  N1-­‐  and  N2-­‐  isoclines  

Similarly,  we  can  graph  the  ZNGI  for  species  2  on  the  same  axis  (Fig  2(b))  Fig  2(b)  shows  that  
species  1  drives  species  2  into  extinction.  When  both  isoclines  are  parallel,  the  species  with  
a  higher  carrying  capacity  will  win.  

 
 

Fig.3:  Isoclines  for  N1-­‐  and  N2-­‐  showing  outcomes  of  competition  

Fig.  3  (a)  shows  that:  

K2  >  K1/  α12   and     K1  >  K2/  α21  

i.e.     K1  <  K2  α12   and     K2  <  K1α21  

This  means  that  interspecific  effects  are  more  important  than  intraspecific  effects:  both  
species  are  strong  interspecific  competitors.  There  are  two  stable  points  (N1  =  K1;  N2  =  0  and  
N2  =  K2;  N1  =  0)  and  an  unstable  equilibrium  combination  of  N1  and  N2.  This  means  that  one  
species  always  drives  the  other  into  extinction,  but  what  species  will  be  excluded  depends  
on  initial  densities.  In  other  words,  if  α  >  1  and  isoclines  intersect,  one  species  will  exclude  
the  other,  but  which  species  excludes  the  other  depends  on  initial  densities  of  both  
populations.    

Fig.  3  (b)  shows  that:  

   K1/  α12  >  K2   and     K2/  α21  >  K1  

i.e.     K1  >  K2  α12   and     K2  >  K1α21  

This  means  that  intraspecific  effects  are  now  more  important  than  interspecific  effects:  both  
species  are  weak  interspecific  competitors,  and  the  species  can  coexist  at  a  particular  
equilibrium  combination  of  N1  and  N2.    

Species  coexistence  is  possible  if  intraspecific  competition  is  stronger  interspecific  
competition.  This  occurs  if  competing  species  have  different  preferences  in  resource  usage.  

Using  the  Lotka  –  Volterra  model,  we  can  therefore  envision  4  possible  relationships  
between  the  two  species  each  with  a  different  outcome:  
1   Species  1  strong  competitor,  species  2  weak  competitor:  

-­‐   Species  1  wins,  species  2  is  excluded….no  matter  what  the  initial  densities  are.  

2   Species  2  strong  competitor,  species  1  weak  competitor:  

-­‐   Species  2  wins,  species  1  is  excluded….no  matter  what  the  initial  densities  are.  

3   Both  species  strong  competitors:  

-­‐   Which  species  wins  depends  on  initial  densities.  Coexistence  is  possible,  but  the  
equilibrium  will  be  unstable  –  diverges  at  slightest  disturbance.  

4   Both  species  weak  competitors:  

-­‐   Result  is  stable  coexistence,  stable  equilibrium.  

-­‐   Get  same  equilibrium  densities,  irrespective  of  initial  densities.  

Problems  with  the  Lotka  -­‐  Volterra  Model  


-­‐   Simplest  possible  model;  biological  unrealistic.  

-­‐   Linear  effect  of  interspecific  competition.  

-­‐   How  to  determine  α  values  in  the  field  or  lab  

-­‐   Does  not  explicitly  define  resources  or  nature  of  interaction  (e.g.  exploitation  vs  
interference)  –  underlying  explanation  may  be  very  complex.  

Theory:  The  Competitive  Exclusion  Principle  


-­‐   Complete  competitors  cannot  coexist  at  equilibrium  in  a  stable  environment.  

-­‐   If  two  competing  species  coexist  in  a  stable  environment,  and  there  are  no  
differences  in  their  realized  niches,  then  one  will  be  competitively  excluded.  

-­‐   If  they  do  coexist  in  the  field,  there  must  be  some  niche  differentiation.  

-­‐   Can  we  disprove  the  theory?  

-­‐   If  we  find  coexistence,  we  look  for  niche  differentiation  –  if  we  do  not  find  any,  it  
means  we  did  not  look  hard  enough.  

Examples  of  competitive  exclusion  from  laboratory  experiments  


1.   Gause’s  work  with  Paramecium  

P.  aurelia  vs  P.  caudatum  

P.  aurelia  excludes  P.  caudatum  from  the  test  tube.  

2.          Park  (1954)’s  work  with  flour  beetles  Tribolium  confusum  and  T.  castenium  

In  a  series  of  simple,  sterilised  cultures,  Park  held  most  environmental  variables  
constant  but  varied  the  climate.  In  all  conditions,  both  species  were  able  to  survive  in  
monospecific  cultures:  the  fundamental  niches  of  both  species  spanned  the  whole  
climatic  range.  In  mixed  cultures,  however,  T.  castenium  completely  excluded  T.  
confusum  under  hot  –  moist  conditions  and  T.  confusum  excluded  T.  castenium  
completely  under  cold-­‐dry  conditions.  

Examples  of  competitive  exclusion  in  the  field  


1.   Connell’s  barnacles  in  Scotland  

-­‐   Balanus  excluded  Chthamalus  from  the  zone  of  their  potential  overlap.  

2.   Bedstraws  

-­‐   Galium  hercynicum  excluded  Galium  pumilum  on  acidic  soils  while  G.  pumilum  
excluded  G.  hercynicum  on  basic  soils.  

-­‐   Competitive  exclusion,  in  this  case,  was  very  localised.  

Coexistence  
When  competitive  exclusion  does  not  occur  (i.e.  when  coexistence  occurs),  we  must  have:  

1.   Niche  differentiation  (i.e.  competition  not  complete,  niche  overlap  not  complete).  

2.   Unstable  environment  (not  at  equilibrium).  

3.   Spatial  patchiness  (not  at  equilibrium)  

We  can  now  consider  these  in  greater  detail:  

(a)  Niche  differentiation  


Species  can  coexist  when:  

-­‐   they  are  not  complete  competitors,  and  

-­‐   when  differences  exist  in  their  realized  niches.  


For  example,  Paramecium  in  the  Lab  exhibited  niche  differentiation  with  one  species  
restricted  at  the  bottom  of  the  test  tube  and  the  other  at  the  top.  

Field  example:  coexistence  of  three  species  of  birds  (tits)  in  English  broadleaved  woodlands  
–  work  of  Lack  (1971):  

Lack  concluded  that  the  three  species  of  birds  are  separated  from  each  other  at  most  times  
of  the  year  by  their  feeding  station  and  the  size  of  their  insect  prey:  

-­‐   Blue  tit  (Parus  caeruleus):  feeds  in  oak  trees  on  insects  <2mm  in  length.  

-­‐   Marsh  tit  (P.  palustris):  feeds  in  shrubs  on  insects  3  -­‐4mm  in  length.  

-­‐   Great  tit  (P.  major):  feeds  on  the  ground  on  insects  >6mm  in  length.  

So,  we  have  two-­‐dimensional  niche  differentiation.  

The  niche  differences  are  also  reflected  in  differences  in  body  size,  and  size  and  shapes  of  
the  bills.  

Other  examples  of  niche  differential  include  mixed  herds  of  grazers  that  graze  together  in  
the  savannah  grasslands.  For  instance  mixed  herds  of  zebras  and  wildebeest  graze  together  
on  the  vast  plains  of  the  Serengeti  but  do  not  compete.  Wildebeests  eat  tender  grass  
shoots,  while  zebras  crop  tall,  tough  grass  that  the  wildebeest  cannot  bite  off.  

Why  do  niche  differences  exist?  


Niche  differences  exist  either  because  of  current  competition,  competition  in  the  past  or  
some  other  reasons  with  nothing  to  do  with  competition.  

(i)   Current  competition  

The  species  can  restrict  each  other  to  smaller  realized  niches  through  
completion.  

Lack  did  not  do  field  experiments  to  show  competition.  He  could  have  removed  
one  species  to  show  increased  growth,  survivorship  and/or  fecundity  of  the  other  
species  –  or  show  expansion  of  niches  as  these  differences  may  reflect  choice  or  
behavioural  limits.  

(ii)   Competition  in  the  past  

The  species  might  have  evolved  niche  differences  to  avoid  competition.  

It  could  be  that  natural  selection  favoured  those  individuals  with  less  overlap,  
and  so  with  less  inter-­‐specific  competition.  Individuals  of  both  species  in  the  area  
of  overlap  could  have  been  negatively  impacted  resulting  in  lower  fitness.  If  
niche  dimension  is  a  heritable  trait,  those  in  overlap  will  decrease  over  time.  

Character  displacement  –  species  can  evolve  morphological  adaptations  which  


allow  them  (largely  if  not  totally)  to  avoid  competition.  

Niche  differentiation  due  to  natural  selection  could  have  arisen  in  order  to  avoid  
competition.  In  Connell  (1980)’s  words  we  can  explain  niche  differences  in  the  
present  by  invoking  the  “ghost  of  competition  past”.  

(iii)   Evolved  niche  differences  for  some  other  reasons  

During  the  course  of  their  evolution,  species  might  have  adapted  to  their  in  
environment  in  different  ways,  but  in  ways  with  nothing  to  do  with  interspecific  
competition.  

Its  quite  possible  that  the  species  did  not  compete  at  any  time  in  the  past,  and  do  
not  compete  at  present.  

We  cannot  distinguish  between  these  hypotheses  based  on  field  observations  alone.  
We  need  to  carry  out  experiments  to  determine  whether  the  species  are  currently  
competing.  

It  is  impossible  to  demonstrate  competition  in  the  past  because  observed  niche  
differences  may  be  consistent  with,  but  not  sufficient  to  demonstrate  current  or  past  
competition.  

(b)  Unstable  Environment  


E.G.  Hutchinson’s  “Paradox  of  the  Plankton”  

When  the  principle  of  competitive  exclusion  became  widely  known  among  ecologists,  it  
seemed  to  contradict  with  some  well-­‐known  facts  and  this  contradiction  was  formulated  as  
“paradoxes”  for  example  the  “paradox  of  the  plankton”  focused  on  the  variability  of  
plankton  organisms  which  all  seemed  to  use  the  same  resources.  

All  phytoplankton  species  use  the  same  essential,  limiting  resources:  light,  P,  N,  and  trace  
elements.  There  are  not  so  many  mineral  components  dissolved  in  water  as  compared  to  
the  large  variability  in  the  phytoplankton  species.  This  means  that  there  is  a  broad  overlap  in  
niches.  

In  some  systems  spatial  heterogeneity  (or  patchiness)  allows  species  to  coexist,  but  no  such  
spatial  patchiness  (no  physical  niche  differentiation)  exists  in  the  case  of  the  lake  as  all  the  
mineral  components  occur  throughout  the  mixed  zone  of  lakes.  

Apparent  paradox  –  many  species  coexist,  but  how?,  with  no  competitive  exclusion.    
The  answer  is  that  the  environment  changes  before  exclusion  can  take  place  (i.e.  the  
environment  is  unstable).  As  the  environment  changes,  different  species  become  favoured  
and  a  state  of  equilibrium  will  never  be  reached  –  rate  of  environmental  change  is  faster  
than  the  rate  of  exclusion.  

(c)  Spatial  patchiness  


1.   Ephemeral  habitats  –  environment  only  exists  temporarily  –  e.g.  vernal  ponds  
(which  dry  up),  rotting  corpses,  fruits  and  dung.  

The  inferior  competitor  may  be  able  to  coexist  with  a  superior  competitor  by  being  a  
better  colonizer  (which  finds  new  habitats  faster)  and/or  being  able  to  complete  its  
life  cycle  before  the  habitat  vanishes  (early  age  at  first  reproduction;  matures  at  
smaller  size,  etc.)  

-­‐   the  system  not  at  equilibrium  

2.   Unpredictable  gaps  –  e.g.  light  gaps  in  tropical  rainforests.  Light  intensity  is  very  low  
under  the  canopy  (due  to  interception  by  the  canopy)  –  seedlings  cannot  establish  
under  a  canopy,  but  only  grow  to  adults  in  “light  gaps”.  

Light  gaps  are  caused  by  tree  falls,  e.g.  with  high  winds,  the  tallest  canopy  trees  can  
be  blown  down.  

One  possible  strategy  then  is  to  be  a  “fugitive”  species  which  colonises  light  gaps  (or  
newly  formed  ephemeral  ponds,  etc.)  first  and  the  produce  mobile  or  long-­‐dispersed  
seed.  E.g.  fast  growing  vines  can  coexist  with  superior  competitors  if  enough  light  
gaps  are  formed.  

In  light  gaps,  species  are  not  at  equilibrium  –  eventually,  the  gap  closes  in  and  
dominant  canopy  trees  win.  An  inferior  competitor  can  “pre-­‐empt”  space  (by  getting  
an  earlier  start,  the  inferior  competitor  can  persist  longer  than  it  would  if  it  started  
to  grow  at  the  same  time  as  the  superior  competitor).  

General  conclusions  about  competition  


-­‐   the  interference/exploitation  dichotomy  in  interspecific  competition  is  somewhat  
similar  to  that  between  scramble  and  contest  (in  intraspecific  competition)  in  that  
contest  always  involves  interference,  while  scramble  usually  involves  exploitation.  
Interference  and  exploitation,  however,  have  none  of  the  extreme  threshold  
characteristics  associated  with  scramble  and  contest  competition.  
-­‐   competitive  exclusion  happens:  

●   winner  often  depends  on  environment.  

-­‐   coexistence  is  possible  if:  

●   there  is  niche  differentiation,  and  

●   the  system  is  not  at  equilibrium.  

   
Lecture  8:  RELATIONS  AMONG  SPECIES  
Module:  Principles  of  Ecology  
Dept.  of  Applied  Biosciences  and  Biotechnology,  Midlands  State  University  

8.1.   Introduction  
8.2.   Commensalism  
8.3.   Mutualism  

___________________________________________________________________________  
   
RELATIONS  AMONG  SPECIES  

Introduction  
We  define  ecology  as  the  scientific  study  of  the  interactions  that  determine  the  abundance  
and  distribution  of  organisms.  

These  interactions  are  with:    

-­‐   the  abiotic  environment  (e.g.  temperature,  pressure,  pH,  pO2),  and    

-­‐   the  biotic  environment  (i.e.  other  organisms).  

Theoretically,  populations  of  two  species  may  interact  in  basic  ways  that  correspond  to  
combinations  of  0,  +  and  –  as  follows:  

00,  ++,  +0,  -­‐0,  and  +-­‐  

Where:  

+  stands  for  benefits  

-­‐stands  for  suffers,  and  

0  stands  for  unaffected.  

The  five  major  types  of  interactions  are  as  follows:  

COMPETITION   -­‐-­‐   BOTH  COMPETITORS  SUFFER  


PREDATION   +-­‐   PREDATOR  BENEFITS,  PREY  SUFFERS  
PARASITISM   +-­‐   PARASITE  BENEFITS,  HOST  SUFFERS  
COMMENSALISM   +0   ONE  BENEFITS,  THE  OTHER  UNAFFECTED  
MUTUALISM   ++   BOTH  MUTUALISTS  BENEFIT  
 

Three  of  these  combinations  (++,  -­‐-­‐  and  +-­‐)  are  commonly  subdivided,  resulting  in  9  types  of  
interactions.  These  are  as  follows:  

(i)   Neutralism  (00)  =  in  which  neither  population  is  affected  by  association  with  the  
other.  
(ii)   Competition:  Direct  interference  type  (-­‐-­‐)  =  in  which  both  populations  actively  
inhibit  each  other.    
(iii)   Competition:  Resource  use  type  (-­‐-­‐)  =  in  which  both  populations  suffer  when  a  
common  resource  is  in  short  supply.    

(iv)   Ammensalism  (-­‐0)  =  in  which  one  population  is  inhibited  and  the  other  is  not  
affected.    
(iv)   Predation  and  Parasitism  (+-­‐)  =  in  which  one  population  adversely  affects  the  
other  by  direct  attack  but  nevertheless  dependent.    
(v)   Commensalism  (+0)  =  in  which  the  commensal  benefits  while  the  host  is  not  
affected.    
(vi)   Protocoorperation  (++)  =  in  which  both  populations  benefit  by  the  association  
but  relations  are  not  obligatory.  
(vii)   Mutualism  (++)  =  in  which  the  interaction  is  favourable  to  both  populations  and  
the  association  is  obligatory.  

COMMENSALISM  
Commensalism  is  an  interaction  in  which  one  organism  (or  species)  beneficially  affects  the  
second  organism  (or  species),  but  the  second  has  no  effect  on  the  first.  

Commensalism  often  involve  one  organism  that  is  already  dead,  which  is  then  consumed  by  
another  (decomposition,  detritivory),  or  the  consumption  of  the  waste  products  of  one  
species  by  another  (e.g.  coprophagy).  

Decomposition  is  the  breakdown  of  complex  energy  rich  organic  molecules  to  simple  
inorganic  constituents  –  also  called  “mineralisation”  and  is  an  important  step  in  material  
recycling  (nutrient  regeneration).  

Decomposers  are  typically  bacteria  and  fungi.  The  later  may  dominate  in  acidic  habitats,  
while  the  former  may  dominate  in  more  basic  sites.  

Bacteria  dominate  in  the  decomposition  of  the  more  easily  degradable  “parts”  (or  
molecules)  of  dead  organisms  (e.g.  simple  carbohydrates,  lipids,  proteins  and  nucleic  acids);  
while  fungi  dominate  in  the  decomposition  of  hard  to  degrade  molecules  (e.g.  celluloses,  
hemicelluloses  and  lignins  of  woody  plants).  Forest  mushrooms  are  fruiting  bodies  of  the  
fungi  that  breakdown  wood  and  leaves.  

Detritivory  is  the  consumption  of  dead  organic  matter  (detritus)  usually  together  with  the  
associated  decomposers  (bacteria  and  fungi).  

Detritivory  can  also  include  carrion  feeders.  These  range  from  scavengers  (like  vultures)  
most  of  which  can  also  act  as  predators  if  the  opportunity  arises,  to  carrion  feeding  insects  
like  blowflies  in  which  the  adults  find  carcasses  by  smell,  lay  their  eggs  on  or  in  the  carcasses  
after  which  the  maggots  (larvae)  consume  the  flesh,  pupate  and  emerge  as  adults.  

Burying  beetles  of  the  genus  Necrophorus  are  also  carrion  feeders.  They  will  burry  small  
mammal  carcasses  they  will  find  (to  reduce  competition  with  blowflies)  and  then  tear  off  
and  consume  pieces  of  flesh  and/or  eat  any  blowfly  larvae  they  encounter.  
Most  streams  and  some  lakes  derive  most  of  their  organic  matter  from  the  surrounding  
watershed,  especially  dead  plant  parts  (mostly  leaves)  of  terrestrial  plants  that  fall  into  
streams.  These  support  a  diverse  assemblage  of  aquatic  insect  larvae  and  other  
invertebrates.  

Coprophagy  

 
While  the  guts  of  herbivores  are  longer  than  those  of  carnivores,  plant  matter  is  typically  
more  resistant  to  digestion  than  are  animal  tissues  and  the  faeces  of  herbivorous  
vertebrates  still  contain  significant  quantities  of  organic  matter.  

Some  animals  have  specialised  on  feeding  on  the  faeces  of  other  animals,  e.g.  the  large  
African  dung  beetle  (Heliocopris  dilloni)  feeds  on  the  dung  of  African  elephants.  

The  winged  adults  of  the  beetle  smell  fresh  dung,  fly  to  the  site,  eat  some  of  the  dung  
immediately,  and  cut  off  pieces  that  they  roll  away  and  burry  in  the  ground.  

The  adult  female  lays  a  single  egg  in  each  of  the  dung  balls.  The  larva  then  hatches  out  and  
consumes  the  dung  ball,  pupates  and  then  emerges  as  an  adult  to  look  for  fresh  dung  to  lay  
its  eggs.  

MUTUALISM  
Mutualistic  relationships  are  mutually  beneficial  relationships,  in  which  individuals  of  both  
species  have  higher  fitness  in  the  presence  of  the  other  species.  

Fitness  is  the  relative  contribution  that  an  individual  makes  to  the  gene  pool  of  the  next  
generation.  Fitness  is  very  difficult  to  determine  in  most  cases  –  rather,  we  take  short-­‐cut  
measures  to  estimate  fitness,  such  as:  

-­‐   growth,    

-­‐   survivorship,  and  


-­‐   fecundity  (reproductive  output).  

Mutualisms  generally  involve  one  species  gaining  access  to  resources  (food  or  inorganic  
nutrients)  while  the  other  is  provided  with  protection  from  enemies,  a  favourable  
environment,  or  a  service  (e.g.  pollination,  seed  dispersal,  rid  of  parasites,  etc.).  

Coevolution  of  mutualistic  relationships  


-­‐   Interacting  species  often  show  reciprocal  changes  through  time.  

-­‐   Coevolution  –  each  species  must  change  in  response  to  the  changes  in  the  other  
species.  

-­‐   Symbiosis  (=same  lives)  is  the  ultimate  expression  of  coevolution  –  two  species  live  
as  one,  often  one  within  the  cells  or  tissues  of  the  other.  

-­‐   a  symbiosis  is  an  intimate,  long-­‐lived  relationship  and  often  obligatory  association  of  
two  species.    

Mutualisms  can  be  divided  into  two  broad  categories:  Non-­‐symbiotic  and  symbiotic  
mutualisms.  

Facultative  vs  Obligate  Mutualisms  


-­‐   in  facultative  mutualisms,  each  species  gains  a  benefit  from  the  association  but  can  
live  on  its  own.  

-­‐   in  obligate  mutualisms,  on  the  other  hand,  each  species  can  only  live  in  the  presence  
of  the  other.  

Nonsymbiotic  Mutualisms  
Protection  (or  Defensive)  Mutualisms  
In  this  case,  one  species  is  defended  against  predators,  parasites  or  competitors  by  the  
other,  which  may  in  turn,  receive  shelter,  etc.  

Ant/Plant  Protection  Mutualisms  

-­‐   Seen  in  many  taxa  of  plants  but  especially  well  developed  in  the  South  American  
Acacia  and  its  associated  ant  species.    

-­‐   Ants  protect  Acacia  trees  from  herbivorous  insects  by  actively  attacking  any  that  
land  on  the  plant.  Ants  also  protect  the  Acacia  from  competitors  by  cutting  off  any  
leaves  and  branches  of  nearby  plants.  

-­‐   In  return,  Acacia  trees  provide  the  ants  with:    


(i)   nesting/brooding  sites  in  the  form  of  hollow  thorns  into  which  the  queen  
ant  bores  

(ii)   protein-­‐rich  food  in  the  form  of  “beltian  bodies”  produced  at  the  tips  of  the  
leaves,  and  

(iii)   carbohydrate-­‐rich  food  in  the  form  of  “nectar”  secreted  by  nectaries  at  the  
leaf  bases.  

-­‐   South  American  Acacias  lack  more  traditional  plant  defences  such  as  spines  and  
secondary  metabolites  to  deter  herbivores  and  so  they  depend  on  ants.  

Farming  of  Fungi  by  Leafcutter  Ants  

-­‐   Leafcutter  ants  defoliate  trees,  carry  the  leaf  pieces  back  to  underground  burrows  
they  have  dug,  chew  the  leaves  and  hang  the  chewed  matter  from  the  roofs  of  their  
galleries  where  they  become  infected  with  a  fungus.  The  fungus  “digests”  the  tough  
leaves  and  is  harvested  for  food  for  larval  and  adult  ants.  

-­‐   The  fungus  gains  a  constant  supply  of  “food”  and  an  agent  for  dispersal.  When  the  
ants  move  to  the  new  burrow,  the  queen  takes  a  supply  of  the  fungus  with  her  to  
start  new  “gardens”.  The  fungus  is  never  found  in  the  absence  of  the  ants.  

-­‐   Harvester  termites  which  lack  gut  symbionts  found  in  other  termites  to  break  down  
cellulose,  produce  subterranean  fungal  gardens  in  a  similar  manner.  

Farming  of  Fungi  by  “Ambrosia”  beetles  


-­‐   One  group  of  beetles  exhibits  mutualisms  with  fungi  similar  to  that  of  leafcutter  ants,  
however,  here  the  food  for  the  fungus  is  the  wood  of  the  dead  and  dying  trees.  The  
beetles  burrow  into  the  wood  forming  a  network  of  tunnels,  which  it  infects  with  
fungi  carried  in  its  gut  or  attached  to  specialised  “hairs”.  The  beetles  visit  the  tunnels  
periodically  to  harvest  the  ”ambrosia”  for  food.  

-­‐   Again  the  fungus  gains  a  supply  of  food  and  a  means  of  dispersal.  

-­‐   The  larvae  of  wood  wasps  live  similarly.  

Farming  of  Aphids,  Leafhoppers  Scale  insects  and  other  bugs  (Homoptera)  by  
Ants    

-­‐   Numerous  species  of  true  buds  (Homoptera),  such  as  aphids,  treehoppers,  
leafhoppers,  and  scale  insects,  feed  on  phloem  sap  of  vascular  plants.  However,  
since  they  must  pierce  the  plant  with  tubular  mouthparts  which  cannot  be  rapidly  
withdrawn,  they  are  vulnerable  to  predation  by  other  insects  when  feeding.  
-­‐   Many  of  these  homopterans  have  mutualistic  relationships  with  ants.  The  ants  
protect  the  homopterans  from  predators.  

-­‐   In  return,  the  ants  drink  the  “honeydew”  (dilute  sugary  faeces)  constantly  produced  
by  the  bugs.  The  ants  often  stimulate  the  bugs  to  release  the  honeydew  by  stroking  
them  with  their  antennae.  The  bugs  must  process  large  quantities  of  sap  so  as  to  
ensure  that  they  obtain  enough  protein  and  are  able  to  leave  some  sugars  
undigested  for  the  ants.  

Farming  of  Caterpillars  (Lepidoptera)  by  Ants    

-­‐   Caterpillars  of  the  butterfly  family  Lycaenidae  have  mutualistic  relationships  with  
ants.  These  relationships  are  most  complex  in  the  blue  butterflies  where  the  
relationships  are  between  single  species  pairs  of  butterflies  and  ants.  

-­‐   The  caterpillars  feed  on  plants  up  to  their  third  instar,  then  fall  to  the  ground.  The  
caterpillars  mimic  ant  larvae  chemically  and  behaviourally  and  they  are  picked  up  by  
ants  and  carried  back  to  their  nest.  

-­‐   The  ants  protect  the  caterpillars  from  predators  and  parasitoids,  as  well  as  feeding  
them  with  plant  material.  

-­‐   In  return,  the  caterpillars  supply  the  ants  with  sugary  secretions  produced  by  the  
“honey  glands”  when  they  are  stroked  by  the  ant’s  antennae.  The  caterpillar  
secretions  may  also  contain  an  addictive  chemical.  

-­‐   The  relationship  between  the  butterfly,  ant,  and  the  plant  on  which  they  feed  may  
be  so  restrictive  that  loss  of  any  one  species  may  cause  extinction  of  all  the  three,  
e.g.,  loss  of  grasslands  in  England  has  dramatically  reduced  the  amount  of  wild  
thyme,  the  plant  on  which  the  blue  butterfly  and  its  two  species  of  attendant  ants  
depend.  The  butterfly  is  now  extinct  in  England.  

Goby  fish/Burrowing  Shrimp  Mutualism    

-­‐   Marine  shrimp  of  the  genus  Alpheus  have  mutualisms  with  goby  fish  of  the  genus  
Cryptocentrus.  

-­‐   The  shrimp  digs  burrows  which  the  fish  use  for  shelter.  

-­‐   The  shrimp,  which  is  blind,  keeps  one  antennae  constantly  in  contact  with  the  fish  
when  outside  the  burrow.  This  provides  the  shrimp  with  early  warning  of  danger.  

Mixed  Species  Aggregations  


-­‐ Mixed  species  aggregations  such  as  the  herds  seen  on  the  Serengeti  Plains  (with  
Zebra,  Wildebeest  and  Gazelle,  for  example),  achieve  increased  vigilance  against  
predators  as  the  different  species  combined  will  have  better  vision,  hearing,  smell,  
etc.  

Sea  Anemone/Clown  fish  Mutualisms  

 
Many  species  of  clown  fish  associate  with  sea  anemones.  The  clown  fish  gains  protection  by  
retreating  into  the  sea  anemone’s  tentacles  when  threatened.  They  accumulate  slime  from  
the  sea  anemone  to  prevent  triggering  the  nematocysts  themselves.  

The  sea  anemones  also  gain  protection,  as  the  clown  fish  attacks  fish  species  that  normally  
eat  sea  anemones.  

Cleaning  Mutualisms  

-­‐   A  number  of  coral  reef  fish  species  function  as  “cleaners”  of  other  often  larger  fish.    
They  set  up  “cleaning  stations”  on  the  reef  to  which  they  attract  “customers”  with  
their  body  posture.  

-­‐   The  customers  allow  the  cleaners  to  move  over  their  body  surface  and  even  into  
their  mouth  and  gill  chambers  to  pick  out  and  eat  parasites  attached  to  them.  

-­‐   The  cleaner  gains  a  source  of  food  while  the  customer  is  cleaned  of  parasites.  

-­‐   In  Africa  ,  birds  known  as  the  oxpeckers  remove  ticks  and  other  parasites  from  large  
mammals.  Recently,  oxpeckers  have  disappeared  from  certain  areas  as  they  have  
been  poisoned  by  feeding  on  ticks  attached  to  cattle  that  were  dipped  to  kill  
parasites.  The  wild  animals  have  suffered  tremendously  as  ticks  have  since  increased  
dramatically.  

Transport  (or  Dispersive)  Mutualisms  


Plants  have  developed  mutualistic  and  parasitic  relationships  with  animals  for  the  
movement  of  their  gametes  (especially  sperm  in  pollen)  and  propagules  (especially  seeds).  

Pollination  Mutualisms  
-­‐   Wind  pollination  is  very  wasteful.  Most  of  the  pollen  never  reaches  another  flower.  
Inducing  an  animal  (insect,  bird,  bat,  rodent,  etc.)  to  visit  flowers  by  providing  a  food  
reward  in  the  form  of  nectar  (a  sugary  solution)  or  excess  pollen  (high  protein)  
provides  a  more  efficient  way  of  delivering  pollen  from  one  flower  to  the  next.  

-­‐   Pollinators  and  flowers  may  be  generalists  (each  interacting  with  many  species  of  the  
other)  or  specialists.  If  pollinators  and  flowers  are  specialists,  efficiency  in  pollen  
delivery  is  even  greater  as  none  is  wasted  on  flowers  of  the  “wrong”  species.  
However,  specialisation  which  can  even  lead  to  obligate  mutualisms  (e.g.  between  
the  figs  and  the  fig  wasps),  makes  each  species  prone  to  extinction  should  the  other  
be  wiped  out.  

Seed  Dispersal  Mutualisms  


-­‐   Many  plants  produce  fleshy  fruits  as  an  inducement  for  birds  and/or  mammals  to  
consume  their  seeds.  The  seeds  are  then  carried  for  some  time  or  distance  before  
being  regurgitated  or  excreted,  hence  dispersing  the  seeds.  The  seeds,  however,  
must  be  digestion  resistant.  

-­‐   Most  plants  have  their  seeds  consumed/dispersed  by  a  great  variety  of  species.  A  
few  cases  of  specialisation  are  known,  however.  For  example,  some  Australian  plants  
produce  seeds  with  edible  appendages  (called  elaiosomes)  which  are  eaten  only  by  
ants.  Many  seeds  are  carried  some  distance  by  the  ants  before  elaiosomes  are  eaten.  

A  Miscellaneous  Nonsymbiotic  Mutualism  


Honey  Guide/Honey  badger  Mutualism  

 
 
When  the  African  bird  known  as  the  honey  guide  (Indicator  indicator)  finds  a  bee’s  nest,  it  
flies  around  looking  for  a  honey  badger  (also  known  as  the  ratel,  Melliovora  capensis),  or  a  
human  being,  and  guides  it  back  to  the  nest.  The  powerful  mammal  rips  open  the  nest  and  
eats  the  honey  and  larvae.  When  it  is  done,  the  bird  then  comes  in  and  eats  the  beeswax  
and  the  remaining  larvae.  

The  bird  has  gut  mutualists  capable  of  digesting  wax.  It  was  first  described  stealing  candles  
from  mission  churches.  

Symbiotic  Mutualisms  
Digestive  Symbioses  (Gut  Mutualisms)  
While  technically  outside  the  tissues  of  their  hosts,  gut  symbionts  or  mutualists  aid  the  
digestion  of  complex  molecules  in  a  number  of  animals  and  synthesize  important  organic  
molecules  like  vitamins  needed  by  their  host  using  the  host’s  food  as  raw  materials.  

Ruminant  Mammals/Ciliates/Bacteria  
-­‐ Ruminants  (dear,  cattle,  antelope,  etc.)  which  feed  primarily  on  hard-­‐to-­‐digests  
grasses,  have  four-­‐chambered  stomachs.  The  rumen  contains  a  complex  community  
of  cellulolytic  bacteria  and  protozoans,  most  of  which  are  specialists  that  cannot  live  
elsewhere.  The  ruminants  use  the  fermentation  products  of  their  mutualists  as  food.  

Termites/Flagellates/bacteria  
-­‐   While  75%  of  termite  species  produce  their  own  cellulose  to  digest  their  food  
(wood),  and  some  species  cultivate  fungal  gardens,  other  species  possess  a  diverse  
assemblage  of  mutualists  within  out-­‐pockets  of  their  guts.  

-­‐   Anaerobic  flagellate  protozoans  ingest  wood  particles  and  ferment  them,  releasing  
acetic  acid  and  other  organics  used  by  the  termite.  

-­‐   The  flagellates  are  covered  by  two  types  of  bacteria:  spirochaetes,  which  provide  
locomotion  for  the  flagellate  in  exchange  for  “food”  (there  are  so  many  on  the  
surface  of  the  protozoans  that  they  were  initially  classified  as  ciliated)    and  nitrogen-­‐
fixing  bacteria  which  fix  gaseous  nitrogen  in  exchange  for  food.  

General  Gut  flora  


-­‐ Bacteria  in  the  guts  of  animals  may  synthesize  vitamins  and  other  molecules  which  
their  hosts  may  need  but  cannot  make  on  their  own.  

Photosynthetic  Symbioses  
In  some  “marginal”  habitats,  inorganic  nutrients  needed  by  photoautotrophs  are  scarce  as  
are  “prey”  for  heterotrophs.  Mutualisms  between  these  groups  provide  a  close  coupling  
with  inorganic  nitrogen,  phosphorus,  etc.  passing  from  animals  (which  secrete  them  as  
wastes)  to  “plants”  (where  they  are  nutrients)  and  some  photosynthate  passing  from  
“plants”  to  animals.    These  mutualisms  typically  involve  algae  living  within  tissues  or  cells  of  
animals.  

Reef-­‐building  Corals/”Zooxanthellae”  
-­‐   Corals  (coelenterate)  polyps  contain  symbiotic  algae  in  their  tissues  to  “harvest  
sunlight”.  The  symbiotic  algae  of  most  cnidarians  are  non-­‐motile  dinoflagellates,  
called  Zooxanthellae.  

-­‐   The  zooxanthellae  also  provide  their  hosts  with  excess  photosynthate.  

Sea  Anemones&  Jellyfish/Zooxathellae  &  Zoochlorellae”  


-­‐ A  number  of  species  of  sea  anemones  and  jellyfish  (coelenterate  or  cnidarians)  
contain  symbiotic  dinoflagellates  (“Zooxanthella”)  or  green  algae  (“Zoochlorella”).  In  
cnidarians,  these  are  contained  in  the  gastrodermal  cells  and  they  pass  on  to  their  
host,  excess  photosynthate,  in  addition  to  harvesting  sunlight.  

“Green  Hydra”  (Hydra  viridis)  /  Chlorella  sp.  


-­‐ The  freshwater,  solitary  coelenterate  Hydra  viridis  contains  symbiotic  green  algae  of  
the  genus  Chlorella  within  its  cells.  The  algae  can  provide  their  hosts  with  oxygen  as  
well  as  photosynthate.    

Paramecium  bursaria/Chlorella  sp.  


-­‐ The  single-­‐celled  ciliate  protozoan  Paramecium  bursaria  contains  numerous  
intracellular  symbiotic  cells  of  green  algae  of  the  genus  Chlorella.  Not  only  do  the  
algae  provide  their  hosts  with  photosynthate,  they  also  supply  them  with  oxygen  in  
hypoxic/anoxic  environments.    

“Lichens”:  Fungi/Algae  

-­‐   About  25%  of  fungal  species  can  be  lichenized,  i.e.  infected  with  algae  to  form  
lichens.  These  can  occupy  extremely  poor  habitats  such  as  bare  rock  and  tree  trunks.  

-­‐   The  algae  provide  photosynthate  (and  possibly  fixed  nitrogen)  to  the  fungus.  The  
fungus  might  serve  the  alga  by  efficiently  taking  up  inorganic  nutrients  from  the  
environment,  even  from  dilute  rainwater.  

Nutrient  Uptake  Symbioses  


“Lichens”:  Fungi/Algae  

Mycorrhizae:  Plants/Fungi  

The  roots  of  many  plants  form  intimate  relationships  with  various  fungi.  The  fungi  typically  
penetrate  into  the  tissues  of  the  root  or  even  into  individual  cells.  

Fungal  hyphae  extend  into  the  organic  upper  layers  of  the  soil  and  take  up  inorganic  
nutrients  from  the  soil,  passing  them  on  to  the  plant.  In  exchange,  the  plant  passes  some  
photosynthate  to  the  fungus.  

Mycorrhizal  fungi  can  extend  between  many  plants,  even  trees  of  different  species.  

Chemosynthetic  Symbioses  
Red  tube  worm  (Phylum  Pogonophora)/chemoautotrophic  Bacteria  

-­‐   Deep  sea  hydrothermal  vents  release  sulphides  which  support  unique  communities  
of  pogonophorans  (red  tube  worms),  polychaete  worms,  crabs  and  large  molluscs.  

-­‐   The  red  tube  worms,  which  lack  a  mouth  and  a  digestive  tract,  have  a  mass  of  tissue,  
called  the  trophosome  that  is  packed  with  symbiotic  bacteria  (chemosynthetic  
bacteria).  The  bacteria  oxidize  sulphides,  producing  carbon-­‐containing  compounds  
which  they  share  with  their  pogonophoran  host.  

-­‐   The  vascular  haemoglobin,  which  accounts  for  the  red  colour  of  these  large  worms,  
is  important  in  delivering  the  large  amounts  of  oxygen  required  by  the  bacteria.  

-­‐   The  worm  thus  appears  to  provide  a  safe  habitat  with  appropriate  conditions  for  the  
bacteria.  

-­‐   Molluscs  have  their  own  chemosynthetic  bacteria  in  their  gills.  

Nitrogen-­‐fixation  Symbioses  
In  many  habitats  photoautotrophs  are  limited  by  nitrogen  (or  phosphorus).  While  
eukaryotes  lack  the  enzymes  to  ‘fix”  the  abundant  atmospheric  nitrogen  (N2)  into  an  organic  
form,  this  ability  is  widespread  in  the  prokaryotic  bacteria.  Mutualism  between  nitrogen-­‐
fixing  bacteria  and  some  plants  ensures  that  the  later  has  an  adequate  supply  of  nitrogen.    

Legume/Rhizobia  

-­‐   A  number  of  bacterial  genera  collectively  known  as  Rhizobia  form  complex  
relationships  with  legumes.  
-­‐   The  presence  of  the  bacteria  triggers  the  root  cells  to  produce  root  nodules,  in  which  
the  nitrogen-­‐fixing  bacteria  will  be  contained.  The  nodules  are  kept  hypoxic,  as  the  
nitrogenase  enzyme  only  functions  at  very  low  oxygen  concentrations.  

-­‐   “fixed”  nitrogen  (mainly  asparagine)  is  passed  on  to  the  plant.  In  return  the  plant  
passes  photosynthate  on  to  the  bacteria.  

Actinomycetes/non-­‐legume  symbioses  

-­‐ Actinomycets  of  the  genus  Frankia  form  nodules  on  the  roots  of  non-­‐legumes  such  
as  Alder  (Alnus),  which  is  an  important  early  colonist  in  some  habitats.  

Cyanobacteria/Plant  Symbioses  
-­‐ A  number  of  plant  species  from  different  groups  form  symbioses  with  “blue-­‐green  
algae”  (prokaryotic  cyanobacteria),  some  of  which  can  fix  N2.  The  cyanobacteria  are  
housed  in  special  structures  on  the  plant  thallus,  leaves  or  roots.  

Cyanobacteria/Heterotrophic  Bacteria  Symbioses  

-­‐ Some  free-­‐living,  filamentous  cyanobacteria  like  Aphanizomenon  contain  special,  


thick-­‐walled  cells  called  heterocyst  which  are  the  sites  of  nitrogen  fixation.  To  ensure  
low  oxygen  concentration  in  the  heterocysts,  they  leak  photosynthate  which  attracts  
heterotrophic  bacteria  which  consume  most  of  the  oxygen  in  the  surrounding  water.  

Luminescence  Symbioses  
Luminescence  Bacteria/Animals  
-­‐   A  large  number  of  marine  fish  and  invertebrate  species  are  capable  of  producing  
light  (bioluminescence).  In  many  cases  the  fish  or  invertebrate  itself  produces  the  
chemicals,  but  in  other  cases,  light  is  produced  by  luminescent  bacteria  housed  in  
special  regions  of  the  animal.  

-­‐   The  light  produced  serves  the  anima  in  “counter-­‐shading”,  finding  food  (e.g.  the  
lantern  fish),  attracting  food  (e.g.  deep-­‐sea  Anglerfish),  attracting  mates,  etc.  

-­‐   The  animal  provides  the  bacteria  with  appropriate  conditions  and  food.  

Serial  Endosymbiosis  Theory  of  Eukaryotic  Evolution  


-­‐   In  1975,  Lynn  Margulis  proposed  that  eukaryotic  cells  evolved  by  intracellular  
symbiosis  between/among  prokaryotes.  

-­‐   Fermentative  anaerobes  could  become  infected  by  Krebs  cycle-­‐containing  


eubacteria.  The  later  becoming  mitochondria.  
-­‐   Motile  bacteria  (such  as  spirochaetes)  could  infect  the  cells  to  become  cilia  and  
flagella  (like  in  the  flagellate  in  the  termite  gut).  

-­‐   “ingestion”  of  photosynthetic  cyanobacteria  could  give  rise  to  chloroplasts,  etc.  

-­‐   The  fact  that  both  mitochondria  and  chloroplasts  have  their  own  DNA  is  support  for  
their  origin  as  intracellular  symbionts.  

-­‐   Given  the  success  of  eukaryotes,  these  symbiotic  mutualisms  (if  indeed  that  is  what  
they  are)  can  be  viewed  as  the  most  important  of  all.  

   
Lecture  9:  PARASITISM  
Module:  Principles  of  Ecology  
Dept.  of  Applied  Biosciences  and  Biotechnology,  Midlands  State  University  

9.1.   Introduction  
9.2.   Micro-­‐parasites  vs  Macro-­‐parasites  
9.3.   Endo-­‐parasites  vs  Ecto-­‐parasites  

___________________________________________________________________________  

   
PARASITISM  AND  INFECTIOUS  DISEASE  
In  parasitism  and  infectious  disease,  one  organism  (the  parasite  or  pathogen)  benefits  at  the  
expense  of  its  host.  

Taken  together  parasitism  and  infectious  disease  have  devastating  effects  on  the  health  of  
individuals  and  size  and  dynamics  of  most  (if  not  all)  wild  populations.  

A  parasite  is  an  organism  that  consumes  part  of  the  host’s  tissues,  usually  without  killing  the  
host  (at  least  immediately).  As  such,  parasitism  can  be  regarded  as  partial  predation.  

The  term  disease  refers  to  an  unhealthy  condition  of  an  individual  that  impairs  a  vital  
function.  An  infectious  disease  is  thus,  one  caused  by  an  infection  by  a  foreign  organism,  
and  is  different  from  autoimmune,  metabolic  or  genetic  diseases.  

A  disease  organism  is  an  organism,  usually  endoparasitic,  that  causes  a  disease  condition  in  
another  organism.  

Microparasites  vs  Macroparasites  


-­‐   Microparasites  (including  most  disease  organisms)  are  very  small  and  typically  can  
complete  their  entire  life  cycle  within  a  single  host.  They  may  rely  on  a  
vector”organisms,  however,  to  transport  their  propagules  to  infect  new  hosts.  

-­‐   One  example  of  a  microparasite  is  Plasmodium  spp.,  a  protozoan  which  causes  
malaria  in  humans.  The  vector  organisms  of  malaria  are  the  females  of  a  number  of  
species  of  mosquitoes.  

-­‐   Macroparasites  are  large  and  typically  live  on  the  surface  of  their  host  or  in  the  
host’s  gut  cavity.  However,  the  blood  flukes  (Schistosoma  spp.),  which  are  classified  
as  microparasites,  live  in  the  blood  vessels  of  their  hosts  and  cause  a  devastating  
disease  in  humans  called  schistosomiasis.  –e.g.  the  Lamprey  and  leeches  on  fish  and  
ticks  and  fleas  on  mammals.    

-­‐   Macroparasites  often  cannot  complete  their  entire  life  cycle  on  or  within  a  single  
host.  They  often  produce  eggs  or  dispersive  larvae  and  frequently  involve  an  
intermediate  host  in  their  life  cycle.  For  example,  schistosomes  infect  humans  as  
adults  and  fresh  water  snails  (the  intermediate  host)  as  one  larval  form.  The  adults  
(infecting  humans)  release  eggs  in  ponds  and  streams,  which  develop  into  ciliated  
miracidium  larvae,  which  in  turn,  seek  out  and  penetrate  snails,  transforming  into  
sporocysts  which  reproduce  asexually  to  produce  rediae  and  cercariae,  the  later  of  
which  leave  the  snail  to  infect  humans  (either  actively  or  passively  by  encysting  and  
being  eaten  by  the  host).  
-­‐   While  the  hosts  can  be  viewed  as  habitats  for  the  parasites  which  supply  them  with  
controlled  conditions  and  a  constant  supply  of  resources,  the  host  possess  defence  
mechanisms  in  the  form  of  resistant  surfaces,  immune  systems,  macrophages,  etc.  

-­‐   These  defences  co-­‐evolve  with  “escape”  systems  in  the  parasite  in  a  continual  “arms  
race”.  For  example,  immune  systems  of  hosts  produce  antibiotics  in  response  to  
“foreign”  proteins  on  the  surface  of  the  parasites.  These  then  bind  to  the  surface  of  
the  parasites  and  target  them  for  destruction  by  macrophages.  However,  parasites  
can  escape  this  control  by  coating  themselves  with  host  proteins,  living  inside  cells  
where  the  immune  system  cannot  reach  them,  or  constantly  changing  their  surface  
proteins  so  that  the  host  cannot  ‘keep  up”.  

-­‐   To  illustrate  the  normal  effect  host  defences  have  we  need  only  consider  the  cases  
where  disease  and  parasites  have  reached  new  host  populations  that  evolved  in  
their  absence.  The  diseases  may  devastate  these  ‘naïve”  populations  (e.g.  European  
humans  bringing  small  pox  and  other  diseases  to  the  native  Americans).  

Endoparasites  vs  Ectoparasites    


-­‐   Endoparasites  live  within  the  tissues  or  cells  of  their  host  –e.g.  most  infectious  
disease  organisms  including  Plasmodium  and  Schistosoma.  

-­‐   Ectoparasites  live  on  the  surface  of  their  host  –e.g.  the  Lamprey  and  leeches  on  fish  
and  ticks  and  fleas  on  mammals.  

   
Lecture  10:  PREDATION  
Module:  Principles  of  Ecology  
Dept.  of  Applied  Biosciences  and  Biotechnology,  Midlands  State  University  

10.1.   Introduction  
10.2.   Predation  
10.3.   Types  of  predators  
10.4.   Predator/Prey  coevolution  
10.5.   Adaptations  of  predators  and  prey  

___________________________________________________________________________  

 
   
PREDATION  
Introduction  
Predation  and  parasitism  are  examples  of  antagonistic  interactions  between  two  species  
which  result  in  negative  effects  on  the  growth  and  survival  of  one  of  the  populations.  
Predators  use  their  prey  as  a  source  of  food  only,  whereas  parasites  use  their  hosts  both  as  
a  food  source  and  a  habitat.  

Importance  of  the  study  of  predation  and  parasitism:  

-­‐   In  many  species  predation  and  parasitism  are  dominating  among  ecological  
processes.  The  dynamics  of  the  se  populations  cannot  be  predicted  and  understood  
without  considering  natural  enemies.  

-­‐   Pest  species  of  insects  and  weeds  can  be  suppressed  by  introduction  of  natural  
enemies  (biological  control).  

-­‐   Natural  enemies  may  cause  side  effects  in  pesticide  applications.  The  numbers  of  
arthropod  natural  enemies  may  be  reduced  due  to  pesticide  treatment  which  may  
result  in  increasing  pest  populations.  

Predation  
Predation  is  the  consumption  of  one  organism  (the  prey),  in  whole  or  in  part,  by  another  
organism  (the  predator),  in  which  the  prey  is  alive  when  the  predator  first  attacks  it  (i.e.  not  
detritivory).  

Predation  benefits  the  predator  (which  gains  a  resource  for  survival,  growth  and  
reproduction)  but  is  costly  to  the  prey  (which  suffers  mortality  or  injury).  

A  predator  will  require  more  than  one  prey  individual  in  order  to  grow  so  that  its  search  for  
prey  is  continuous.  

Types  of  Predators  


a  “functional”  classification  

-­‐   True  Predators  –  are  animals  that  consume  other  animals.  They  kill  prey  more  or  less  
immediately  after  attack;  eat  many  prey  individuals  per  life  time;  usually  consume  
the  entire  prey  –  e.g.  lions,  cheetahs,  eagles,  snakes,  etc.    

-­‐   Herbivores  or  grazers  –  are  animals  that  eat  plants;  rarely  kill  prey;  ear  only  parts  of  
many  prey  individuals  per  life  time,  called  browsers  if  they  eat  woody  parts  of  
woody  plants,  e.g.  sheep,  cattle,  goats,  etc.  
-­‐   Parasites  –  are  organisms  (animals  or  plants)  that  live  in  an  obligatory,  close  
association  with  a  host  individual;  rarely  killing  prey  (hosts)  outright;  eat  only  part  of  
one  or  very  few  prey  individuals  per  life  time,  e.g.  tapeworms,  liver  flukes,  leaf-­‐
mining  and  leaf-­‐galling  insects,  various  bacterial  and  fungal  diseases.  

-­‐   Parasitoids  –  are  insects,  mostly  wasps  (Hymenoptera)  and  flies  (Diptera),  which  are  
free-­‐living  in  their  adult  stage,  but  lay  eggs  on  or  within  their  host.  The  larval  
parasitoid  begins  eating  the  host,  initially  doing  little  harm,  but  eventually  almost  
totally  consuming,  and  therefore  killing  the  host.  Parasitoids  eat  one  prey  individual  
per  life  time.  

a  taxonomic  classification  
-­‐   Carnivores  –  consume  animals  

-­‐   Herbivores  –  consume  plants  

-­‐   Omnivores  –  consume  both  plants  and  animals  

Predator/Prey  Coevolution  
The  act  of  predation  affects  the  survival  and  ultimately  fecundity  of  both  predators  and  
prey.  The  negative  effects,  however,  tend  to  be  quantitatively  small  where  the  interacting  
populations  have  had  a  common  evolutionary  history  in  a  relatively  stable  ecosystem.  
Natural  selection  tends  to  lead  to  a  reduction  in  detrimental  effects  or  to  the  elimination  of  
the  interaction  altogether,  since  continued  severe  depression  of  the  prey  population  by  the  
predator  can  only  lead  to  the  extinction  of  one  or  both  populations.  Consequently,  severe  
interactions  are  most  frequently  observed  when  the  interaction  is  of  a  recent  origin  or  when  
there  has  been  large-­‐scale  or  sudden  changes  in  the  ecosystem.  

Predators  and  their  prey  are  likely  to  have  co-­‐evolved  and  there  is  a  continuous  selection  
pressure  on  prey  to  avoid  death  at  the  hands  of  their  predators,  and  a  reciprocal,  
continuous  selection  pressure  on  predators  to  increase  their  predation  efficiency.  

Adaptations  of  Predators  and  Prey  


Since  the  act  of  predation  affects  the  survival  and  fecundity  of  both  predators  and  prey  (i.e.  
affects  their  fitness),  we  see  strong  selection  pressure  for  adaptations  that  increase:  

-­‐   the  efficiency  of  predation  in  predator  populations,  and    

-­‐   the  changes  of  escape  or  evasion  in  the  prey  populations.  

Predation  can  be  viewed  as  a  multi-­‐step  (or  multi-­‐stage)  process  and  the  adaptations  to  
predation  of  predators  and  prey  are  stage-­‐specific  (co-­‐evolution):  
STAGE   PREDATOR   PREY  COUNTER-­‐
ADAPTATIONS   ADAPTATIONS  
Encounter   Encounter  behaviours   Rarity;  apparent  rarity  
(ambush  or    search);  search  
(hiding  and  refuge  use,  
or  set  up  an  ambush  in  a  
being  active  when  
limited  area  where  prey  is  
predators  are  not);  
abundant   Movement,  e.g.  
seasonal  migration  
away  from  centres  of  
predator  density);  
Aggregation  
Detection   Sensory  acuity;  search  image   Crypsis;  Immobility;  
Polymorphism;  detect  
predator  first  and  move  
away;  Dispersion  
Recognition/Selection   Image  of  “food:;  assess   Polymorphism;  
profitability  (discrimination)   masquerading  as  an  
learning   inedible  object;  
Warning  colourations  
(Aposematism);  
Batesian  mimicry,  
Mullerian  mimicry  
Attack/Capture   Motor  skills  (speed,  agility);   Motor  skills,  escape  
offensive  weapons  (see   behaviours  
below)   (unpredictable  
movements;  flee  to  
cover;  startle,  bluff  or  
threaten  predator);  
Aggregation  (predator  
saturation)  
Handling/Subduing   Subduing  skills  and  offensive   Strength  to  escape;  
weapons  (e.g.  teeth,  claws,   Active  defence  
filter  sieves);  strength  to   (offensive  weapons);  
detain;  killing  skills;  venom   Group  defence;  Spines;  
Tough  integument;  
Resistance  to  venom  
Consumption   Tools  to  dismember  prey   Protection  for  safe  
(teeth,  jaws  and  claws);   passage  through  the  
detoxification  ability;   gut;  Toxins;  Spines;  
digestive  tract   Emetics  
Encounter  stage  
-­‐   In  order  to  encounter  prey  predators  may  lie  in  wait  or  move  around  in  search  of  
prey.  In  other  words  Ambush  predators  may  set  up  in  a  localised  area  where  prey  is  
abundant,  e.g.  at  water  holes  while  others  have  to  move  around  in  search  of  prey.  
Which  mode  is  used,  however,  depends,  in  part,  on  the  movement  (or  lack  of  
movement)  of  the  prey  (i.e.  it  makes  no  sense  for  a  herbivore  to  be  an  ambush  
predator).  

-­‐   If  prey  are  rare  or  simply  “appear”  to  be  rare  (due  to  hiding  or  by  being  active  when  
predators  are  not),  the  chances  of  an  encounter  are  reduced.  

-­‐   Therefore  prey  species  can  reduce  encounters  with  predators  through  refuge  use.  
For  example:  

-­‐   Small  organisms  may  hide  under  rocks  or  in  holes.  

-­‐   Zooplankton  may  remain  in  deep  waters  during  the  day  to  avoid  detection  (at  
least  when  predators  are  most  active).  

-­‐   Some  herbivores  may  take  cover  in  the  dense  bushes.  

-­‐   Some  species,  however,  avoid  cover  so  that  they  can  keep  a  good  lookout  for  
predators,  but  others  need  it,  paradoxically,  for  the  purpose  of  hiding  from  
predators.  Thus,  while  the  Thompson  Gazelles,  Wildebeest  and  Zebras  like  the  open  
plains  and  may  well  be  less  vulnerable  there,  others  such  as  the  waterbucks,  
reedbucks  and  duikers,  take  shelter  in  places  with  denser  vegetation.  

-­‐   Refuge  use,  however,  may  be  costly  to  the  prey  in  that  they  may  be  unable  to  search  
for  their  own  food  when  in  hiding.  

-­‐   Prey  may  also  reduce  encounters  with  predators  by  being  active  when  predators  are  
not,  e.g.  prey  species  can  adopt  nocturnal  feeding  habits  in  order  to  reduce  
encounters  with  predators  which  hunt  during  the  day.  

-­‐   The  movement  of  the  prey  may  “be  designed”  (i.e.  shaped  by  natural  selection)  to  
limit  encounters  with  predators,  e.g.  seasonal  migration  away  from  centres  of  
predator  density.  In  some  cases,  herbivores  migrate  and  their  predators  do  not.  This  
means  that  the  predators  only  live  in  plenty  when  the  game  passes  through.  

Detection  stage  
-­‐   Predators  often  have  acute  senses  (sight,  smell,  hearing,  etc.)  to  help  them  detect  
prey;  they  also  develop  “search  images”  (e.g.  symmetry,  contrast,  etc.)  to  help  them  
detect  prey  “signals”  against  the  environmental  “background”.  
-­‐   Conversely,  prey  often  have  adaptations  to  reduce  their  chances  of  being  detected  
by  predators.  These  include:  

-­‐   Cryptic  colouration  or  camouflage  –  this  is  a  passive  but  essential  defence  
strategy  that  makes  potential  prey  difficult  to  spot  against  its  “background”.  
A  camouflaged  animal  needs  only  remain  still  on  an  appropriate  substratum  
to  avoid  detection.  Camouflage  may  also  include  transparency  and  counter-­‐
shading  in  pelagic  species.  

-­‐   Immobility  (or  Akinesis)  makes  prey  less  detectable  by  predators  whose  
vision  responds  best  to  motion.  

-­‐   Polymorphism  –  if  prey  are  polymorphic  (i.e.  the  population  or  species  
consists  of  individuals  with  many  different  forms  or  colours)  this  can  
confound  the  predator’s  search  image.  

-­‐   Sensory  acuity  –  if  prey  have  acute  senses  as  well,  they  may  detect  the  
predators  first  and  move  away.  

-­‐   Interspecific  co-­‐operation  (or  mixed  species  aggregations)  e.g.  Impala  and  
Baboon  vs  leopard  and  the  mixed  species  aggregations  of  Zebras,  
Wildebeests  and  Gazelles  (in  the  Serengeti)  ensures  increased  vigilance  
against  predators  as  the  species  combined  will  have  better  vision,  hearing,  
smell,  etc.  for  example,  the  Impala/baboon  combination  relies  on  the  
impala’s  good  sense  of  smell  and  the  baboon’s  good  eye  sight  to  escape  from  
the  leopard,  i.e.  they  are  able  to  detect  the  leopard  first  and  move  away.  

-­‐   Dispersion  –  if  the  predators  require  more  than  one  individual  for  detection,  
dispersion  by  the  prey  may  be  advantageous.  

-­‐   Distraction  displays  -­‐  distraction  displays  direct  the  attention  of  the  predator  
away  from  a  vulnerable  prey,  such  as  a  bird  chick,  to  another  potential  prey  
that  is  more  likely  to  escape  such  as  the  chick’s  parent.  

Recognition  (selection)  stage  


-­‐   Typically  predators  do  not  “attack”  all  of  the  “potential”  prey  items  they  encounter.  
Most  are  selective  predators  to  some  extent  (some  extremely  so)  and  they  asses  the  
profitability  of  the  detected  prey  item  before  they  attack  –  what  constitutes  “good  
food”  can  be  shaped  by  learning.  

-­‐   Again  polymorphism  (in  the  prey  population)  may  foil  the  predator’s  image  of  “good  
food”  as  can  masquerading  as  an  inedible  object.  
-­‐   In  some  cases  prey  that  have  chemical  defences  that  affect  latter  stages  (e.g.  are  
distasteful  or  toxic  and  hence  rejected  as  they  are  being  consumed)  advertise  this  
fact,  often  by  being  brightly  coloured,  a  warning  to  predators  known  as  aposematic  
colouration  (or  aposematism).  This  is  prevalent  in  several  groups,  e.g.  “poison  
arrow”  frogs,  butterflies,  etc.  the  fire  Salamander,  which  can  squit  a  nerve  poison  
from  glands  on  its  back,  is  brightly  coloured  in  black  and  yellow  to  deter  would-­‐be  
predators.  

-­‐   Mimicry  –  since  predators  learn  to  associate  warning  colourations  with  
distastefulness/toxicity  and  the  refrain  from  attacking,  some  other  species  have  
evolved  to  mimic  the  colourations  of  these  protected  species  and  hence  gain  
protection.  

-­‐   Mimicry  is  a  phenomenon  in  which  the  mimic  bears  a  superficial  resemblance  to  
another  species,  the  model.  Defensive  mimicry  in  prey  often  involves  aposematic  
models.  

-­‐   In  Batesian  mimicry,  a  palatable  or  harmless  species  mimics  an  unpalatable  or  
harmful  model.  The  larva  of  the  hawkmoth,  for  example,  puffs  up  its  head  and  
thorax  when  disturbed,  looking  like  the  head  of  a  small  poisonous  snake  complete  
with  eyes.  The  mimicry  even  involves  behaviour,  the  larva  weaves  its  head  back  and  
forth  and  hisses  like  a  snake.  

   
-­‐   Additional  examples  of  Batesian  mimicry  are  the  many  harmless  snakes  that  mimic  
the  conspicuous  red,  white  and  black  markings  of  the  poisonous  coral  snake.  

-­‐   For  Batesian  mimicry  to  be  effective,  however,  the  models  must  generally  
outnumber  the  mimics,  otherwise  predators  would  learn  that  animals  with  a  
particular  colouration  are  good  rather  than  bad  to  eat.  

-­‐   Batesian  mimicry  was  named  after  its  discoverer,  Henry  Bates,  a  19th  century  English  
naturalist  and  explorer.  

-­‐   In  Mullerian  mimicry,  which  was  named  after  another  19th  century  explorer,  the  
German  Fritz  Muller,  two  or  more  unpalatable  aposematically  coloured  potential  
prey  species  resemble  each  other.  Each  acts  as  both  a  model  and  a  mimic.  Each  
species  gains  an  additional  advantage,  because  the  pooling  of  numbers  causes  
predators  to  learn  more  quickly  to  avoid  any  prey  with  a  particular  appearance.  

-­‐   Prey  may  employ  crypsis  and  akinesis  at  this  (and  the  following)  stage  too.  
Once  detected,  the  prey  may  stop  moving,  for  example,  if  the  predator’s  
vision  relies  on  motion  for  detection,  the  predator  may  lose  track  of  the  prey.  

Attack  (capture)  stage  


-­‐   Once  the  predator  encounters  a  prey  item  and  deems  it  acceptable,  it  will  “attack”.  
This  may  range  anywhere  from  a  lion  chasing  a  Gazelle  to  a  squirrel  picking  up  an  
acorn.  

-­‐   Predators  may  exhibit  some  rapid  locomotion  and/or  offensive  weapons  (teeth,  
claws,  filtering  sieves,  etc.)  to  help  them  approach  and  capture  prey.  

-­‐   Predators  may  also  use  mimicry  to  capture  prey.  For  example,  some  snapping  turtles  
have  tongues  resembling  a  wriggling  worm,  thus  luring  small  fish,  any  fish  that  tries  
to  eat  the  bait  is  itself  quickly  consumed  as  the  turtle’s  strong  jaws  snap  shut.  

-­‐   Motile  prey  may  attempt  to  escape.  Fleeing  is  perhaps  the  most  direct  antipredator  
response,  although  it  can  be  energetically  very  expensive.  During  the  course  of  
evolution  there  has  been  selection  for  fast  locomotion  and  some  prey  species  can  
exhibit  evasive  behaviours,  e.g.  unpredictable  movements.  

-­‐   Aggregation  by  prey  can  “saturate”  predators,  i.e.  diluting  risk  to  prey.  

-­‐   Prey  may  attempt  to  startle,  bluff,  or  threaten  the  predator,  e.g.  impalas,  the  ink  in  
the  Cephalopod  molluscs  can  confuse  predators.  

Handling  (subduing)  stage  


-­‐   Predators  show  adaptations  to  increase  ability  to  subdue  (or  even  kill)  prey,  e.g.  
physical  strength,  claws,  pointed  teeth,  etc.  lions,  for  instance,  will  use  their  physical  
strength,  claws  and  pointed  teeth  to  subdue  a  Wildebeest  and  then  go  for  the  neck  
to  suffocate  it.  

-­‐   Some  predators  (e.g.  many  snakes)  even  employ  venoms  (often  neurotoxins)  to  
immobilise  prey.  

-­‐   Prey  may  use  their  strength  or  “defensive  weapons”  (e.g.  some  prey  can  attack  
predators  using  horns,  spines,  gelatinous  sheaths,  etc.)  in  an  attempt  to  escape.  
-­‐   Prey  may  employ  group  defence  –  because  as  a  group  they  are  more  likely  to  be  able  
to  fend  off  predators.  

-­‐   Some  prey  show  adaptations  to  make  them  hard  to  handle  (manipulate  in  
preparation  for  consumption)  by  predators,  e.g.  spines,  gelatinous  sheaths,  etc.  

-­‐   Some  potential  prey  escape  simply  by  being  too  large  for  predators  to  handle  or  
consume,  e.g.  whales  and  elephants.  This  is  called  “escape  in  size  from  size  selective  
predation”.  

Consumption  stage  
-­‐   Predators  may  have  to  dismember  prey  (mechanical  preparation,  i.e.  convert  it  into  
smaller  pieces)  before  they  can  ingest  it  –  they  may  use  teeth,  jaws  and/or  claws.  

-­‐   Since  prey  may  be  toxic,  the  predator  may  have  to  be  able  to  detoxify  the  prey,  e.g.  
some  pythons,  such  as  the  yellow  python,  can  eat  rattle  snakes  because  they  can  
detoxify  rattle  snake  venom.  

-­‐   The  digestive  tract  of  the  predator  must  be  able  to  chemically  digest  (hydrolyse)  the  
food  prior  to  absorption.  

-­‐   Even  when  captured,  prey  will  show  the  last  line  of  defence  against  being  eaten.  This  
can  be  in  the  form  of  armour  or  spines  or  bad  taste  (including  toxics  or  emetics).  
Predators  may  learn  from  bad  experience  with  one  individual  not  to  prey  on  similar  
items  in  future.  

-­‐   But  how  does  this  help  the  dead  organism  (i.e.  how  can  such  a  trait  be  selected  for  if  
the  organism  has  to  die  for  the  defence  of  other  individuals?)  –  by  kin  selection.  This  
could  feedback  on  the  predator’s  selection  of  prey  to  attack.  

   
Lecture  11:  PREDATION  (CONTINUED)  
Module:  Principles  of  Ecology  
Dept.  of  Applied  Biosciences  and  Biotechnology,  Midlands  State  University  

11.1.   Effects  of  Predation  on  Prey  Populations  


11.2.   Regulation  of  Prey  Populations  by  Predation  
11.3.   Effects  of  Predation  on  Predator  Populations  
11.4.   Numerical  Response  
11.5.   Functional  Response  

___________________________________________________________________________  

   
EFFECTS  OF  PREDATION  ON  PREY  POPULATIONS  
We  can  ask  two  questions  on  the  effects  of  predation  on  prey  populations:    

(i) can  predators  regulate  prey  populations?  


(ii) Why  don’t  predators  typically  drive  prey  populations  to  extinction?  

Regulation  of  Prey  Populations  by  Predation  


A  number  of  field  studies  have  indicated  that  predator  populations  can  regulate  (limit)  prey  
species  abundance:  

Huffaker  and  Kenett  (1956):  


-­‐   Their  work  involved  two  species  of  mites  found  in  association  with  strawberry  plants.  
The  Cyclamen  mite  Tasonemus  pallidus  is  herbivorous  on  strawberries.  The  other  
species  Typhlodramus  occidentalis  is  carnivorous  and  predatory  upon  the  Tasonemus  
pallidus.  

-­‐   In  the  presence  of  its  predator  (Typhlodramus),  the  cyclamen  mite  maintained  low  
population  levels,  but  in  the  absence  of  its  predator  there  was  a  25-­‐fold  increase  in  
its  population  size  (treatment  with  a  chemical  insecticide  killed  the  “predator”  but  
not  the  “prey”).  

-­‐   This  was  a  removal  experiment  and  it  illustrated  that  predators  normally  keep  prey  
densities  low.  

Cactus  “prickly  pear”  (Opuntia  sp.)  and  the  cactus  moth  (Cactoblastis  
cactorum)  

-­‐   Caterpillars  of  the  cactus  moth  (Cactoblastis  cactorum  prey  upon  the  cactus  (and  
also  carry  pathogens  and  rot  causing  organisms).  

-­‐   The  cactus  was  introduced  in  Australia  from  South  America.  It  spread  rapidly  and  
dramatically  increased  in  abundance,  i.e.  the  cactus  expanded  its  distribution  and  
abundance  in  the  absence  of  its  natural  enemy  (or  predator).  

-­‐   The  introduction  of  the  cactus’  natural  enemy,  the  cactus  moth  (from  South  
America)  to  control  the  cactus  saw  a  dramatic  decline  in  abundance  and  distribution  
and  restrictions  in  the  range  of  the  cactus.  

-­‐   This  an  example  of  an  addition  experiment  which  shows  that  predators  can  regulate  
or  limit  prey  populations.  
-­‐   The  cactus  now  persists  as  a  “fugitive”  species  as  it  can  disperse  more  quickly  than  
the  moths  (some  seeds  reach  areas  where  the  moth  is  not  established,  and  can  grow  
and  reproduce  before  the  moth  catches  up  with  them).    

Why  don’t  predators  typically  drive  prey  populations  to  extinction?  


As  a  prey  species  becomes  rarer,  it  becomes  harder  for  predators  to  find  individuals  to  
consume.  The  predator  will,  therefore,  spend  a  longer  time  searching  for  that  prey  type.  
This  means  that  the  “costs”  of  predation  will  go  up  and  the  “returns”  will  go  down.  Some  
predators  will  then  switch  their  search  image  to  a  more  abundant  (and  hence  more  
profitable)  prey  type.  This  prey  switching  by  the  predator  reduces  the  predation  pressure  on  
the  rare  prey  species.  The  rare  prey  species  is  thus,  not  driven  into  extinction.  

On  the  effect  of  predation  on  prey  populations,  we  need  to  note:  Prey  populations  reduced  
by  their  predators  will  experience  a  compensatory  decline  in  the  depressant  effects  of  
intraspecific  competition,  while  those  that  grow  large  through  the  rarity  of  predators  will  
suffer  the  consequences  of  intraspecific  competition  all  the  more  intensely.  

Effects  of  Predation  on  Predator  Populations  


Since  prey  are  an  obvious  resource  for  their  predators,  we  can  ask:    

-­‐ are  predators  limited  by  prey  density?    

Or    

-­‐ how  do  predator  populations  respond  to  prey  density?  

Holling  (1959)  studied  the  predation  of  small  mammals  on  pine  sawflies  and  he  found  that  
predation  rates  increased  with  increasing  prey  density.  This  resulted  from  two  effects:  (1)  
each  predator  increased  its  consumption  rate  when  exposed  to  a  higher  prey  density,  and  
(2)  predator  density  increased  with  increasing  prey  density.  Holling  considered  these  effects  
as  two  kinds  of  responses  of  predator  population  to  prey  density,  with  the  increase  in  
consumption  rate  being  a  functional  response  and  an  increase  in  predator  density  being  a  
numerical  response.  

Numerical  response  
Numerical  response  means  that  predators  become  more  abundant  as  prey  density  
increases.  This  may  result  from  two  different  mechanisms:    

(i) increased  rate  of  predator  reproduction  when  prey  are  abundant  (numerical  
response  per  se).  
(ii) Attraction  of  predator  to  prey  aggregations  (“aggregational  response”).  

The  reproduction  rate  of  predators  naturally  depends  on  their  predation  rate.  The  more  
prey  consumed,  the  more  the  energy  the  predator  can  allocate  for  reproduction.  Mortality  
rate  also  decreases  with  increased  prey  consumption.  

 The  most  simple  model  of  predators’  numerical  response  is  based  on  the  assumption  that  
the  reproduction  rate  of  predators  is  proportional  to  the  number  of  prey  consumed.  This  is  
like  the  conversion  of  prey  into  new  predators.  Lets  say,  for  example,  as  10  prey  are  
consumed,  a  new  predator  is  born.  

We  can  define  the  following:  

-­‐   Threshold  –  same  as  threshold  level  or  light  compensation  point  –  is  the  minimum  
density  needed  to  sustain  a  predator  population.  

-­‐   Limited  region  –  increases  in  prey  density  over  this  range  cause  increases  in  predator  
fitness.  

-­‐   Saturated  region  (“unlimited  region”)  –  at  prey  densities  above  the  “saturation  
point”,  further  increases  in  prey  density  do  not  further  increase  predator  fitness.  

-­‐   Satiation  –  predators  can  only  consume  prey  at  a  limited  rate  (in  the  zone  
“saturation”  above).  

Satiation  leads  to  the  prey  strategy  of  explosive  reproduction,  in  which  the  prey  try  to  
swamp  predators.  

Food  quality  

Another  factor  complicating  the  relationship  between  predation  rate  and  predator  fitness  is  
food  quality.  It  may  take  more  of  a  lower  quality  prey  to  support  the  same-­‐sized  predator  
population.  In  particular,  herbivores  are  greatly  affected  by  the  quality  of  their  food,  
especially  nitrogen  content.  

Sinclair  (1975)  monitored  the  protein  content  of  the  food  available  to  the  Wildebeest  in  
Serengeti  during  1971.    He  also  monitored  the  fat  reserves  in  the  bone  marrow  of  the  live  
males,  and  of  males  that  had  died  from  natural  causes  (these  reserves  being  the  last  to  be  
utilised).  

It  can  be  concluded  from  Sinclair’s  work  that  the  wildebeest  consumed  food  that  was  below  
the  level  necessary  even  for  maintenance  (5  –  6%crude  protein)  in  the  dry  season  and  
judging  by  the  depleted  fat  reserves  of  the  dead  Wildebeest,  poor  food  quality  was  an  
important  cause  of  mortality.    
McNeil  (1978)  also  showed  that  seasonal  peaks  in  the  densities  of  insects  feeding  on  the  
grass  Holcus  mollis  are  related  to  peaks  in  food  quality  (measured  as  soluble  nitrogen)  in  the  
leaves  and  stems.  

Aggregational  Response  
Aggregation  of  predators  to  prey  density  is  often  called  “aggregational  response”.  
Aggregational  response  is  very  important  for  several  predator-­‐prey  systems.  Predators  
selected  for  biological  control  of  insect  pests  should  have  a  strong  aggregational  response,  
otherwise  they  would  not  be  able  to  suppress  prey  populations.  Also  aggregational  response  
increases  the  stability  of  spatially-­‐distributed  predator  –  prey  systems.  

Functional  Response  
The  density  of  prey  is  of  crucial  importance  to  the  predator.  As  prey  density  increases,  in  
general  the  consumer  eats  more  prey  (because  the  prey  becomes  easier  to  find,  etc.).  

The  relationship  between  an  individual’s  consumption  rate  and  food  density  is  known  as  its  
functional  response.  In  other  words,  functional  response  is  the  way  in  which  the  predation  
rate  of  predators  is  influences  by  prey  availability.  

Holling  (1959)’s  model  of  functional  response  illustrates  the  principle  of  time  budget  
behavioural  ecology.  It  assumes  that  the  predator  spends  its  time  on  two  kinds  of  activities:  

(i) Searching  for  prey  


(ii) Prey  handling  (which  includes  chasing,  killing,  eating  and  digestion).  

Consumption  of  the  predator  is  limited  in  Holling’s  model  because  even  if  prey  are  so  
abundant  that  no  time  is  needed  for  search,  a  predator  still  needs  to  spend  time  on  prey  
handling.  

Holling  (1959  identified  three  basic  types  of  functional  responses:  

“Type  2”  functional  response  


-­‐ Most  frequently  observed  and  most  typical  
 
Fig.11.1  Type  2  functional  response  

-­‐   In  this  type  of  functional  response  the  response  is  curvilinear.  Consumption  rate  
initially  rises  with  increasing  prey  density  –  the  rate  of  increase  then  slows,  and  
finally,  a  plateau  is  reached  at  maximum  ingestion  rate.  Holling  attributed  the  form  
taken  by  the  curve  to  the  existence  of  the  predator’s  handling  time  (the  time  the  
predator  spends  persuing,  subduing  and  consuming  each  prey  item  it  finds,  and  then  
preparing  itself  for  further  search).  

-­‐   Examples:  dragonfly  nymphs  eating  waterfleas  (Daphnia),  slugs  eating  grass,  etc.  

Explanation:  
-­‐   A  predator  has  to  devote  a  certain  amount  of  handling  time  to  each  prey  item  it  
consumes,  during  this  time  it  cannot  search  for  additional  items.  

-­‐   As  prey  density  increases,  finding  prey  becomes  increasingly  easy  (search  time  
becomes  trivial),  but  handling  individual  prey  still  takes  the  same  amount  of  time,  
and  handling  takes  up  an  increasing  proportion  of  the  predator’s  time.  

-­‐   Thus,  at  high  prey  densities,  the  predator  effectively  spends  all  its  time  handling  
prey,  and  the  predation  rate  reaches  a  maximum  determined  by  the  maximum  
number  of  handling  times  that  can  be  fitted  into  the  total  time  available.  Therefore,  
even  when  prey  is  superabundant  and  search  time  is  nil,  predation  rate  is  limited  by  
the  finite  handling  time  →plateau.  

Implications  of  “type  2”  functional  response  

-­‐ The  plateau  represents  predator  saturation.  Prey  mortality  declines  with  prey  
density.  Predators  of  this  type  cause  maximum  prey  mortality  at  low  prey  density.  
For  example,  small  mammals  destroy  most  of  the  gypsy  moth  pupae  in  sparse  
populations  of  the  gypsy  moth.  However,  in  high  density  defoliating  populations,  
small  mammals  kill  a  negligible  proportion  of  the  pupae.  

“Type  1”  functional  response  


-­‐ much  less  common  

 
Fig.  11.2  Type  1  functional  response  

In  this  type,  there  is  a  linear  increase  in  predation  rate  with  increasing  prey  density  until  the  
maximum  predation  rate  is  reached,  then  abrupt  plateau  at  the  maximum.  

The  type  1  functional  response  is  found  in  passive  predators  like  spiders  and  filter  feeders.  

Examples:  

-­‐   Spiders  –  the  number  of  flies  caught  in  the  net  (web)  is  proportional  to  fly  density.  

-­‐   Daphnia  magma  (waterflea)  filter-­‐feeding  on  yeast  cells  (Saccharomyces  cerevisiae)  
–  work  of  Rigler  (1961)  

-­‐   Rigler  (1961)  studied  the  feeding  rate  of  Daphnia  magma  with  the  yeast  
Saccharomyces  cerevisiae  as  its  prey.  Daphnia  magma  is  a  filter-­‐feeder  and  it  can  
extract  yeast  cells  from  a  constant  volume  of  water  washed  over  its  filtering  
apparatus.  

-­‐   At  low  prey  density  (below  105  yeast  cells  ml-­‐1)  the  predation  rate  is  directly  
proportional  to  food  concentration  because  the  Daphnia  can  filter  and  swallow  yeast  
cells  simultaneously.  

-­‐   Above  105  yeast  cells  ml-­‐1,  however,  Daphnia  are  unable  to  swallow  (i.e.  handle)  all  
the  food  they  filter.  At  such  concentrations,  therefore,  they  ingest  food  at  a  maximal  
rate,  limited  by  their  handling  time.  

 
 

Fig.  11.3.   Functional  response  of  D.  magma  to  different  concentrations  
of  yeast  cells    

Explanation  
-­‐   Passive  predators  (e.g.  spiders  which  use  webs,  and  filter  feeders)  can  search  for  and  
handle  prey  simultaneously.  

-­‐   Below  the  plateau,  search  time  (filtering  rate  in  this  case)  is  limiting.  Daphnia,  for  
example,  filter  water  at  the  maximal  rate  –  ingestion  rate  increases  with  increasing  
prey  density  (as  all  that  are  found  are  captured  and  consumed  rapidly).  

-­‐   When  the  plateau  is  reached,  ingestion  rate  is  at  its  maximum  and  “suddenly”  this  is  
the  limiting  factor.  

“Type  3”  Functional  Response  


-­‐ In  this  type,  the  response  of  predation  rate  to  prey  density  is  sigmoidal.  The  
predation  rate  initial  increases  exponentially  with  increasing  prey  density,  but  with  
further  increases  in  prey  density,  ingestion  rate  declines  until  it  plateaus  at  a  
maximum  (i.e.  slope  increases,  then  decreases)(Fig  11.4).  

 
Fig  11.4.  Type  3  functional  response  

-­‐ The  type  3  functional  response  occurs  in  predators  which  increase  their  search  
activity.  In  other  words,  type  3  functional  response  occurs  whenever  attack  rake  
increases,  or  handling  time  decreases,  with  increasing  prey  density.  

Exapmles:  

-­‐   Blue  bottle  flies  feeding  on  sugar  droplets,  

-­‐   Shrews  feeding  on  pine  sawflies,  

-­‐   Ichneumonid  wasps  parasitizing  their  insect  hosts.  

Explanation:  
-­‐   Seen  whenever  an  increase  in  prey  density  increases  the  predator’s  searching  
efficiency  and/or  decreases  its  handling  time.  Either  or  both  will  increase  ingestion  
rate  “faster”  than  linear.  

-­‐   These  generally  involves  learning  on  the  part  of  the  predator-­‐i.e.  with  practice  the  
predator  will  become  better  able  to  find  and  handle  prey.    

-­‐   Also,  type  3  functional  responses  result  from  switching  preferences.  When  the  
predator  switches  to  alternative  prey,  the  mortality  of  the  initial  prey  declines.  
Polyphagus  vertebrate  predators  (e.g.  birds)  can  switch  to  the  most  abundant  prey  
species  by  learning  o  recognise  it  visually.  Mortality  first  increases  with  increasing  
prey  density,  and  then  declines.  

Functional  responses  and  the  regulation  of  Prey  densities  


-­‐ Can  predators  regulate  the  population  sizes  of  their  prey?  

Or  

-­‐   Are  prey  population  dynamics  stable?  

-­‐   In  order  to  have  regulation,  the  mortality  of  the  prey  must  increase  with  increasing  
density  N,  i.e.  only  positive  density  dependent  mortality  can  regulate  the  dynamics  
of  prey  populations.  

-­‐   If  predator  density  is  constant  (e.g.  birds,  small  mammals),  then  they  can  prey  
density  only  if  they  have  a  type  3  functional  response  because  this  is  the  only  type  of  
functional  response  for  which  mortality  can  increase  with  increasing  prey  density.  
-­‐   However,  the  regulating  effects  of  predators  are  limited  to  the  initial  portion  of  the  
type  3  functional  response  where  there  is  a  positive  slope  in  the  prey  mortality-­‐vs-­‐
prey  density  curve,  i.e.  the  interval  of  the  prey  density  where  mortality  increases.  
Increases  in  prey  density  at  this  interval  lead  to  increases  in  predation  pressure  on  
the  prey  population.  This  process  is  density  dependent  and  has  a  potentially  
important  stabilising  effect  on  the  interaction.  If  prey  density  exceeds  the  upper  limit  
of  this  interval,  then  mortality  due  to  predation  starts  declining,  and  predation  will  
cause  a  positive  feedback.  As  a  result  the  number  of  prey  will  get  out  of  control.  
They  will  increase  in  numbers  until  some  other  factors  (intraspecific  competition,  for  
example)  will  stop  their  reproduction.  This  phenomenon  is  known  as  “escape  from  
natural  enemies”  and  was  first  discovered  by  Takahashi.  

-­‐   Wildlife  managers  are  familiar  with  type  3  functional  responses:  e.g.  in  the  presence  
of  wolves  ,  whitetail  deer  densities  are  relatively  stable.  When  wolves  are  
exterminated,  whitetail  population  densities  may  fluctuate  wildly.  This  can  lead  to  
the  collapse  of  the  population  if  it  seriously  overgrazes  its  range.  

-­‐   Type  3  functional  responses,  which  are  capable  of  stabilizing  prey  population  
densities,  have  traditionally  been  referred  to  as  “vertebrate  functional  responses”.  It  
was  believed  that  because  they  involve  learning  only  vertebrates  would  exhibit  
them.  However,  insects  can  also  learn  hence  blue  bottle  flies  and  ichneumonid  
wasps  also  exhibit  type  3  functional  responses.  

-­‐   Type  2  functional  responses  have  traditionally  been  referred  to  as  “invertebrate  
functional  responses”.  

-­‐   Hunting  can  be  viewed  as  another  “type  3  functional  response”  because  you  can  
become  a  better  hunter  with  experience  –  more  efficient  at  locating  and  capturing  
prey.  

   
Lecture  12:  PREDATION  
Module:  Principles  of  Ecology  
Dept.  of  Applied  Biosciences  and  Biotechnology,  Midlands  State  University  

12.1.   Predator  Behaviour  


12.2.   Diet  Composition  -­‐  Preferences  
12.3.   Diet  Width  –  The  Optimum  Foraging  Theory  
12.4.   Foraging  Behaviour  and  Predation  Risk  

___________________________________________________________________________  

 
 
 
 
 
 

 
 
 
 

 
 
 
 
PREDATOR  BEHAVIOUR  
-­‐   Organisms  face  a  multitude  of  problems  through  their  lives,  but  dominated  by  three  
things:  eating,  avoiding  being  eaten  and  reproducing.  Those  organisms  that  do  any  of  
these  better  than  their  fellows  will,  on  average,  leave  more  progeny.  If  the  trait  
which  makes  them  better  has  a  genetic  component,  it  will  become  more  common,  
i.e.  it  will  be  “selected”  by  “natural  selection”.  So,  over  time  we  would  expect  that  
organisms  become  as  good  as  possible  at  doing  the  jobs  they  have  evolved  to  do  
(finding  food,  mates,  and  avoiding  predators).  

-­‐   Using  simple  arguments  based  on  some  basic  biological  assumptions  and  the  
premise  that  natural  selection  happens  we  can  make  predictions  about  hoe  
organisms  should  behave.  For  the  process  of  food  gathering  and  eating  (collectively  
known  as  “foraging”)  we  can  predict  that  organisms  will  have  been  selected  to  
maximise  the  profitability  of  their  feeding  (energy  per  unit  time)  whilst  minimising  
the  risks  of  being  eaten.  

-­‐   The  field  of  behavioural  ecology  is  concerned  with  looking  at  the  behaviour  of  
organisms  and  seeing  if  their  behaviour  makes  “sense”  in  the  light  of  evolutionary  
biology.  By  looking  at  foraging  behaviour,  we  are  also  trying  to  understand  how  
natural  selection  has  favoured  particular  patterns  of  behaviour  in  particular  
circumstances.  

-­‐   Predator  behaviour:  

-­‐   is  a  product  of  evolution  (natural  selection),    

-­‐   coevolves  with  prey  defences  

-­‐   affects  predator-­‐prey  population  dynamics.  

Diet  Composition  –  Preferences  


-­‐   We  can  start  by  asking  the  question:  

-­‐   What  determines  the  diet  composition  for  a  particular  predator?  

-­‐   Predators  are  classified  as  either  monophagus  (feeding  on  a  single  prey  type),  
oligophagus  (feeding  on  few  prey  types)  or  polyphagus  (feeding  on  many  prey  
types).    

-­‐   An  equally  important  classification  is  between  specialists  (broadly,  monophagus  and  
ologophagus)  and  generalists  (polyphagus).    

-­‐   It  should  be  noted  that  polyphagus  and  oligophagus  predators  are  not  indiscriminate  
in  what  they  choose  from  their  acceptable  range.  Many  predators  are  selective  –  
they  show  preferences  for  some  prey  items,  i.e.  they  eat  a  prey  out  of  proportion  of  
its  abundance.  -­‐Thus,  to  measure  food  preference  in  nature,  it  is  necessary  not  only  
to  examine  the  animal’s  diet  (usually  by  analysis  of  gut  contents)  but  also  to  access  
the  availability  of  different  food  items.  

-­‐   Examples:  
1.   Bluegills  eating  Daphnia  of  different  sizes  (the  work  of  Werner):  

PREY  SIZE   %  IN  THE  ENVIRONMENT   %  IN  THE  GUT  


1.4mm   25   10  
1.9mm   25   20  
2.5mm   25   31  
3.6mm   25   39  
 

This  is  an  example  of  a  RANKED  PREFERENCE  –  predators  prefer  prey  with  
the  highest  energy  content  (but  these  also  require  greater  handling).  Prey  can  
be  ranked  on  a  simple  scale:  

Benefit/cost  =  Energy  gain/handling  time  


Attacking,  capturing,  and  ingesting  prey  cost  time  and  energy.  

2.   BALANCED  PREFERENCE  –  based  on  the  need  for  a  balanced  diet  of  
complementary  prey  items:  

-­‐   Fitness  higher  with  a  mixed  diet,  

-­‐   Complementary  resources,  or  avoid  eating  too  much  of  any  one  
plant/animal  toxin.  

3.   SWITCHING  PREFERENCE  –  predators  tend  to  prefer  the  most  common  prey.  
If  the  most  common  prey  changes  (maybe  due  the  action  of  the  predator),  
the  predator  switches  to  the  “new”  most  common  prey  type:  

When  prey  is  abundant,  the  predator  “prefers”  it  and  consumes  it  in  greater  
proportion  than  its  availability.  

When  this  prey  becomes  scarce,  the  predator  switches  to  a  more  abundant  
prey  item  and  consumes  it  in  lesser  proportion  than  its  availability.  

Predator  switching  is  based  upon  a  learnt  ability  to  specialise.  Certain  
predators  especially  vertebrates,  develop  specific  search  images  which  
enable  them  to  search  more  effectively  (since  they  know  what  they  are  
looking  for)  and  result  in  them  concentrating  on  their  “image”  prey  to  the  
relative  exclusion  of  their  “non-­‐image”  prey.  

Predators  exhibit  RANKED  PREFERENCES  in  discriminating  between  resource  


types  that  are  perfectly  “substitutable”  and  exhibit  BALANCED  PREFERNCES  
between  resource  types  that  are  complementary.  

Ranked  preferences  predominate  when  food  items,  can  be  classified  on  a  
single  scale.  Many  predators  exhibit  a  combination  of  ranked  and  balanced  
preferences.  They  select  food  that  is  generally  of  high  quality,  but  they  also  
select  items  to  meet  specific  requirements.  

Diet  Width  -­‐  Optimal  Foraging  theory  


-­‐   Knowledge  of  the  diet  width  among  predators  and  of  the  ways  in  which  predators  
distribute  their  effort  amongst  prey  items  will  give  insights  to  the  ways  in  which  
predators  tend  to  maximise  their  profits.  

-­‐   Natural  selection  has  shaped  foraging  behaviour  and  all  the  foraging  behaviours  tend  
to  converge  on  the  optimum  (=the  best  possible)  –  “the  ghost  of  predation  past”.  

-­‐   The  optimal  foraging  theory  allows  us  to  predict  the  optimum  foraging  behaviour  for  
a  predator  (e.g.,  actively  searching  for  prey,  or  ambush  predation).  In  other  words,  
the  aim  of  the  optimal  foraging  theory  is  to  predict  the  foraging  strategy  to  be  
expected  under  specified  conditions.  We  can  compare  data  from  the  field  to  the  
theory’s  predictions  and  judge  whether  the  predator  is  truly  optimal.  

Model  of  Diet  Width  


-­‐   Once  a  prey  is  encountered,  should  it  be  eaten?  

-­‐   The  model  of  diet  width  addresses  which  prey  to  include  in  the  diet  –  preferences.  

-­‐   We  rank  prey  items  by  profitability.  

-­‐   Then  ask  if  the  predator  should  expand  its  diet  by  including  the  next,  most  profitable  
prey  item?  

-­‐   We  can  divide  the  time  spend  by  the  predator  (with  respect  to  foraging)  into  two  
categories:  

1.   Search  time  –  the  time  spent  trying  to  locate  prey  items  (S  =  mean  time  spent  
searching  for  food).  
2.   Handling  time  (hi  for  the  ith  item;  h  =mean  handling  time  for  all  items)  –  this  
is  the  time  spent  between  the  location  and  the  final  consumption  for  all  prey  
items,  and  includes:  

(a)   time  spent  pursuing  the  item  until  capture,  

(b)   time  spent  manipulating  and  consuming  the  item  after  capture.  

-­‐   If  a  predator  is  a  SPECIALIST,  it  will  pursue  and  consume  only  profitable  prey  items,  
but  it  may  expend  a  great  deal  of  time  and  energy  searching  for  them.  

-­‐   If  a  predator  is  a  GENERALIST,  then  it  will  spend  relatively  little  time  searching  but  it  
will  pursue  and  consume  both  profitable  and  unprofitable  prey  items.  

-­‐   An  OPTIMAL  FORAGER  “should  balance  the  pros  and  cons”  of  whether  or  not  to  
include  a  given  item  so  as  to  maximize  its  overall  rate  of  energy  intake.  

Calculating  profitability:  

For  the  ith  species:  

Profitability   =   Energy  Content  of  Prey/handling  time          =   Ei/hi  

Profitability   =     E/h   =    Mean  Energy  Content/Mean  handling  time  

If  a  predator  eats  prey  type  i  its  rate  of  energy  gain  is:  

Ei/hi  

If  a  predator  ignores  prey  type  i,  it  must  search  for  time  S  to  find  another  prey  item  
during  which  its  rate  of  energy  gain  is:    

E/(S  +  h)  

We  predict  that  the  predator  will  include  the  item  i  in  its  diet  only  if:    

Energy  Gain  if  taken     ≥   Energy  Gain  if  ignored  


or  

Ei/hi   ≥   E/(S  +  h)  

So,  a  predator  should  add  item  i  to  its  diet  if  it  increases  its  average  rate  of  energy  
gain  by  doing  so.  

Predictions  from  the  Optimal  Foraging  Theory  


Some  predictions  of  the  Optimal  Foraging  Theory  include:  

1.   Predators  with  handling  times  that  are  typically  short  compared  to  their  search  times  
should  be  generalists.  

-­‐   hi  small  →  Ei/hi  is  large  

-­‐   while  hi  small,  S  is  large,  and    Ei/hi  remains  unchanged  

then,  for  many  prey  items,  Ei/hi  will  be  greater  than  E/(S  +  h),  so  they  will  be  
included  in  the  diet,  e.g.,  birds  that  “glean”  stationary  insects  from  foliage  are  
generalists.    

2.   Predators  with  handling  times  that  are  long  compared  to  their  search  times  should  
be  specialists.  

-­‐   S  is  small,  so  E/(S  +  h),  is  effectively  =  E/h    

-­‐   Therefore,  maximizing  E/(S  +  h),  is  effectively  equal  to  maximizing    E/h    
which  is  done  simply  by  including  the  most  profitable  prey  items  (i.e.  those  
with  the  largest  Ei/hi  )  in  the  diet,  lions  spend  much  time  pursuing,  
capturing  and  consuming  their  prey.  As  a  result  they  select  the  old,  lame  and  
weak  which  are  easier  to  catch  (hi  small  →  Ei/hi  large).    

3.   In  unproductive  environments  (i.e.,  where  prey  are  scarce  and  hence  S  is  large),  
predators  should  have  broader  diets  than  in  productive  environments  (where  S  is  
small):  

e.g.,  this  is  observed  in  both  bluegill  sunfish  and  the  Great  tit.  Within  each  species,  
diet  width  depends  on  productivity.  

4.   Whether  or  not  a  predator  should  include  an  ith  item  in  its  diet  should  depend  on:  

(a)   The  profitability  of  that  item  (Ei/hi)  

(b)   The  profitability  of  items  already  in  the  diet  (E/h)  

(c)   The  search  times  for  items  already  in  the  diet.  

…………………………….but  it  should  not  depend  on:  

the  search  time  of  the  ith  item  (Si)  because  it  is  being  encountered  
during  the  “normal”  search  (S),  nor  therefore  should  it  depend  on  i’s  
abundance  or  the  rate  at  which  the  predator  encounters  it.  
………………………………therefore  

(a)   Predators  should  specialise  when  profitable  prey  items  are  common  
and/or  differences  in  profitability  are  great.  

(b)   Predators  should  not  indiscriminate  when  profitable  prey  items  are  
rare  and/or  differences  in  profitability  are  slight.  

(c)   Predators  should  ignore  insufficiently  profitable  prey  types    


    irrespective  of  their  abundance.  

What  are  the  constraints  on  reaching  optimality?  

(i) Other  aspects  of  the  predator’s  behaviour  may  influence  fitness  more  
than  foraging  behaviour,  such  as  avoidance  of  the  predator’s  
predators.  
(ii) For  many  predators  (especially  herbivores  and  omnivores),  energy  
intake  may  not  be  the  crucial  factor.  Rather,  it  may  be  some  other  
factor  (e.g.,  nitrogen  content)  o  obtaining  a  balanced  diet  by  
consuming  a  mixture  of  prey.  

(iii)   While  foraging  behaviour  evolved  under  one  set  of  circumstances  
(e.g.,  to  limit  predator  or  to  supply  a  balanced  diet),  present  emphasis  
may  be  on  energy  –  there  may  have  been  not  enough  time  to  evolve  
the  optimal  strategy  (phylogenetic  constraint).  

However,  if  energy  intake  rate  is  the  crucial  factor  and  other  aspects  like  
predation  risk  are  taken  into  account,  the  optimal  foraging  theory  is  a  
powerful  tool  –  lands  insights  into  the  foraging  decisions  a  predator  must  
make.  

Foraging  Behaviour  and  Predation  Risk  


-­‐   Organisms  face  a  multitude  of  problems  in  their  lives,  chief  among  them  eating,  
reproducing  and  avoiding  being  eaten.  Foraging  strategies  will  not  always  be  
strategies  for  simply  maximizing  feeding  efficiency.  Natural  selection  actually  favours  
foragers  that  maximize  the  profitability  of  their  feeding  whilst  minimizing  the  risks  of  
being  eaten.  

-­‐   The  modifying  influence  of  predators  on  foraging  behaviour  was  studied  by  Werner  
et  al  (1983)  working  on  bluegill  sunfish.  

-­‐     Here  we  have  three  trophic  levels:  

Prey     →   our  predator    →   our  predator’s  predator    


Our  predator  is  the  forager  and  our  predator’s  predator  is  the  predator  

The  forager  must  balance  eating  –  vs  –  being  eaten  →  conflicting  demands  

Example  –  Bluegills  (work  of  Werner  et  al)  

Predator  =  largemouth  bass  (fish)  

Forager  =  bluegills  

Prey  =  aquatic  insects,  zooplankton,  etc  

In  the  absence  of  the  predators,  the  bluegills  spent  much  of  their  time  for  aging  in  
“open  water”  (few  weeds)  where  food  was  easy  to  spot  and  capture.  

When  predators  (largemouth  bass)  were  present,  the  bluegills  shifted  to  
increasingly  forage  in  more  heavily  vegetated  areas  where  there  is  more  cover  to  
avoid  detection  by  the  bass  and/or  to  facilitate  escape.  

In  the  absence  of  the  predators,  the  bluegills  move  amongst  the  available  sites  to  
maximize  their  energy  intake  →  makes  optimal  foraging  predictions.  

In  the  presence  of  predators,  the  bluegills  spent  much  more  time  in  prey  poor  
“refuges”  –  energy  intake  much  less  (decreased  growth  and  reproduction),  but  
overall  survival  is  increased.  

When  we  looked  at  the  niche  concept  and  interspecific  competition  we  saw  that  
realized  niches  can  be  highly  constrained  by  the  effects  of  competitors.  The  bluegills  
example  also  illustrates  that  realized  niches  are  also  constrained  by  predators.  

   
Lecture  13:  PREDATOR  BEHAVIOUR  
Module:  Principles  of  Ecology  
Dept.  of  Applied  Biosciences  and  Biotechnology,  Midlands  State  University  

13.1.   Lotka-­‐Volterra  Predator-­‐Prey  Model  


13.2.   Do  Predator-­‐Prey  Cycles  occur  in  nature?  
13.3.   Conclusions  about  Lotka-­‐Volterra  Predator-­‐Prey  Model  

___________________________________________________________________________  

   
THE  POPULATION  DYNAMICS  OF  PREDATION  
We  can  look  at  the  effects  of  predation  on  the  population  dynamics  (i.e.,  changes  in  density  
over  time)  of  both  the  predator  and  its  prey.  

We  will  begin  by  considering  a  relatively  simple  model  based  on  differential  equations  
considering  just  one  predator  and  one  prey  species.  

Lotka-­‐Volterra  Predator-­‐Prey  Model  


The  Lotka-­‐Volterra  model  is  the  simplest  model  of  predator-­‐prey  interactions.  The  model  
was  developed  independently  by  Lotka  in  1925  and  Volterra  in  1926.  

The  model  has  two  important  variables:  

C  =  number  (or  density)  of  the  predator  population  

N  =  number  (or  density  or  biomass)  of  the  prey  or  plant  population  

Here  we  have  two  simultaneous  differential  equations,  one  for  the  prey  and  the  other  one  
for  the  predator.  

1)Prey  Population  growth  rate:  

Prey  Population  growth  rate  =  exponential  birth  rate  –  predator  caused  death  rate  

Assumption  #1:     in  the  absence  of  predators,  the  prey  population  increases  
exponentially,  i.e.,  no  density-­‐dependence:  

dN/dt  =rN  

However,  in  the  presence  of  predators  prey  individuals  are  removed  (i.e.  killed)  by  
predators.  The  rate  of  prey  removal  depends  on  the  rate  of  predator-­‐prey  encounters  and  
the  efficiency  of  the  predator.  

Encounters  increase  as  (1)  the  number  of  predators  (C)  increase  and,  (2)  the  number  of  Prey  
(N)  increases.  

The  predator’s  efficiency  can  be  designated  a’  the  “attack  rate”  of  the  predator.  

a’  increases  with  (1)  increasing  searching  efficiency  and  (2)  increasing  capture  efficiency.  

Assumption  #2:     a’  comes  from  a  linear  functional  response.  

 
The  rate  of  successful  encounters  (i.e.,  the  consumption  rate)  will  then  be:  a’CN    

and  the  overall  equation  for  the  prey  population  dynamics  becomes:    

dN/dt  =  rN  –  a’CN         equation  1.  

2)Predator  Population  Growth  Rate  


Predator  population  growth  rate  =  birth  rate  –  background  m ortality  

In  the  absence  of  prey  (i.e.,  food),  predator  individuals  lose  weight  and  starve  to  death.  
Therefore,  in  this  model,  predators  are  assumed  to  suffer  “background”  mortality.  Thus,  the  
predator  population  declines  exponentially  through  starvation  in  the  absence  of  prey.  

dC/dt  =  -­‐  qC  

where:    

q  =  mortality  rate  

This  loss  of  predators  is  counteracted  by  predator  birth  which  depends  on:  

(i) the  rate  at  which  food  (prey)  is  consumed:  a’CN  
(ii) the  predator’s  efficiency(f)  of  turning  this  food  into  predator  offspring.  

There  are  two  additional  assumptions  here:  

Assumption  #3:     there  is  no  mutual  interference  or  cooperation  between  predators.  

Assumption  #4:     the  conversion  of  prey  into  predator  offspring  has  a  linear  numerical  
response  –  no  thresholds  or  saturation:  

-­‐ The  predator  birth  rate  then  becomes:  

fa’CN  

and  the  overall  equation  for  the  predator  population  dynamics  becomes:  

dC/dt  =  fa’CN  –  qC     equation  2  

equations  (1)  and  (2)  above  together  constitute  the  Lotka-­‐Volterra  Predator-­‐Prey  Model.  

To  begin  to  investigate  the  properties  of  this  model,  we  start  with  a  plot  of  prey  abundance  
(x  –  axis)  vs  predator  abundance  (y  –  axis)  and  find  the  ZERO  NET  GROWTH  ISOCLINES  for  
each  species  (similar  to  what  we  did  in  the  Lotka  –  Volterra  Model  of  Interspecific  
competition).  
ZNGI’s  are  lines  along  which  the  population  just  maintains  itself,  neither  increasing  nor  
decreasing.  On  one  side  of  the  isocline,  the  population  will  increase,  on  the  other  sid,  it  will  
decrease.  

1)  Prey  Zero  Net  Growth  Isocline  –  this  is  when:  

dN/dt  =  0,       or     dN/dt  =  rN  –a’CN  =  0  

rearranging  gives:  

rN  =  a’CN  

or  

C  =r/a’  

Graphically:  

 
 

2)  Predator  Zero  Net  Growth  Isocline  –  this  is  when:  

dC/dt  =  0,       or     dC/dt  =  fa’CN  –  qC  =  0  

rearranging  gives:  

fa’CN  =  qC  

or  

N  =q/fa’  
Graphically:  

To  get  the  behaviour  of  both  species  (i.e.,  the  whole  system)  we  can  plot  both  isoclines  
together  and  use  vector  addition  as  before.  

We  end  up  with  a  number  of  cycles  =  coupled  oscillations:  These  coupled  oscillations  
continue  indefinitely  (frictionless  pendulum)  i.e.  the  cycle  does  not  degrade  over  time.  

For  a  single  predator  species  and  a  single  prey  species,  the  predator  tends  to  increase  in  
abundance  when  there  are  large  numbers  of  prey.  However,  there  should  then  be  an  
increase  in  predation  pressure  on  the  prey,  leading  to  a  decrease  in  prey  abundance.  This  
will  ultimately  lead  to  food  shortage  for  predators,  a  decrease  in  predator  abundance,  a  
concomitant  drop  in  predation  pressure,  an  increase  in  prey  abundance  and  so  on.  Thus  
predators  and  their  prey  undergo  coupled  oscillations  in  abundance,  with  predator  
numbers  “tracking”  those  of  the  prey.  

“Neutral  stability”  –  populations  follow  precisely  the  same  cycles  indefinitely  unless  some  
external  influence  (i.e.,  some  disturbance)  shifts  them  to  new  values,  after  which  they  
follow  the  new  cycles  indefinitely,  e.g.  

 
So,  if  a  given  cycle  is  disturbed,  there  is  no  tendency  to  return  to  the  original  cycle  –  (i.e.,  
there  is  no  regulation).  

The  Lotka  –  Volterra  predator  –  prey  model  is  useful  in  that  it  points  out  the  underlying  
tendency  for  these  interactions  to  generate  cycles.  The  fact  that  we  do  not  generally  see  
perfectly  regular  cycles  “in  nature”  could  then  be  explained  by  the  frequent  disturbances  
which  set  up  new  cycles  (due  to  changes  in  climate,  food  availability  to  prey,  etc.).    

The  cycling  results  from  a  series  of  time  lags:  

(i) Between  “many  prey”  and  “many  predators”  –  time  needed  to  produce  new  
predator  offspring  after  prey  abundance  increases.  
(ii) Between  “many  predators”  and  “few  prey”  –  time  needed  for  predator  to  
reduce  the  size  of  the  prey  population.  
(iii) Between  “few  prey”  and  “few  predators”  –  time  needed  for  the  predator  to  
starve  after  their  resource  has  declined.  
(iv) Between  “few  predators”  and  “many  prey”  –  time  needed  for  the  prey  
population  to  recover  following  reduction  in  predation  pressure.  

Do  Predator  –  Prey  Cycles  Occur  in  Nature?  


Snowshoe  hare  (Lepas  americanus)  and  the  Lynx  (Lynx  canadensis)  

The  snowshoe  hare  follows  a  roughly  10  year  cycle,  with  10  to  100  fold  changes  in  
abundance.  The  cycle  is  indirectly  driven  bt  predation  by  the  lynx.  

First  note  that  there  also  cycles  in  other  prey  species  taken  by  the  lynx,  such  as  a  bird  known  
as  the  ruffed  grouse.  

The  decline  in  hare  abundance  is  only  indirectly  due  to  predation  by  the  lynx,  but  more  
directly  due  to  overgrazing  of  winter  food  (woody  browse).  

When  the  hares  are  abundant,  they  overgraze  woody  browse,  leading  to  food  shortages,  
low  birth  rates,  low  juvenile  survivorship,  high  weight  loss  and  low  growth  rates  in  the  hares  
result.  

In  addition,  when  the  woody  plants  are  heavily  overgrazed,  they  produce  high  levels  of  
toxins  (takes  2  –  3  years  for  these  to  build  up).  This  results  in  a  decreased  quality  of  food.    

These  two  factors  (low  food  quality  and  quantity)  together  lead  to  a  poor  “condition”  in  the  
hares.  This  makes  them  more  susceptible  to  predation  by  predators.    

So,  the  cycles  in  the  prey  population  (hares)  are  really  generated  by  a  time-­‐lag  in  the  hare-­‐
plant  interactions  and  have  little  to  do  with  the  lynx.  
When  hare  numbers  decline,  woody  browse  can  increase  in  quantity  and  quality  –  woody  
browse  is  released  from  predation.  

Also,  the  cycles  in  the  predator  (lynx)  population  simply  track  the  cycles  in  their  prey  (the  
hares),  they  do  not  generate  them:  

When  the  hare  population  recovers,  the  predator  to  prey  ration  declines.  This  means  more  
food  for  the  lynx,  higher  lynx  birth  rate  and  therefore  an  increase  in  the  abundance  of  the  
lynx.  

When  the  hare  population  declines,  the  predator  to  prey  ration  increases.  This  means  less  
food  for  the  lynx,  lower  lynx  birth  rate  (and/or  death  rate  due  to  starvation),  and  therefore  a  
decline  in  numbers  of  the  lynx.  

 But  additionally,  when  hares  become  scarce,  the  lynx  may  switch  to  the  ruffed  grouse  and  
other  (alternate)  prey.  Thus,  the  lynx  can  drive  cycles  in  the  grouse  but  not  in  the  hare.  

Conclusions  about  the  Lotka  -­‐  Volterra  Predator  –  Prey  Model:  


-­‐   You  have  to  be  very  careful  when  applying  simplistic  models  to  (often  complex)  field  
situations.    

-­‐   Recall  that  the  Lotka  –  Volterra  Model  makes  some  very  basic  assumptions  that  are  
rarely  expected  to  be  met,  e.g.:  
(i)   Density  –  independent  population  growth  of  prey.  

(ii)   Simple  linear  functional  response.  

(iii)   Linear  numerical  response.  

-­‐   Other  models  have  been  developed  (such  as  the  “Theta  logistic”)  which  allow  logistic  
population  growth  of  prey,  type  1/type  2/type  3  functional  responses,  etc.  

-­‐   These  models  still  indicate  that  cycles  are  likely  under  most  conditions  (but  not  seen  
if  there  is  type  3  functional  response).    

-­‐   But  can  get  dumped  oscillations,  etc.  as  well.  

   
Lecture  14:  POPULATION  REGULATION  
Module:  Principles  of  Ecology  
Dept.  of  Applied  Biosciences  and  Biotechnology,  Midlands  State  University  
14.1   Density  dependence  and  Population  regulation  

___________________________________________________________________________  

   
POPULATION  REGULATION  
Recall  that  the  size  (or  density)  of  a  population  can  vary  temporarily  due  to  four  
demographic  processes:    

1.   Mortality  –  death  of  individuals  (decreases  population  size).  

2.   Natality  (or  fecundity)  –  birth  of  individuals  (increases  population  size).  

3.   Emigration  –  movement  of  individuals  out  of  a  population  (by  migration,  dispersal,  
etc.)  (decreases  population  size).  

4.   Immigration  –  movement  of  individuals  into  a  population  (by  migration,  dispersal,  


etc.)  (increases  population  size).  

Begon  and  Mortimer  (1986)  define  regulation  as  the  ability  to  decrease  the  size  of  
populations  which  are  above  a  particular  level  and  to  allow  an  increase  in  the  size  of  
populations  which  are  below  a  particular  level.  

This  particular  level  will  therefore  be  a  point  of  equilibrium.  Populations  below  it  increase,  
those  above  it  decrease,  and  populations  actually  on  it  neither  increase  nor  decrease:  
population  size  is  subject  to  negative  feedback.  In  the  case  of  the  effects  of  intraspecific  
competition,  this  equilibrium  level  is  called  the  carrying  capacity  of  the  population.  

In  other  words,  there  is  a  tendency  for  N  to  increase  when  it  is  low  and  to  decline  when  it  is  
high.  This  results  in  logistic  growth.  

Regulation  has  a  well-­‐defined  meaning  and  by  definition  can  only  occur  as  a  result  of  a  
density  dependent  process.  

The  regulation  of  most  populations  results  from  density  dependence  of  mortality  and/or  
fecundity.  There  are  three  possible  patterns:  

(i) both  birth  and  death  rates  are  density-­‐dependent.  


(ii) only  birth  rate  is  density-­‐dependent.  
(iii) only  death  rate  is  density-­‐dependent.  

Examples  of  density-­‐dependent  mortality  (i.e.,  mortality  which  increases  as  N  increases):  

-­‐   Daphnia,  in  the  Lab.  

-­‐   Daphnia,  in  the  field.    

-­‐   Flour  beetle,  in  the  Lab.  

-­‐   Buffalo,  in  the  field  (Serengeti).  


Examples  of  density-­‐dependent  fecundity  (i.e.,  fecundity  which  increases  as  N  increases):  

-­‐   Daphnia,  in  the  Lab.  

-­‐   Annual  plant,  in  the  field.    

-­‐   Sparrow,  in  the  field  (Serengeti).  

Unlike  regulation,  which  can  only  result  from  density  dependent  processes,  abundance  will  
be  determined  by  the  combined  effects  of  all  the  factors  and  all  the  processes  that  impinge  
on  a  population  (Begon  and  Mortimer  (1986).  

In  conclusion,  it  should  be  noted  that  some  density-­‐independent  factors  such  as  climate  can  
drive  density  dependent  processes,  for  example,  Andrewartha  and  Davidson  felt  that  
weather  acted  as  a  density-­‐dependent  component  of  the  environment  during  winter,  by  
killing  the  proportion  of  thrips  (Thrips  imaginis)  population  inhabiting  less  favourable  
situations  (if  the  number  of  safe  sites  is  limited  and  remains  roughly  the  same  from  year  to  
year,  then  the  number  of  individuals  outside  these  sites  killed  by  weather  will  increase  with  
density.  

   
POPULATION  REGULATION  Recall  that  the  size  (or  density)  of  a  population  can  vary  
temporarily  

Recall  that  the  size  (or  density)  of  a  population  can  vary  temporarily  due  to  four  
demographic  processes:  

Vous aimerez peut-être aussi