Download as pdf or txt
Download as pdf or txt
You are on page 1of 73

EC/075

~ l:l •
541.5
' • S7
T6
39 How Trees Grow Ph1hp R Morey
BOOKS IN THE SERIES 40 Endogenous Plant Growtli Substance:
2nd edn T A Hill
1 Ecological Energetics John Phillipson 41 Hair · Michael Ryder
2 Life in the Soil 42 The Structure and Function of Enzyme!
R M . Jackson and Frank Raw 2nd edn Cohn H Wynn
3 The Study of Behaviour 2nd edn 43 Introductory Statistics for Biology
John D. earthy 2nd edn R E Parker
4 An Introduction to Parasitology 2nd edn 44 Biology of Aphids A F G Dixon
R Alan Wilson 45 Biology of the Food Industry
6 Plant Taxonomy 2nd edn H. R Barn ell
V. H.Heywood 4 6 An Introduction to Animal Breeding
6 Microecology J . L. Cloudsley-Thompson John C. Bowman
7 Guts 2nd edn John Morton 47 Experimental Psychology : An
8 The Body Fluids and Their Functions Introduction fo r Biologi sts
2nd edn Ga~h Chapman Malcolm Jeeves
9 The Electron Microscope in Biology 4 8 Plants and Mineral Salts
2nd edn A V. Grimstone James F Sutcliffe and Dennis A Baker
1 O Translocation in Plants 2nd edn 49 Estuarine Biology R S K Barne s
Michael Richardson 50 Pest Control and Its Ecology
11 Muscle 2nd edn D. R. W1lk1e H F van Emden
12 Plant Breeding Wilham J C. Lawrence 51 The Ecology of Small Mammals
13 Understanding the Chemistry of the M J Delany
Cell Geoffrey R Barker 52 Phytoplankton A D Boney
14 Plants and Water 2nd edn 53 Bone and Bi omineral ization
James F. Sutcliffe K S1mk1ss
15 Developmental Plant Anatomy 54 The Biology of Plant Phenolics
Ala n R. Gemmell John R L Walker
16 Plan • .,ymbiosis George D Scott 55 Arthropod Vectors of Diseasa
17 Fungal Parasitism Brian J . Deverall James R Busv1ne
18 Population Dynamics 2nd edn 56 The Biology of Slime Moulds
Maurice E Solomon J M Ashworth and Jennifer Dee
19 Chemical Communication 57 Dor mancy and the Survival of Plants
John Ebling and Kenneth C. H1ghnam Trevor A Villiers
lO Human Genetics and Medicine 58 Ecology of Plants in the Tropics
2nd edn Cyril A Clarke Daniel H Janzen
21 Cell Division and Heredity Roger Kemp 59 The Optical M icroscop e in Biology
22 Animal Skeletons John D Currey Sav1le Bradbury
23 Investigation by Experiment 60 The Secretion of Milk Ben Mepham
0 V S. Heath 61 The Biology of Eucalypts
24 Animal Growth and Development Lindsay D Pryo r
David R. Newth 62 Marine Zooplankton John H W1ckste
25 Animal Photoperiodism Brian Lofts 63 Homeostasis Richard N Hardy
26 Natural History of Infectious Disease 64 Diseases in Crops B E J W heeler
J A. Boycott 65 Plant Tissue Culture
27 The Membranes of Animal Cells Dennis N Butcher and David S Ingram
2nd edn A P. M Lockwood 66 Lichens es Pollution Monitors
28 The Biology of Respiration 2nd edn David L Hawksworth and Fra ncis Rose
Christopher Bryant 67 Animal Asymmetry A Charles Nevil If
29 Size and Shape R. McNeil Alexander 68 Phytochrome and Plant Growth
30 Cellular Radiobiology Richard E Kendrick and Barry Frankland
Christopher W . Lawrence 69 Genetics and Adaptation E B Ford
31 Chloroplasts and Mitochondria 70 Population Cytogenetics Bernard Joi
Michael A Tribe and Peter A Whittaker 71 A Biology of Locusts R ~- Chap man
32 Fungal Saprophytism 2nd edn 72 The Dynamics of CompetItIon and
H J. Hudson Predation M ichael P Hassell
33 Animal Flight Colin J. Pennycu1ck 73 Mammalian Odours and Pheromone s
34 How Grasses Grow 2 nd edn D M ichael Stoddart
R. H M Langer 74 Decomposit ion C F Mason
36 Temperature and Animal Life 2nd edn 75 Viviparity Peter J Hogarth
Richard N. Hardy 76 Ecology of Fish es in Tropical Waters
36 Nervous Systems R H. Lowe-McConnell
Peter N R. Usherwood 77 Ecology and Archaeology
37 Photosynthesis 2nd edn Geoffrey W D1mbleby
D O Hall and K K Rao
38 The Biology of Pollution 2nd edn
Kenneth M ellanby l see al so 1ns1de rear cover)
The Institute of Biology's
Studies in Biology no. 122

The Ecology of
Streams and Rivers

Colin R. Townsend
B.Sc .• D.Phil.
Lecturer in Biology. University of East Anglia. Norwich

Edward Arnold
© Colin R. Townsend, 1980

First published 1980


by Edward Arnold (Publishers) Limited
41 Bedford Square, London WC 1B 3DQ.

All Rights Reserved. No pan of this publication may be


reproduced, stored in a retrieval system, or transmitted,
in any form or by any means, electronic, mechanical,
photocopying, recording or otherwise, without the prior
permission of Edward Arnold (Publishers) Limited.

British Library Cataloguing in Publication Data

Townsend, Colin R
The ecology of streams and rivers. - (Institute
of Biology. Studies in biology; 1u
ISSN 0537-9024).
1. Stream ecology
I. Title I I. Series
574.5'263 Q.H541.5.S7

Primed and bound in Great Britain al


The Camelot Press Ltd, Sou1hamp1on
General Preface to the Series

Because it is no longer possible for one textbook. to cover the whole field
of biology while remaining sufficiently up to date, the Institute of Biology
has sponsored this series so that teachers and students can learn about
significant developments. The enthusiastic acceptance of 'Studies in
Biology' shows that the books are providing authoritative views of
biological topics.
The features of the series include the attention given to methods, the
selected list of books for further reading and, wherever possible,
suggestions for practical work..
Readers' comments will be welcomed by the Education Officer of the
Institute.
1980 Institute of Biology,
41 Q.ueen's Gate,
London SW7 5HU

Preface

This book. is intended to give an introduction to the ecology of running


waters. The principal theme is· pattern - in the physico-chemical
environment, in the distribution of organisms and in community
organization. The aim is to define some general principles and unifying
concepts in lotic ecology.
Streams and rivers are important features of all landscapes. Their roles
in water supply, waste disposal and navigation make them especially
susceptible to human impact, and the need for effective river
management has provided a strong impetus for their study. They also
offer an admirable medium for fundamental ecological investigation at
.illl academic levels.
I wish to thank. DrsJ. H. R. Gee, R. D. Hey, A. G. Hildrew, M.A. Lock.
and K. H. Mann who kindly read parts of the text and made valuable
suggestions.
Norwich, 1980 C. R.T.
Contents

General Preface to the Series iii

Preface iii

Introduction
1. 1 The lotic ecosystem 1.2 Spatial pattern in ecology 1.3 The
biota oflotic ecosystems

2 The Flowing Water Environment 7


2.1 Introduction 2.2 The catchment area as a unit of s1udy
2.3 Pattern in the physico-chemical environment

3 Longitudinal Pattern in the Distribution of Organisms 18


3.1 Introduction 3.2 Abiotic factors, adaptation and dis1ribu-
tion 3.3 Biotic facwrs and distribution 3.4 The evolu1ionary
process 3.5 Adaptive suites 3.6 Conclusion

4 Patterns in Community Organization 37


4.1 AgeneralmodelofloticcommunityenergyAow 4.2 Longi-
tudinal pattern in community energy Aow 4.3 Conclusions

5 Microdistribution Patterns .50


5.1 Introduction 5.2 Vertical microdistribution in the sub-
stratum 5.3 Microdistribution on the substratum surface

6 Benthic Community Dynamics .56


6.1 Introduction 6.2 Experiments on temporal changes in
microdistribution 6.3 Invertebrate drif1 6.4 Milllc-r's
colonization cycle

References
1 Introduction

1. 1 The lotic ecosystem


Flowing freshwater environments are referred lO as lo1ic (lotus, washed)
to contrast them with standing water, lentic systems (leniJ, calm). In fan,
1he 1wo often grade into one another. For example, lakes which have a
lo1ic input and ou1pu1 are subject 10 a slow bu1 continuous 1hroughpu1 of
water. Thus, 1he major dis1inc1ion between lo1ic and lentic environments
is 1he relative residence times of water within them. The much more open
nature of lo1ic water bodies, with their continuous and rapid 1hroughpu1
of water and other materials, is a fundamental feature of such ecosvs1ems
and has consequences for all aspects of their ecology. '
Ecology has been defined as 1he experimental analysis of dis1ribu1ion
and abundance. This book is concerned with questions of spatial
dis1ribu1ion in 1he lo1ic ecosystem and will nm dwell on primarily
temporal phenomena such as life histories and population dynamics.
Dis1ribu1ional ecology can be inves1iga1ed on a geographical scale by
comparing the properties of rivers which are influenced by different
geological, topographical and climatic factors. The focus here will be
narrower still: by looking in detail at dis1ribu1ional phenomena within
individual rivers, 1he aim is lO identify some general principles in 1he
ecology of running waters.

1. 11 Spatial pattern in ecology


Each species has a res1ric1ed pauern of distribution on a global and
local scale. h does nol occur everywhere, in all habi1a1s. A principal aim
of ecology is 10 understand 1he factors 1ha1 determine the dis1ribu1ion of
organisms. These fall into three categories.
First, species may simply fail to reach an area because of restricted powers
of dispersal. This will be a significant influence for ecosystems which are
relatively isolated or al a distance from a source of colonists. Snakes are
believed to be absent from Ireland because the wave of recolonization
from Europe after the last glaciation only reached Scotland when the
land-bridge to Ireland had disappeared. This was also true for many
species of freshwater fish (VARLEY, 1967). It was probably only those
migratory species such as salmon (Salmo salar), sea trout (Salmo trutla) and
eels (Anguilla anguilla), that spend part of their life cycle in the sea, which
have since recolonized naturally. The other species which now occur there
have undoubtedly been introduced by man. Nevertheless, diversity in
freshwater fish species is still lower than on mainland Britain, which in
turn is less than on the European continent.
SI THE BIOTA OF LOTIC ECOSYSTEMS § 1.3
Secondly, biotic factors (influences of a biological nature) can be
significant in determining whether species can occur. Do particular
competitors or predators preclude occurrence and arc suitable food
resources available?
The final category, and invariably one of fundamental imponann.·,
consists of the abiotic factors in the environment (i.e. physical and
chemical). It has long been recognized that the communiry existing at ,1
given location is composed of species adapted to live under the a biotic
conditions that prevail there. Abiotic factors vary in intensity from place
to place. For each important abiotic factor a species possesses a range of
tolerance within which it can survive and reproduce. For example, it may
not thrive where the temperature is too high or too low or where water
velocity is too fast or too slow. Individual species display individual sets ol
tolerance ranges and it can be predicted that these will be reflected in their
different distributions in relation to patterns in the abiotic environment.
It is important to discover which environmental factors are significant
and how they affect distribution, on both a large and a micro-scale.
Chapter SI deals with patterns in both abiotic and biotic factors in the
environment and Chapters 3. 5 and 6 consider how far they are reflected
in the distribution of species. Chapter 4 is concerned with the con-
sequences of patterns in the distributions of individual species for
organization of the community as a whole.

1.3 The biota oflotic ecosystems


It will be useful at this stage to introduce the living organisms which
commonly occur in streams and rivers. These consist of the plants which
convert the sun's radiation into energy-rich compounds, bacteria and
fungi which are of crucial importance in the decomposition of dead
organic matter, and animals which feed on plants, on microorganism~
and on each other.

1 .J. I Plants
The large plants of freshwater environments are known as macrophJleJ.
The major groups are the flowering plants, mosses and liverworts, some
encrusting lichens and a few unusually large algal forms such as Chara.
Most of the aquatic higher plants are members of diverse families in
which the majority of species are terrestrial. An adaptive feature which fits
many of them for life under water is a good supply of aerenchyma tissue,
contributing to adequate internal aeration in the relatively low oxygen
levels and rates of diffusion which ocrur. Another feature of a submerged
way oflife is the low light intensity. Adaptations allowing relatively high
rates of photosynthesis under these conditions include absence of cuticle
from stems and leaves and high concentrations of chloroplasts in the
epidermal layer. Many plants, however, have floating leaves which are in
direct contact with the atmosphere. Some macrophytes, such as mosses,
3
ATTACHED ALGAE

(al
-~~
~

' -~'

Achnanthes Cladophora

Gloeotrichia tJ Lemanea
PHYTOPLANKTON
(fl (gl (hi (ii

X /lj
Scenedesmus Anabaena

(j)

,+)I
Pediastrum
(kl
Chlamydomonas

Diatoma
Asterionella

Fig. 1-1 Representative anached algae and phytoplankton of streams and


rivers. (a), (j) and (k) diatoms; (b) and (c) filamentous green algae; (d) and (g) blue-
green algae; (e) red algae; (f), (h) and (i) green algae. (Drawn 10 various scales.)

liverwons and lichens, may be found attached IO rocks, while others, like
the Aowering duckweeds are free-Aoating. Most of the Aowering plants,
on the other hand, are rooted in the substrate. These can usually be found
in lemic water bodies as well as in streams and rivers. In fact, no rooted
plant shows any special adaptation to running water. Those that occur
can do so because of their tough but Hexible stems and their creeping
stoloniferous or rhizomatous growth habit.
Algae are of great importance as primary producers in streams and
4 THE BIOTA OF LOTIC ECOSYSTEMS § 1.3

rivers. Atlachtd algae are associated with all kinds of sol id objects incl ucl ing
rocks (tpilithic algae) and macrophytes (epiphyticl, while others occur on
mud and silt surfaces (tpiptlic). This attached community consists
primarily of microscopic forms, but filamentous algae such as Cladoplrora
also occur (Fig. 1-1 ). The other main algal category is the phytoplankton.
This consists of free-floating microscopic forms and in most marine and
lake environments is responsible for the major part of plam production.
It assumes a much less important role in most regions of streams and
rivers where the rapid throughput of water makes it impossible for
phytoplankton populations to build up. In fact, any free-floating
microscopic algae found in upstream stretches are almost certain to be
attached algae which have become dislodged. However, in deeper
downstream regions phytoplankton may become important, with
diatoms often the dominant taxon.
r .3.2 Fungi and bacteria
Some bacteria are also able to photosynthesize. However, the most
crucial role played by fungi and bacteria (Fig. 1-2) is as decomposers of
dead organic matter.

(al
(b) (cl

".I..

Fig. 1-1 Characteristic microorganisms of streams and rivers. (a) The fungus
ClavariopsiJ (U/uatu:a showing release of tetra-radiate spores. (b) Developing spore
of the fungus Gyoer.lfJtlla sptciosa. (c) The aquatic actinomycete Actinoplanei.
(d-i) Some freshwater bacteria: Id) Micrococcw; (e) Pstudomonas; (0 Vibrio; (g)
Spirillum; lhl Chromalium; (i) Btggialoa. (Drawn to various scales.)
§ 1.3 THE BIOTA OF LOTIC ECOSYSTEMS 5
All organic substrates, whether derived from dead bodies or faeces and
whether they die in the water or fall in from outside, will be metabolized.
The microorganisms best suited to deal with solid matter are the fungi
and actinomycetes. These arc usually able to adhere to, or penetrate the
surface, and may produce extracellular degrading enzymes. The small
molecules released during this metabolism, and any soluble organic
compounds entering the stream from its catchment area, are rapidly
consumed by bacteria. Fungal spores and bacteria are very abundant in
stream and river water, so dead material is rapidly colonized.
Many aquatic fungi have branched spores, and the commonest type
consists of four long, straight arms diverging from a common point. This
tetra-radiate structure is believed to act like a minute anchor allowing the
spore to attach efficiently 10 a substrate even in turbulent water. The
spores are sometimes concentrated by air bubbles into a persistent foam
which may be seen below rapids and waterfalls. If the foam is collected in
a jar it quickly breaks down into bubble-free liquid. It is a straightforward
matter to mount a drop of this on a microscope slide and to examine the
spores it contains. Nearly all can be identified to species under a low
power, 16 mmobjeccive(1NGOLD, 1975).

t .J.J Animals
The animals oflotic ecosystems (Fig. 1-3) can be categorized according
to where they live. A few species are associated with the water surface, such
as water striders (GerriJ sp.): these are referred to as newton. Some species
dwell in mid-water. Powerful swimmers which occur are called nekton.
Those with limited powers of movement and more or less at the mercy of
the current are called z.ooplankton. Like the phytoplankton, zooplankton
will usually be most abundant in downstream regions. The commonest
zooplankton in lotic ecosystems are the Rotifera and the micro-
crustacean Cladocera and Copepoda. However, these usually contribute
little to the invertebrate biomass even in rivers. The majority of
invertebrate species are associated with the scream bed and are referred to
collectively as the benthos. Insects are more numerous than any other
invertebrate class. Other groups are generally less important, although
there may be many species or large biomasses in certain habitats. These
include sponges, flatworms, oligochaete worms, leeches, molluscs,
crustaceans and mites.
The principal vertebrate group is the fish. The majority of fishes which
occur in streams and rivers are also to be found in lentic ecosystems.
6

ZOOPLANKTON

Difflugia Bosmina
Brachionus

BENTHIC INVERTEBRATES

l11li lh~

~
~
~ . (IJ

,-.. .,.,...,,_;, ,,,,.,


Gammarus

Stenophylax
Simulium

(ml

Baetis

Chironomus
di'?::::
Epeorus

Fig. 1--g Representalives of the invertebrale fauna of streams and rivers. (a1
rotifer; (b) and (cl protozoan; (d) cladoceran; (cl copcpod; (f) oligochae1e worm;
(g) leech; (h) bivalve mollusc; (i) amphipod; ij) blackffy laiva; (k) mayffy nymph.
(I) cased caddis laiva; (ml stoncffy nymph; (n) chironomid laiva. (Drawn to
various scales.)
2 The Flowing Water Environment

11.1 Introduction
The one-way flow generated by gravity in streams and rivers has far-
reaching consequences for every aspect of their ecology. Characteristics
of flow which are influential are water velocity, stream discharge and
turbulence. Flow velocity is itself influenced by a combination of channel
characteristics such as slope, width and depth. The nature of the mineral
substrate, the concentrations of dissolved and suspended materials, and
temperature, are further factors which play important ecological roles.
The objectives of this chapter are two-fold. First, to show that these
intrinsic physico-chemical features of the lotic environment are partly
determined by extrinsic forces, and that it is not appropriate to view the
stream or river as an isolated entity. Secondly, to identify the patterns
which exist in the physico-chemical environment so that in later chapters
it may be possible to explain some associated patterns in the distribution
of species and in community functioning.

2. 2 The catchment area as a unit of study


A moment's reflection reveals that lotic ecosystems are not isolated
entities. Streams and rivers impinge on the atmosphere at their surface
and on the land at their edges, and it is across these boundaries that
movements of materials and energy take place. The principal materials
that enter the stream channel from the atmosphere are gases and water.
The latter is in the form of rain, sleet or snow, and contains dissolved
chemicals. However, most of the water input to stream channels drains in
.from the catchment area of the stream. On its way, the water picks up
dissolved and particulate inorganic matter derived from weathered rock.
A further input of material from the terrestrial system takes the form of
dead organic matter from plants and animals. This may be blown or
washed or may fall into the stream channel. Autumn-shed tree leaves are
of particular significance as an energy source for. freshwater micro-
organisms and animals, especially in streams running through wooded
countryside, SoJ'l"I-: less natural inputs, but often with just as profound
ecological consequences, come from man's activities. These include
sewage, agricultural run-off of fertilizers and pesticides, industrial
effluents, heat from power stations and rubbish of all kinds.
Much of the material entering streams and rivers passes through the
system to be discharged into the sea. Some, however, finds its way into
sediments on the river bed, and some becomes involved in ecological
processes and passes along food chains. Emerging organisms, water
8 THE CATCHMENT AREA AS A UNIT OF STUDY

vapour and gases move back from the wa1er 10 1he surrounding
environment.
It is easy to become engrossed in the biology of s1reams and rivers and
to ignore the properties of their catchment areas. However, 1he ecologisl
needs a broader perspective and in the last decade or so inves1iga1ors have
begun to treat the catchment area as a unit of study.
The most fundamental interactions between a lolic ecosys1em and i1s
catchment area concern the input of water and chemicals. The elements of
hydrology and water chemistry are discussed in a general way in 1hc
following section before proceeding 10 a more detailed 1rea1mem of a
particularly well-studied catchment in the U .S.A.

Derivation of strtam wattr


, ., . t
The water in stream channels is, of course, derived from precipi1a1ion ,
although only a proportion of this reaches 1he stream. The rest returns IO
the atmosphere through evaporation and through transpira1ion . A ven·
small proportion of stream water enters as rain falling directly inw 1hc
channel (channtl precipitation). The remainder derives from three indirect
pathways: surface run-off, interflow and groundwater discharge (Fig .
2-1).

(i) Surjact run-off. Precipitation falling on the land will ei1her infihrare
into the ground or move over the surface as surface run-off. The

Precipitation
il il tl tl
Evaporation
and
transpirati

Fig. 2-1 Diagram of the various pathways of water after arrival in a catchment
area.
THE CATCHMEl'IT AREA AS A UNIT OF STUDY 9
proportion in each category will depend on the infiltration capacity of the
soil surface and the rainfall intensity. In the case of impermeable ground,
such as hard bedrock or soil which is frozen solid, all the water will move
downhill as surface run-off. In highly permeable soils, such as sand, all
the incoming water is likely to infiltrate.
(ii) lnler.flow. Permeability often decreases with depth because of soil
compaction and filling of pores by washed down particulate material. If
an impermeable horizon underlies a permeable surface horizon, water
will How downhill parallel to the surface above the interface between the
horizons. This is referred to as interAow.
(iii) Groundwater discharge. Everywhere at some depth is rock that is
impermeable and watertight, because cementation and pressure have
closed up the pores. Water can sink so far and no further. It collects as
groundwater in the porous material over this impermeable layer, filling all
the pores and forming something analogous lO a saturated sponge. The
waler lablt - the top of this zone of saturation - rises until it is exposed in
the bottom of the deepest depression in the area, usually a stream
channel. Thus, the groundwater discharges into the stream.

2. 2. 2 Variations in flow
Groundwater discharge forms the major input of water 10 many
streams and rivers. This has important consequences for the nature ofthe
flowing water environment as a place to live. Water moves very slowly
through the ground. Therefore, the input of water lO stream channels lags
days or weeks behind the rainfall from which it is derived. The input will
be regular because it represents the 'overflow' from the slowly changing
groundwater reservoir. It is particularly important during long, dry spells
and ensures that most lotic ecosystems provide permanent habitats.
Of course, many stream channels do dry up periodically, particularly
in arid regions of the world. Such intermittent streams nevertheless
usually possess diverse communities. Some of the species they contain
survive by burrowing down into the damp substratum, while others
produce eggs or resistant stages which can withstand long periods of
drought. Often the stream channel retains isolated pools of water and
many species persist in these and recolonize the rest of the stream when
flow is restored.
The other extreme pattern of water flow is the flood. In times of very
high water discharge, the substratum of the stream bed is likely to be
scoured out and many animals and plants with it.

2.2.J Waler chemistry


The water of streams and rivers contains chemicals which influence the
nature or the biological communities they contain. Plant growth requires
IO THE CATCHMENT AREA AS A UNIT OF STUDY

a large number of elements, including significant amounts of nitro-


gen, phosphorus and potassium. Animals also have specific chemical
requirements (e.g. molluscs need sufficient calcium for shell growth),
while dissolved organic molecules provide an important energy source.-
for bacteria.
Dissolved gases play a crucial role in the metabolic processes or all
organisms. A proportion of gases enters by diffusion from the at-
mosphere or by natural aeration in turbulent 'white water'. However, life
processes within the stream have significant effects on gases in solution.
For example, the marked diurnal pattern of photosynthesis may be
reflected by considerable changes in oxygen and carbon dioxide
concentrations, and in situations where large amounts of organic waste
enter a stream channel as sewage its decomposition by microorganisms
causes severe depletion of dissolved oxygen.
Other dissolved and particulate chemical substances are derived from
two sources: (a) From rainwater which is not pure and may contain quite
appreciable amounts of many inorganic compounds; and (b) from 1he
water entering stream channels, much of which will have been in intima1e
contact both with soil and unweathered parent rock and as a result it
will have taken up inorganic and organic substances in solution and
suspension.
The chemistry of stream water is largely determined by the miner,11
nature of the catchment area. If the parent material consists of easily
eroded sedimentary rock, such as chalk, the concentration of minerals in
1hewaterwill be high. On the other hand, parent materials such as granite
are resistant to weathering and the associated stream water is low in
dissolved minerals.

::1.::1.4
The Hubbard Broolc Jludy
LIKENSand BORMANN (1972), in their important work on several small
catchments in the Hubbard Brook Experimental Forest, Nn,·
Hampshire, U.S.A. (Fig. 2-2), have quantified the inputs of water ,111d
chemicals from the various possible sources. Precipitation gauges were
placed throughout the catchment to monitor the amount of water
arriving and to determine the quantity of chemicals contributed in thio;
form. The stream outflow from each catchment area was measured at
specially constructed weirs, and its dissolved and particulate chemica I
loads were analysed.
The average annual precipitation was 12 3. 2 cm. Of this, 40. 796 returned
to the atmosphere by evapotranspiration, while 59.396 found its way into
the stream channels as groundwater discharge. Surface run-off and
interAow were insignificant because the watertight bedrock was overlain
by highly permeable sandy loam soil and glacial till. StreamAow varied
markedly with season. Sixty-eight per cent of annual How occurred
during snow-melt in March, April and May. In the warm summer
months, when forest growth was at a maximum and transpiration rates
30
co C=:J Precipitation
B
<ii - Streamflow
::,
C:
20
C:
co
0
Q)
CJ)
co 10
c
Q)
~
Q)
a.

~
(a) 0 1 km (bi

Fig. 2-2 (a) Map of the Hubbard Brook Experimental Forest in New Hampshire, U.S.A. The stream C',Hchmt·llls used i11 dwmkal
budget studies are indicated by dashed outlines. (b) Average annual prccipiiation ancl streamtlow, 1955-1969. (From t.lKFNS and
BORMANN, 1972.)
12 THE CATCHMENT AREA AS A UNIT OF STUDY

were high, streamffow was low. In autumn, transpiring surfaces were lost
when the deciduous trees shed their leaves: slreamffow increased
dramatically at this time.
Chemicals entered the streams via groundwater discharge in two
forms; 8096 dissolved and 2096 associated with particulate mailer. The
nutrient budgets in the catchments for some important elements art·
shown in Table 1.

Table I Average input in precipitation and output in streamRow for various


elements in the six catchments of the Hubbard Brook study ( 1963-1 969). The net
difference comes from weathering of parent material.

Precipitation Stream Net


input output difference
Element (kg ha-• yr-•J (kg ha- 1 yr- 1) (kg ha-• yr- 1J

Silicon very low 16.4 + 16.4


Calcium t.6 I t.7 +9.1
Sulphur u.7 16.2 +3.5
Sodium 1.5 6.8 +5.3
Magnesium 0.7 2.8 + 2.1
Potassium I.I 1.7 +o.6
Nitrogen 5.8 2.3 -3.5

A similar pattern is found each year. It shows that there is a signilican1


input of chemicals from the catchment area which in most casc·s is 11111< h
greater than could be accounted for by the chemicals in rain, sleet and
snow. This nel gain from the terrestrial environment is due IO weath1·ring
of parent material. Nitrogen is atypical because there are mechanisms ,11
work in the terrestrial system which tend to retain nitrogen-comaining
compounds with a high level of efficiency.
In a large-scale experiment, all the trees were felled in one or the
catchments. The enormous reduction in transpiring surfaces led IO 40%
more water reaching the groundwater and ultimately discharging inlO 1he
stream. The increased outflow contributed to greater rates of leaching of
chemicals and weathering of parent material. The rate of export of
dissolved inorganic substances from the disturbed watershed was thirteen
times the normal.
Clearly, it is not only the physico-chemical nature of the catchmem
area that determines chemical properties of stream water. The terrestrial
community also plays a vital role. One consequence of the dramatic rise
in the concentrations of plant nutrients was dense summer blooms of thl'
filamentous alga Ulothrix zonala. Undisturbed streams were essemiall)"
devoid of algae of this kind.
§ 2.3 PATTERN IN THE PHYSICO-CHEMICAL ENVIRONMENT 13
2.3 Pattern in the physico-chemical environment ~ 1,:, ( ,;
2 .J. r longitudinal pattern in lotic ecosystems k'"""' l ~
There is an interesting contrast in the principal panerns of abiotic
features in lotic and lentic environments. In deep ponds and lakes many
ecologically important physico-chemical factors, including temperature,
light intensity and oxygen concentration, vary in a vertical manner as
distinct depth profiles. In rivers, on the other hand, horizontal, or
longitudinal, patterns in the intensity of abio1ic factors can usually be
discerned from headwaters to mouth. These result largely through
downstream trends of decreasing river slope and increasing depth and
volume of discharge.

Flow velocity. Average How velocity (v) is influenced by river slope (s),
average flow depth (d) and resistance to Aow offered by the river's bed and
banks (0 in the following manner

v=flip
where g is the gravity constant.
Thus, water will Aow faster where the slope is steeper, where the How
depth is greater and where resistance to How is less. River slope is
generally steepest al the headwaters and declines progressively
downstream. If slope were the sole determinant then average Aowvelocity
would decrease with distance from the river's source. However, How
depth increases downstream as a result of increased streamAow as
tributaries join the main river. In addition, resistance to How decreases
(because the river becomes deeper and its bed consists of finer particles,
both reducing frictional resistance) and this also tends to increase average
flow velocity. Because the effects on How velocity of increasing depth and
decreasing resistance outweigh that of declining slope, the outcome of
these opposing forces in most rivers, except during bankfull and Hood
conditions, is for average How velocity to increase downstream (Fig. 2-3).
This may come as a surprise to many readers who believed that
headwater streams flowed faster than large rivers approaching the sea.
Headwater streams only appear to How fast in relation lo their shallow
depth whereas rivers appear to How sluggishly with respect to their
greater depth. However, it is true that many large rivers have been subject
to human manipulation of discharge using dams and lock-gates which
artificially reduce the average flow velocity.
Note that there is much variation in How velocity both across the width
of streams and rivers (usually maximal near the middle and susbstantially
reduced among macrophytes, if present) and also down the length of
short stretches. In small streams, for instance, shallow fast-flowing riffies
alternate with deeper pools. Flow velocity also varies with time, in
response to fluctuations in discharge during wet and dry periods of
weather.
14

Bankfull (once in 5 years)

an nual discharge
Mean

Distance downstream
Fig. 1-3 The relationship between average Row velocity and distance from 1h•·
headwaters.

Shtar strtss on tht strtam btd. Flow velocity declines in a logari1hmi<


manner from water surface to stream bed (Fig. i-4). Average velocity ol
the whole water column is actually equivalent to that occurring al abou1
60116 depth, and may be a significant measure for planktonic and nckwnH
organisms which live in mid-water.
However, avtrage How velocity is of little relevance to the bcndm
organisms which dwell on the bottom. What local forces will they he

Water surface

0·2
~
~ 0·4

Io 0·6
I
!!l I
"i: 1Mean
:::::, 0·8 I velocity
I
I

Units of velocity

Fig. 2-4 The relationship between Row velocity and depth in an open chanrwl
PATTERN IN THE PHYSICO-CHEMICAL ENVIRONMENT 15

subject to, and how will this vary downstream? River morphologists
prefer to talk in terms of shear stress on the bed (T 0 ) where
T0 = yds the constant y is termed the specific weight of water
In this equation, too, the factors How depth (d) and slope (s) vary in an
opposite manner in a downstream direction. However, slope declines
more quickly than depth increases (there could easily be a hundred-fold
decrease in slope but perhaps only a ten-fold increase in depth) and thus
for a given discharge frequency T 0 declines downstream.
The greater the shear stress on the bottom, the more likely an organism
is to be dislodged and swept away. Thus, it is clear that the shear stress
encountered by benthic organisms is likely to be a more influential factor
near the headwaters, despite the lower average How velocity there.

Substratum particle size. It is not just organisms that are affected by shear
stresses on the river bed. Particles in the substratum are also liable to
dislodgement. The particles that remain at a location will be those which
are larger than the flowing water is capable of carrying away. The smallest
ones to be dislodged will be carried in the column of water as suspended
load while the larger ones will be rolled or bounced along the bouom
as bed load. Since shear stress on the bottom generally decreases
downstream, it would be predicted that substratum particle size will
become smaller the further one moves from the source. This is exactly
what is found (e.g. Fig. 2-5).

200 ·'\,
100 \
~
\
.
e_g 50 \
\
\
a, \
N
'i)i 10 \
\
C 5 \
...
'iii
Cl>
\
\
C \
a, \
a, 1 \
~ ·5 \
\.
• \
\
.1
1 50 100 500 1000 5000
Distance from headwaters (km)
Fig. 1-5 Relationship between mean grain size of bed ma1erial and dis1ance
from headwa1ers in 1he Yellows1one-Missouri-Mississippi river sys1em. (Af1er
LEOPOLD, L. e. tt al. (1964). Fluvial Processes in Geomorpholot:J. W. H. Freeman, San
Francisco.)
16 PATTERN IN THE PHYSICO-CHEMICAL ENVIRONMENT

Substratum particle size is an abiotic factor likely to be ofimponann· in


the distribution of many organisms. Thus, we would expect shear strl'ss
on the bed to have indirect effects, as well as the direct effects memiont'd
in the last section, on patterns of distribution of animals and plants.
It should be noted that temporal variation in discharge and thus shear
stress will necessarily lead to changes in size class structure of the mineral
substratum.
Temperature. Water temperature also varies along the lengths of rivers,
mainly in response to the contrasting temperatures of soil and ail-. Thl'
soil is well buffered against the effects of changing air temperature and
incoming radiation and consequently its temperature is much morl'
constant than air temperature. Headwater streams which are fl'd by
groundwater will be at a temperature close to that of the suJTmmding
soil. In summer, this will be very much cooler than air temperature ,rnd
the cool water will be gradually warmed downstream by contact with the
air and by solar radiation. The rise is approximately propor1ional to the
logarithm of distance travelled (HYNES, 1970). In winter, however, soil
temperature near the headwaters may be warmer than air temperatun·
further downstream, in which case spring-fed streams may decrease in
temperature in a downstream direction. These are referred to a~ suml'lwr-
cold/winter-warm streams. The phenomenon does 1101 occur in s11-ca1m
which rise in impermeable ground and so are fed by surface nm-ofl~ and
it is not marked in the many streams which rise at high altitude where air.
soil and stream temperatures are cooler than downsu·eam throughout thr
year. The general pa11ern of increasing temperature holds also in tropical
regions. For example, many cemral African streams on high mountains
arise from ice-water at o°C and warm up as they Aow downstream.
The question of what aspects of temperature are ecologically important
is not.an easy one to answer. In some cases it will be the maximum and/or
minimum within which individuals of a given species can survive or
operate effectively, and then the downstream pattern of 1cmper,11ure may
have a straightforward influence on distribution. In less clear-nu
situations, however, it may be temperature at a particular stage in the life
history, or the diurnal or annual range in temperature which is most
important. Temperatures are most variable and change most rapidly in
response 10 changing weather in small, unshaded streams. The exceptions
arc spring-fed streams near their source where temperatures art'
remarkably constant, as explained above.
Dissolved oxygen. The concentration of oxygen in stream water is
influenced by a number of physical and biological factors. Solubility of
gases decreases with increasing temperature, so a downstream incrcast· in
the laner will cause the oxygen concentration 10 diminish. A second
factor contributing to a decrease in dissolved oxygen with distance from
source is the gradual change from turbulent, well-mixed water to
somewhat less well-mixed water. The biological proccsst•s which
PATTERN IN THE PHYSICO-CHEMICAL F.NVIRO:--.:ME:--.:T 17

inlluence concentrations of oxygen show no dear longirudinal pat1l-rn.


Photosynthetic production frequently leads IO supersaturated conditions
during 1he day, while clccomposi1ion of organic matter significantly
reduces 1he amount of oxygen prcsenl. The inpu1 of organic matter may
lake 1he form of autumn-shed 1ree leaves in headwater streams, or wash-
in of organic matter during lloocls and the discharge of human sewage
into larger rivers.
Dis50lved salts. The concentra1ion of ecologically important inorganic
sales often increases in a downs1ream direc1ion. Such a trend is likely 10 he
particularly marked in streams which rise in mountains composed of
metamorphic rock. This is resistam 10 wea1hering and 1he concentra1ion
of dissolved ma1erials in the scream water will be correspondingly low. As
1he river runs i1s course i1 may move into a region of sedimentary rock
which erodes much more rapidly. Concentra1ion of various inorganic ions
may also be increased by the input of human effluents in heavily
populated downstream s1retches.
Appraisal of tlie concept oflongitudinal patterns. This section has described a
'proto1ype' river which rises in the mountains, runs a relatively long
course and discharges into a lake or the sea. Of course, there is no such
1hing as a typical river. Many rise in relatively low-lying areas and some
moun1ain streams discharge into the sea while still at the 'torrential
stream' stage.
Nevenheless, generalizations about longitudinal patterns do hold for
1he majority of rivers. It would be surprising to find a river whose total
discharge actually decreased downstream, or one whose slope grew
consis1ently steeper. The continuous addition of tributaries and the
concave nature of the river profile rule out these configurations. The
consequent patterns in flow and substrate, and to a lesser extent
temperature and dissolved chemicals, are quite well documented and
may be considered to be acceptable general descriptions of lotic
ecosystems. However, it should not be imagined that the patterns occur as
smooth continua. They are frequently upset by local variations in
topography, causing sharp changes in gradient, and geology, altering the
chemical nature of the water. In addition, the confluence of a river with a
major tributary may result in dramatic modification of flow, substrate,
temperature and chemistry.
2.3.2 Small-scale pattern in lotir ecosystems
Pattern in the a biotic environment is also discernible on a small scale as
a mosaic within a very short stretch of stream or river. For example,
within a few square metres of headwater stream bed some patches may be
one or two centimetres deep, with fast flow and stony substrate, while
others with a depth of 50 cm or more are likely to have a slower flow
velocity, smaller shear stresses and finer substrate. Significant variations
in temperature or chemistry are less likely to occur on this scale.
3 Longitudinal Pattern in the
Distribution of Organisms

g. 1 Introduction
Certain abiotic factors vary more or less regularly down the leng1h of
rivers. The patterns are far from perfect, but nevertheless corresponding
longitudinal sequences in the distribution of many species are found.
Figure 3-1 gives examples from the macrophytes, several invenebra1e
groups and fish. The purpose of this chapter is to identify the abiotic and
biotic factors which are likely to influence such distribu1ion panerns.
There have been surprisingly few detailed studies of the mechanisms
underlying particular serial replacements of species. The bes1 understood
examples will be presented in the appropriate sections.

3.1 Abiotic factors, adaptation and distribution

3.2.1 Current
Although average flow velocity generally increases in a downstream
direction, current exerts its greatest influence on the benthic communi~-
in upstream regions. Here the water is usually turbulent and shallow,
exerting considerable shear stress on the stream bed. Thus, thosr
organisms which possess the most obvious adaptations lO 1·esist being
swept away are usually found upstream and particularly in wrremial
waters.

Plants. The only plants to be found in the most extreme flows arc
encrusting algae, such as Chamatsiphon fwcus, unbranched filamemous
algae, such as Ulothrix z.onata, and mosses and liverworts. However, these
squat species are susceptible to smothering by accumulating sih or sand
and, since they do not possess roots, they are easily washed away if the
substrate is subject to scouring. Thus, they cannot become established on
an unstable substrate. The adage 'a rolling stone gathers no moss' is
highly appropriate!

Fig. s-1 Longitudinal patterns in the distribution of: (a) certain macrophytes in
the River Swale (data from HOLMES, N. T. H. and WHITTON, B. A. (1977). Fmhwaln
Biolog:,, 7, 545-58); (b) stonefty nymphs (plecoptera) of the genus Ltuctra and
mayfly nymphs (Ephemeroptera) of the genera Ctntroptilum and Catnis in the River
Endrick. in Scotland (from MAITLAND, P. s. ( 1966). The fauna of the River Endrick.
Studits on Lode l.mnond II. Blackie, Glasgow); (cl fish in a Virginian stream in thr
U.S.A. (From BURTON, G. W. and ODUM, E. P. ( 1945). Ecolog:,, 16, 1h-g4.)
19

km down River Swale


1•1 o 50 100
~ Lemanea fluviatilis
Chyloscyphus polyanthos

<:::::O==- Rorippa nasturtium

<> ~ - - - - - . Phalaris arundinacea


Veronica beccabunga
- - ~ , Epilobium hirsutum
<:::===:::::~Myriophyllum spicatum
....__ _..., 1Potamogeton pectinatus

L-------- --==-.J Lemnaminor


km down River Endrick
(bi 0 25 50
Leuctra nigra
L. hippopus

L. inermis
L. geniculata

L. fusca

Centroptilum luteolum
- - - ----1 Centroptilum pennulatum
===-------i Caenis rivulorum

km down Little Stony River


(cl o 12 24
> Salvelinus f. fontinalis
c::>O Rhinichthys atratulus obtusus
0 0 Catonotus f. flabellaris
< Salmo gairdnerii irideus
Cottus b. bairdii
Cottus bairdii subspecies
Campostomus anomalum subspecies
Notropis albeolus
Rhinichthys cataractae subspecies
Catostomus c. commersonnii
U ABIOTIC FACTORS, ADAPTATION AND DISTRIBUTION

buoyancy by decreasing swim-bladder volume when exposed to


increased water velocity (BEREZ.AV and GEE, 1978). This compensates for the
lift created due to deflection of water over the dorsal surface of the body.

Studies of the relationship between occurrence of animal species and


flow velocity have generally shown that each is most numerous at a
particular velocity and progressively rarer in slower and faster currents.
This is also the case for certain macrophyte species (Fig. 3-2 ). For
example, the mosses with their firm rhizoidal attachment are com-
monest in fast flow, water celery <Apium nodiflorum) and spiked water-
milfoil in moderate How, and the free-floating duckweed (lemna minor) in
slow How. Nevertheless, it cannot be definitely concluded that flow
velocity is the sole factor determining distribution since other variables of
possible significance have usually not been studied .
.J . 1.1Substratum
'The nature of the substratum is stongly influenced by current and the
shear stress it exerts on the stream bed. Substratum particle size is greatest
Apium
Mosses nodiflorum

40t
-:r
t?
C
~
:i
0
n s m
20

0
n s m
'-'
8 Myriophyllum lemna
o... spicatum minor

j20[i]~tls:~
10

0
~4

0
~
nsmf nsmf

Flow velocity class


Fig. 5-1 Number of occurrences of four macrophyte taxa in each of four Row
velocity classes. n, negligible Row; s, slow- in which trailing plants hardly move;
m, moderate - in which the water surface is slightly disturbed; f, fast - in which
trailing plants move vigorously and the water surface is markedly disturbed.
(From HASLAM, 1978.)
§ 3.2 ABIOTIC FACTORS, ADAPTATION AND DISTRIBUTION 23

in torrential headwater streams, where only boulders arc present, and


gradually diminishes downstream as shear stresses on the bottom lessen.
Thus, flow velocity must often exert its influence on the distribution of
organisms indirectly through its eflcct on substratum.
The physical nature of the substratum can influence invertebrate
distribution in a number of ways. For example, species which arc adapted
to live in the crevices beneath and between stones can only occur where
the substratum is stony, while others that need a firm base for a sedentary
way of life, such as Simulium spp. and net-spinning caddis larvae, are not
found where the substratum is unstable or fine-grained. Conversely,
the burrowing nymph of the mayfly Ephemera simulans requires a
substratum of fine particles into which its modified front legs can dig
effectively. It has been shown in the laboratory that digging in its
preferred substratum of coarse sand involves the expenditure of less
energy than in pebbles or silt.
In regions of silry substrata, rooted macrophytes such as fennel
pondweed (Potamogeton pectinatw) and Canadian pondweed (Elodea
canadensis) are often common. The distribution of many invertebrates is
related, in turn, to the presence of macrophytes. In general, more
individuals and more animal species occur on plants than on nearby
mineral substrata. Some species or groups are confined to macrophytes.
For example, lan-ae of the chironomid Eucricotopus brevipalpis feed only
on broad-leaved pondweed (Polamogeton nalans). Rather surprisingly,
however, most invertebrates associated with macrophytes do not feed on
the plant tissue. Some graze on the epiphytic algae which grow on the
surface of the macrophyte, others use the plant as a stable site from which
to filter food from the passing water, and others are predators. Artificial
weeds, made of plastic or string, generally become colonized by the same
invertebrate community as their genuine counterparts. This emphasizes
the role of macrophytes as a substratum rather than a food supply. The
planes are also of importance to many fish providing a surface on which to
attach their eggs.
A detailed study which demonstrates the significance of substratum in a
particular serial replacement of species is that of BOVBJERC (1970). He
undertook field and laboratory investigations to discover the reasons for
the disjunct distribution of two closely related species of freshwater
crayfish in a short stretch of the Little Sioux River in the U.S.A. 0rconectes
immunis, which normally inhabits ponds, occurred in the headwaters.
These waters were atypical, being somewhat pondlike in character and
drying up in the summer. The lower region of the study section consisted
of a gravel and rock bottom. 0. virilis occurred here. The middle region
consisted of a series of pools alternating with riffles and contained both
species.
Several factors were implicated in the distribution pattern. Only
0. immunis has the capacity to burrow, necessary for sUJvival when
the headwaters dried up. Also, it is less susceptible to the lowered
24 ABIOTIC FACTORS, ADAPTATION AND DISTRIBUTION

oxygen levels which occasionally occurred in the pond-like headw,ll('r\.


The most impor1an1 difference, however, rcla1cd to substratum
prererence. Figure 3-3 gives 1he results of labora1m·y experiments i11
which 1hree 1ypes of subs1ra1a were available: rocks, gravel ,111d mud .
When ei1her species was tested alone, 1he results were almost identiral
showing a s1rong preference for rocky substratum. When both specin
were present, however, the preference of 0. immzmiJ but not 0. vmh .1,,.,1\
apparently reversed. Evidently, a large number of aggressive e1Kou11tns
wok place and 0. viriliJ consistently drove off 0. immzmiJ. Rock crl'\'iln
are important resources for crayfish, as shelter from predation a11d
cannibalism (particularly while moulting), and also during spates.
The distribution of these two species in 1he field can be explainl'd
because 0 . viriliJ was able 10 obtain and successfully defend the primr
rocky sites in the lower region, whereas 0. immrmiJ but not 0. vm/1.1 could
survive under the abiotic conditions which prevailed in the hcadw.ttl'I\
The middle region contained patches of both son of habitat and ,11,o
populations of both species.

J.1.J Temperature
II is well known that temperature regime changes with distancl' from
river source. For example, maximum summer temperatures tend tu
increase downstream, and this is panicularly marked for stream\ whi1 h
rise in the mountains. We may postulate that the serial rcplan·ment ol
some species will be related to a longitudinal pauern in tempcratun·

60
"C
Ill
'E
8 50
-
.,,f
.gC 40
'iii
0
~ 30
l!!
g
o 2
Ill
Cl
"'
-
E 10
~
:. 0
IIICIQ rNl'fM --1 I- _,.
I
- - . .....
,...
0. virilis 0. ,mmunis Together
Fig. 3--3 Substratum preferences (rock, gravel, mud) for the crayfish species
0rcorucles virilis and ·0. immunis when alone and when together, expressed as a
percentage of total positions recorded. (After BOVBJ ERG, 19 7o.)
ABIOTIC FACTORS, ADAPTATION AND DISTRIBUTION 115

Clear rela1ionships be1ween ahimde and 1he occurrence of cenain species


have been demons1ra1ecl in several field s1udies: for example, 1he leeches
Trocl1tla bykowski and Erpobdtlla mono,triata occur only al high ahi1udcs in
Poland, while 01hers, such as Trochtla subviridis and Erpobdtlla octoculata,
occur only al low ahiludcs. Findings such as 1hese have usually been
imerpre1ed as a consequence of different 1cmpcra1urc wlcranccs. How-
ever, ii is dangerous lO assume 1ha1 a corrcla1ion bc1ween dis1ribu1ion
and 1empcraturc implies a causa1ivc link; 01her imerpreta1ions can nm be
ruled out. In particular, 1he significance of 1he imima1c rcla1ionship
between 1empcraturc and dissolved oxygen conccmralion needs 10 be
considered. It is sensible, 1hcrcforc, 10 discuss these iwo faC1ors together.

J.2.4 Oxygtn and ltmptraturt


Jus1 as flow velocity and subs1ra1e type arc imimately related, so are
oxygen conccntra1ion and 1empcra1urc. This is because 1he solubility
of oxygen in water decreases with rise in 1cmpera1ure. Thus, the
conccmration of oxygen is often likely lO be lower downs1ream where
temperature is higher. In addi1ion, water flow 1ends to become less
turbulent downs1ream so that the oxygen removed from the water by
respiration of organisms, and particularly during decomposition pro-
cesses, is less readily replaced 1han in the well-mixed headwaters. This is
especially significant for animals which inhabit the. sediment, with its
abundance of decomposing organic matter. Special adaptations to life
here include, for example, 1he presence of haemoglobin in the bodies of
'bloodworms' (chironomid la,vae). Evidently this makes oxygen uptake
possible ,11 the very low oxygen levels in the soft river sediments in which
they occur.
Despite large daily increases in oxygen concentration which may occur
because of i1s release during photosynthesis, oxygen availability is
generally lower downstream, particularly at night.

Fish. The longitudinal distribution pattern of fish may in part be a


reflection of this oxygen availability. For example, British species typical
of upstream regions, such as brown trout, minnow (Phoxinus phoxinus),
bullhead, grayling (Thymallus thymallus) and chub (Ltuciscus etphalus),
require 5-11 ml 1- 1 of oxygen to survive. Species which occur further
downstream, such as roach (Rutilus rutilus), ruffe (Gymnoctphalus ctmua)
and pike (Esox Lucius), require 4 ml 1- 1• Whereas those, such as carp
(Cyprinus carpio), tench ( Tinca Linea) and bream (Abramis brama), found in the
deepest and least turbulent flows, can survive in water containing as little
as 0.5 ml 1- 1 of oxygen. These are approximate values at average
environmental temperatures for the species in question (VARLEY, 1967).
The precise oxygen requirements of a species are largely dictated by
temperature. Figure 3-4 shows the relationship beiween standard
metabolism and temperature for three species representative of a
downstream series. The rainbow trout (Salmo gairdntri) has a relatively
16

'
.s::.
-;- 100
Cl

<5
§
E
.!!! 30
:g
ea
il
E
-e
{g 10
C
ea
in

Temperature •c

Fig. s-4 The relationship between standard metabolism and temperarun· lo,
three species offish, typical of a downstream series. (From VARU:Y, 19fi7 .1

high metabolic rate which only increases three-fold between sand 20°C
It is able to remain active and therefore to feed in cold water, bur ii\
metabolism is already so high that it cannot suivive as tempera1ull'
increases, and will die at approximately 13°C. It can be described a\ ;1
cold-water stenotherm, able to exist only in a small range ol
temperatures. At the other extreme, the eurythermal goldfish (Cara111111
auralw) shows a considerable rise in metabolism with increase 111
temperature and can suivive over a wide range. Its upper lethal
temperature is approximately 41°C. However, it is probably inactive in
very cold water. The perch (Ptrca Jluvialilis) represents an intermedia1r
category. Its upper lethal temperature is approximately 33°C.
Table I shows the temperarures that are optimal for growth and typical
for spawning of cold-water stenotherms, intermediate species and
eurytherms from temperate regions. These concur with the downstream
distribution of the species.
In physiological terms, the adaptation to life in a particular
temperature range seems to lie in the adjustment of basal metabolism 10
those temperatures. A similar phenomenon has been described for an
invertebrate series.
ABIOTIC FACTORS, ADAPTATION AND DISTRIBUTION 27

Table 2 lmporiant 1empera1ure limi1s charac1eris1ic of 1hree fish categories in


1empera1e regions. (After VARLEY, 1967.l

Upper Optimum Typical


lethal for for
limit gTOwth spauming

Cold-water stenotherms < 28 7-17 <10


(e.g. 1rou1, grayling)
Imermedia1e ca1egory 28-34 14-23 >10
(e.g. perch, pike)
Eurytherms >34 20-28 >15
(e.g. carp, 1ench)

Invertebrates. Six species of net-spinning caddis in the family


Hydropsychidae show a marked regularity in their downstream sequence
in British rivers (Fig. 3-5). In large, unpolluted rivers, the typical order of
first appearance from source to mouth is: Diplectrona Jelix, Hydropsyche
i,utabilis, H. siltalai, H. pellucidula, H. contubemalis and/or Cheumatopsyche
lepida. It is possible that the sequence originally had two further species,
H. guttata and H. e;,cocellata. These were taken as adults from the lower
reaches of the River Thames at the beginning of this century but have
since disappeared, presumably as a result of pollution. In the River
Garonne in France, the sequence does indeed conclude with them.
HILDREw and EDINGTON ( 1979) gathered data on summer temperatures
in a number of stations on the River Usk in Wales. In addition, they
performed laboratory experiments to determine the relationship
between respiratory rate and temperature for three of the species in the
sequence, D.felix, H. instabilis and H. pellucidula. The mean temperatures
to which these species were subjected in nature (June-September) were
11.2, 13.5 and 14.2°C respectively, reflecting their distributions
progressively further downstream. A typical D. Jelix station during the
summer had low temperature maxima and a small range, an H. instabili$
station had higher maxima and a large range, while an H. pellucidula
station had the highest maxima and a small range.
There are striking differences between the species in the relationships
between temperature and respiratory rate (Fig. 3-6). Steepening of the
curve occurs at progressively higher temperatures so that, for example, a
respiratory rate of 1 mg 0 2 g- 1 dry weight h- 1 occurs at 14.5°C for D.felix,
16°C for II. instabilis and 21°C for H. pellucidula. The zone of relative
temperature independence - the flat part of the curve - matches the
progessively higher temperature regimes experienced by each species in
summer. Thus, in a manner similar to the fish discussed above, the
metabolic rate of these invertebrate species is 'adjusted' to the
28 ABIOTIC FACTORS, ADAPTATION AND DISTRIBUTION

temperature regime ora partirular section of river. It is inten·sting to nott·


that since respiration energy losses are lower at the appropriatt'
temperature, more of the energy assimilated from food can be
incorporated into growth and reproduction. This may provide' the ba\i~
for each species possessing a competitive adva111age in its typical river
section.

Abergavenny
N 1 10km

"'

Fig. 3-5 The distribution of hydropsychid larvae (bold lines) in tributaries and
lower reaches of the River Usk in Wales. (From HILD REW, A. G. and EDINGTON ,J. M
( 1979).J. Anim. Ecal., 48, 557-76.)
29

~
feH, I
Diplectrona Hydropsyche
;n,,,o;,;,
Hydropsyche
pellucidula

-7{ -~
5 10 15 20 25
Temperature °C

Fig. 3-6 Respiratory rate/temperature relationships for fifth instar larvae of


three hydropsychid species. (From HILDREWand EDINGTON, 1979.)

J.2.5 Concentration of diHolved chemica/.J


Increasing chemical concentrations along the length of rivers may
influence the distribution of certain species which require particular
levels of inorganic ions or plant nutrients. For example, most molluscs
secrete shells of calcium carbonate, and the majority of species are not
found in water with less than 20 mg calcium 1- 1. Evidently, the greater
abundance of species found in the lower reaches of the River Endrick in
Scotland is due to the higher levels of dissolved calcium downstream. The
distribution of plants may also be related to nutrient concentration. For
example, the alternate-flowered water-milfoil (Myriophyllum allernijlorum),
which is typical of nutrient-po_or (oligotrophic) conditions in mountain
streams, is often replaced downstream by its congener, the spiked water-
milfoil (M. Jpicatum), a species requiring much higher concentrations of
nutrients (eutrophic conditions).

3,3 Biotic factors and distribution


Clearly, the distribution of a species is not simply determined by its
tolerance for particular ranges of physical and chemical factors. Biotic
factors are also involved. For example, the presence of appropriate food
resources is crucial in determining the distribution of animals, bacteria
and fungi. (The feeding relationships of running-water organisms will be
dealt with in the next chapter.) The space which a species can occupy,
within the limits of its tolerance for all abiotic factors in the environment
and where essential biological resources occur, is referred to as its
fundamental niche. However, the boundaries for existence are not
necessarily set by these factors alone. The fundamental niche may be
30 BIOTIC FACTORS AND DISTRIBUTION § 3.3
compressed as a result of the influence of other species competing for the
same resources. The resulting occupied niche space is called the realised
niche .
.J ..J. r Competitive interactions
It is clear from Bovbjerg's work (see § 3.2.2), that although the
fundamental niche of the crayfish 0rconectn immuni5 is determined part Iv
by its tolerance of the pondlike abiotic conditions which happen to occur
in the headwaters of the study site, there appears 10 be no reason why i1
could not survive and reproduce in the lower, stony stretch. In fan, it is
here that its preferred substratum is found. It fails to occur in the lower
section because the aggressive 0. virilu excludes it. Thus, the reali~ed
niche of 0. immunu is smaller than its fundamental niche. On the otlwr
hand, 0. virilis would probably not persist in the headwaters or this
stream, since it is not tolerant of the low oxygen concentrations which
occasionally prevail and it is unable to burrow to survive dry periods.
Thus, its fundamental and realised niches coincide. This kind of cliren
physical interaction is called interference competition. Comparable example,
can be found among plants. For example, if a patch of bur reed cle\'dop,
upstream ofa clump of water crowfoot, the latter grows poorly became 11
is sheltered from the full flow and has silt deposited on it.
A second kind of interaction, exploitation competition, is more subtle and
less easily measured. The term refers to the indirect competition that takn
place between species utilizing the same limited resource. Clear!\',
consumption by one species will reduce the amount of resource available
to the other, and the predicted outcome is that the species whid1 is It',,
efficient at 'converting' the resource to reproductive output will be dri\'l'll
to extinction. This is known as the competitive exclusion pri11cipfr.
Exploitation competition appears to play a significant role in lhl'
distribution of certain stream-dwelling flatworms (Tricladida) - ,1 group
that has received particularly close study.

Competition andflatworm distribution. Three flatworms succeed each other


in reasonably clear-cut zones along mountain streams in southeastern
France: Crenobia alpina occurs above 800-900 m altitude, Polyulis fr/ma
between Soo m and 300 m and Duge5ia gonocephala below 300 m and in
swiftly flowing cool streams in the plains.
Experiments were performed to determine for each species the upper
and lower lethal temperatures and limits for reproduction (PATTEE et al.,
1973). In fact, none of the species showed a lower lethal temperature.
They could all be frozen into blocks of ice and thawed out undamaged!
The upper lethal limits (determined as 50% of individuals dead in
indefinite time) increased with downstream order of distribution of the
species, as might be expected. Thus, D. gonocephala, which is subject in
nature to the highest temperatures, had the highest lethal limit of 21 °C.
The value for P.ftliM. was 16°C, and for C. alpiM., which is subject to the
BIOTIC FACTORS AND DISTRIBUTION 31

lowest temperatures, 12°C. However, these upper lethal limits cannot


usually explain flatworm zonation since most habitats do not become hot
enough to eradicate them.
C. alpina can reproduce both sexually and asexually, while re-
production by D. gonocephala is exclusively sexual, and by P. felina,
asexual. To enable direct comparisons to be made between the species an
auempt was made to quantify reproductive capacities at temperatures
between 5 and 20°C. The parameter used depends on binh rate and
generation time and is calculated as the intrinsic rate of natu:-al increase
of the population in an unlimited environment. Figure 3-7 shows how
the intrinsic rates of increase of the species varied with temperature. The
temperature range over which each species demonstrated reproductive
superiority conforms with that recorded in the literature for the animal's
habitat.

3lIll ·02
I!!
u
·=iii
:i
;;
C
0 ·01
~
:!
u
"iii
·2
.5
Temperature •c
Fig. 3-7 The intrinsic rate of natural increase of three Ratwonn species,
measured in the laborawry at different temperatures. (After PATTEE tl al., 1973.)

A possible basis for reproductive advantage is suggested by the results


of experiments on two of the species, C. alpina and D. gonocephala. C. alpina
had a lower metabolic rate at temperatures between 5 and 8°C, while the
rate for D. gonocephala was lower between 8 and 18°C. These correspond
closely to the temperature ranges where reproductive superiority occurs
in the two species. It can be postulated that, like the caddis larvae
discussed in§ 3.2.4, each species is able to convert a greater proportion
of assimilated food into growth and reproduction at temperatures where
its respiratory losses are minimal. Note that this explanation requires that
feeding rate is not depressed at the temperatures in question.
The conclusion of PATTEE et al. (1973), that the distribution of each
flatworm species is dictated by whether it possesses a competitive
advantage over the others, is persuasive. However, such an explanation
3!! BIOTIC FACTORS AND DISTRIBUTION § 3.3
suffers from two limitations. First, it assumes that other differences in the
flatworms' ecologies are unimportant. Secondly, the experiments were
performed with an excess of food and in the absence of competition by
the other species; conditions in their natural environment are often very
different from these. Work carried out in Wales has a bearing on both
these points.
LOCK ( 197 5), in attempting to explain the downstream replacement of
C. alpina by P. Jelina in Welsh mountain streams, was able to show in the
laboratory that P. Jelina was less proficient at colonizing and moving
within a simulated steep-gradient substrate. He suggests, therefore, that
the absence of P. Jelina from such steep-gradient, upstream reaches of
streams was due to their behavioural inability to tolerate the current
regimes present there. But why should C. alpina be replaced by P.felina in
the shallower gradients downstream? An earlier study had suggested that
distribution could not have been simply limited by temperature in these
cases. For example, the temperature regimes at stations containing P
felina, but not C. alpina, were similar to a few sites where P.felina happened
not to occur and where C. alpina was taken. Thus, the fundamental niches
of both species included the downstream region. Could P. felina be
competitively excluding C. alpina in these stretches, confining it 10 a
more limited realized niche?
This is the question addressed by LOCK and REYNOLDSON ( 1976). It is a
difficult task to obtain conclusive evidence that interspecific competition
is occurring in nature. The following criteria, among others, are con-
sidered necessary to demonstrate the existence of the phenomenon.

(i) The comparative distributions of the two potentially competing specie\


should be explicable on the basis of competition.
(ii) The species should be shown to be utilizing a common resource.
(iii) The outcome of experimental manipulations of both the resomce
and the species populations should be consistent with predictions based
upon the competition hypothesis.

LOCK and REYNOLDSON (1976) managed to gain information largclv


satisfying the criteria in a zone of overlap containing the two flatworm
species. Figure 3-8 shows the distribution of C. alpina and P. Jelina in the
Nam Heilyn stream, illustrating the typical replacement of C. alpina by P.
felina in the shallower gradients downstream. The largely disjunct
distribution could possibly be explained on the basis of a competitive
interaction, so criterion (i) above was satisfied.
The second problem was to determine whether the two species utilized
the same resources. The feeding action of flatworms involves the
ingestion of prey body fluids and minute particles so that a simple
identification of gut contents is not possible. The authors' ingenious
approach was to use a serological technique, the 'precipitin test'. This first
involves injecting a rabbit with a protein solution of the suspected food
33

]:
a;
1"'183
Cl)
II)
Cl)
>
0
.0
"'

a I
1-5
km downstream

Fig. 3-8 The distribution of Crenobia alpina and Polyctlis Jelina in the Nam Heilyn
stream in Wales. (Percentages indicate stream slope.)

organism (antigen), to which antibodies are produced in the blood serum


(antiserum). The potential Aatwonn predator is macerated and used in
the precipitin test. If it contains the appropriate antigen (i.e. if the
predator has recemly eaten the suspected prey), when combined with the
antiserum a white precipitin band will occur.
Lock and Reynoldson analysed the diet of three Aatwonn populations,
a control population of C. alpina which occurred on its own, and an
experimental population of C. alpina which occurred together with P.
felina. The results are shown in Table 3.

Table 3 Food of the three Aacworm populations (positive immunological tests


expressed as a percentage of total positive tests).

Total positive
reactions per
Gammaru.s Ephemeroptera Pltcoptera Trichoptera individual

Control C. alpina 44 0 25 31 1·10


Mixed C. alpina 8 0 38 54 0-58
Mixed P.felina 32 4 24 40 0·7 I

The diets of the control C. alpina population and that of P.Jtlina were
broadly similar, so criterion (ii) was satisfied. In addition, it is clear that
although the diet of C. alpina in both the absence and presence of P.felina
contained comparable proportions of stonefly nymphs (Plecoptera) and
34 THE EVOLUTIONARY PROCESS

caddis larvae (Trichoptera), in the presence of P.felina it failed to gain its


'normal' complement of Gammarus. Note also that the total number of
positive precipitin reactions per individual C. alpina was reduced by
almost half in the presence of P. Jelina. This suggests that C. alpi11a was
under competitive pressure in the presence of P.felina with the reduction
in positive reactions possibly due to the more successful feeding of P.
Jelina on Gammanu.
Finally, the authors were able to take advantage of a natural
experiment, when a sudden decrease in a numerically superior
population of P. ftlina (possibly due to the consequences of severe rains)
resulted in a rapid increase in the numbers of a previously inferior
population of C. alpina. This was considered to go some way 10wards
satisfying criterion (iii), and to be good evidence of interspecific
competition.
Although the evidence for some of the criteria for interspecific
competition is not strong, they argue that taken together their resuhs
suggest that competition was occurring, and 1ha1 it could accoum for the
absence of C. alpina in a majority of shallow-gradient habitats. Evidently,
the competitive superiority of P.felina involved its more effective feeding
performance, but a temperature-related superior rate of reproduction, as
suggested by PATTEE et al. ( 1973), cannot be ruled out in all cases. II i~
possible that the species can coexist in zones of overlap because of
continuous immigration of C. alpina from its optimal environment
upstream, and P.felina from downstream.
The flatworm work illustrates a number of important points. First, it
indicates the significance of competition in determining the distribution
of species. Secondly, it stresses the fact that abiotic and biotic factors ma,·
interact in a rather complex manner. Thirdly, it suggests that species
distribution patterns may be influenced by different factors in di£frre111
geographical locations.

3.4 The evolutionary process


According 10 the competitive exclusion principle, species with identical
niche requirements cannot indefinitely coexist. The predicted outcome is
that one of the species would make more effective use of the common
resources, and if these were in limited supply, would drive the other to
extinction. In theory, therefore, the coexistence of species should
generally only be possible if there is a minimum ecological difference
between them. In practice, coexisting species are often found to differ in
their diets, habitat preferences or life cycles. In many cases, such mource
partitioning will occur simply as a result of a random association of species
with complementary characteristics. However, in other cases resource
partitioning may have arisen as an evolutionary response to competitive
pressure. This is likely particularly in the case of closely related species.
Thus, when two congeneric species came together in the past, although
ADAPTIVE SUITES 35
their ecologies would have been broadly similar it is unlikely 1ha1 the
overlap in resource use would have been absolute. Those individuals
from one species which were most different in their requirements from
the other species would produce more offspring and 1ransmi1 more genes
10 future generations. The species would gradually diverge and the
intensity of competition would be reduced. Thus, a possible evolutionary
consequence of competitive interactions is the diversification of the
fundamental niches of the species in question. This is the explanation
applied 10 the amazing degree of specialization observed in Darwin 's
finches on the Galapagos Islands. These species mainly differ in bill size,
and consequently in the size range of food that each takes.
It can be argued 1ha1 among freshwater organisms the successful
individuals must often have been those 1ha1 differed from related species
in their tolerance limits for particular physical and chemical parameters.
As a result of the longitudinal panerns in lo1ic environments of certain of
these abio1ic characteristics, niche diversification would sometimes have
acted 10 distribute closely related species in a serial downstream
replacement, as observed in the examples discussed above.

3.5 Adaptive suites


II should now be clear that although a single environmental factor may
appear to predominate in determining the distribution of particular
species, nevertheless a suite of adaptations is invariably involved. To be
successful in a given environment a species must possess appropriate
morphological, physiological and behavioural adaptations.
Thus, the metabolic adaptations of the trout are necessary for its
survival in the cool, well oxygenated headwaters of rivers where it occurs.
However, these adaptations are not sufficient by themselves. It is able to
maintain its position in the current because of its streamlined shape and
swimming capability, and breeding is possible because the female can
construct a nest (or redd) in the gravel to protect the eggs. Similarly,
invertebrates may possess morphological adaptations which are clearly
important in a particular environment, such as the streamlining of the
mayfly nymph BaetiJ rhodani or the broad burrowing forelegs of another
mayfly, Ephemera Jimularu. However, these species also possess the
behaviour panerns which render the adaptations effective, and their metab-
olism is adjusted to the environments they inhabit. Similar considerations
apply 10 plants. For example, a species that can exist in oligotrophic con-
ditions will nevertheless not be found in an oligotrophic, upstream river
section unless it can also withstand the rather extreme Row conditions.

3.6 Conclusion
While most studies in particular rivers have revealed longitudinal
patterns in distribution, it is important to note that these patterns are not
36 CONCLUSION

reproduced in different rivers. As was noted in Chapter 11, there is no such


thing as a typical river. Rivers rising in mountainous areas, with resistant
rock, would not be expected to exhibit the same longitudinal changes
as those rising in lowland chalk or clay formations. In addition,
downstream patterns in the abiotic environment are frequently disrupted
by the activities of man. In general, though, it is uncommon for any
species to be represented continuously from river source to mouth. The
aim of this chapter has been to demonstrate, by considering detailed
studies of closely related species, which abiotic and biotic factors are of
importance in determining distribution.
4 Patterns in Community
Organization

Many important ecological questions, such as those discussed in Chapter


3 about distribution pa11erns, need 10 be addressed at the level of
individual organisms or of species populations. However, it is also
important to understand ecological processes characteristic of whole
communities. A fundamental feature of the functioning of communities
is the flow of energy through them, since this is the driving force for the
biological processes taking place. A good introduction to the subject of
ecological energetics is provided in an earlier volume in this series
(PHILLIPSON, I 966).
The energy-flow approach to ecology views organisms as energy
transformers which take part in an intricate web of interactions. The
biotic community relies on an energy base for successive trophic levelJ. In
most terrestrial systems the energy base is contributed by plants which
convert solar radiation by photosynthesis into high-energy organic
molecules. In streams and rivers, however, a substantial proportion of the
energy base is often provided as dead organic matter which has been
elaborated outside the system (allochthonous input), and this can t-e
contrasted with the internal production of organic mailer by
photosynthesizing macrophytes and algae (autochthonous input). The
available organic matter, both living and dead, is processed by a great
variety of heterotrophic organisms, including the microscopic bacteria
and fungi, invertebrates and fish, all interacting with each other in a
complex manner.

4.1 A general model oflotic community energy Bow


For the purposes of a community energetics study it is not particularly
useful to categorize organisms into taxonomic groups. Their functional
role is the significant thing and each may be assigned to a particular
trophic grouping. Macrophytes and algae (microphytes), both attached
and planktonic, make up one group - the autotrophs. The microorganisms,
principally bacteria and fungi, constitute a second. Four functional
categories of invertebrates can be distinguished: graurs - herbivores
feeding on attached algae; shredders-animals feeding on large particles of
plant material; collectors - feeding on fine particles either on the stream
bed or filtered from the water, and finally, predators (CUMMINS, 1974).
Vertebrates can be categorized in a similar way.
Figure 4-1 illustrates the interactions between these groups and three
categories of organic material - dissolved organic matter (DOM,
arbitrarily defined to be <o.ooo 45 mm particle size), fine particulate
38 GENERAL MODEL OF LOTIC COMMUNITY ENERGY FLOW § 4.1

organic matter (FPOM, < 1 mm) and coarse particulate organic matter
(CPOM, > 1 mm and up to whole leaves, twigs, etc.). The CPOM and
FPOM compartments are taken to include the microorganisms which are
intimately associated with them.

LIGHT-Macroph/ytes CPOM~.

fiJ "'!>~,~~hre~dders~
1:.\.":"Ii,
1-,
ALLOCHTHONOUS _ _ _.. DOM11occu1a1, FPOM-Collectors-Predat~
INPUT m,crobial
Kl,on

Fig. 4-1 A general model of community energy Row in lotic ecosystems. (Several
arrow~ have been omitted for clarity-see text for details.)

Allochthonous organic matter enters a particular river stretch eithl"r


from upstream in the water column or from the surrounding catchme111
area in groundwater or by overland flow. The input consists or two
principal components- organic matter dissolved in the water, and rnarsl"
particles such as tree leaves, fruits and twigs. Of the autochthonous
production, only an insignificant proportion of macrophytes arc l"aten
alive. Most enters the CPOM compartment and is processed in the same
manner as its allochthonous counterpart. Aquatic microphytcs may bl"
consumed alive by herbivorous grazers, such as limpets (Ancylw spp.) and
certain stonefly nymphs (e.g. Brachypttra spp.), or they may be filtered or
gathered with other FPOM by the collectors.
A fraction of CPOM is quickly lost to the DOM compartment by
leaching. The remainder is converted by three processes to FPOM. First,
mechanical disruption by battering and abrasion causes comminution or
the particles. Secondly, processing by microorganisms causes gradual
break-up of particles. Thirdly, the invertebrate shredders fragment the
large particles and produce large quantities of faeces which comprise a
component of the FPOM compartment. DOM is also converted into
FPOM, either by a physical process of flocculation or by microbial
assimilation - the microbes being, by definition, fine particles. FPOM
constitutes the food supply for invertebrate collectors. At the top of the
trophic structure predators prey on grazers, shredders and collectors, as
well as each other.
Several arrows have been omitted from the model shown in Fig. 4-1 for
the sake of clarity. All the heterotrophic categories contribute to the
§ 4.1 GENERAL MODEL OF LOTIC COMMUNITY ENERGY FLOW 39
FPOM compartment (and 10 some extent 10 CPOM) in the fonn of faeces,
exuviae and dead bodies. In addition, a small fraction of che
alloch1honous input contributes directly co FPOM. The principal food of
many fish in lo1ic environments consists of invertebrates so the
'predators' category includes fish. However, some feed on macrophyces,
microphytes or detritus.
The various aspects of community energetics which are outlined in Fig.
4-1 will be discussed in the following sections.
4.1.1 Coloniz.alion and conditioning of organic malter by microorgani1ms
Dead organic mauer, whether allochchonous or aucochchonous in
origin, passes through a progression of stages during its decomposition.
Firstly, soluble organic compounds such as sugars and amino acids are
quickly leached out. For example, tree leaves lose between 5 and 27% of
dry weight, depending on species, mostly in the first 24 hours after
entering che water. Within a week, large populations of fungi develop,
particularly the aquatic hyphomycetes such as Clavariop1u aquatica and
other fungi including Phytophthora and Pylhium. Bacteria are always
prese111 in association with the fungi but evidence about their relative
roles in decomposition of large particles is conflicting. The details of
microbial breakdown of organic mauer are notoriously difficult to
elucidate and this is an area where a large research effort is required in the
future. One thing that is clear, however, is that bacteria, in contrast to
fungi, are able to participate in decomposition of very small particles - a
single auached algal or phytoplankton cell can support several bacteria.
h is also likely that bacteria, whether free-living or associated with the
organic matrix auached to the surface of stream substrates, are able to
absorb dissolved substances such as those leached from dead material or
leaked from living cells.
The different parts of tree leaves decompose at varying rates. For
example, midrib, veins and petioles persist much longer than the softer
tissue, producing characteristic skeletonized leaves. The various chemical
components also disappear at different rates. This has been elegantly
illustrated by work on leaves of white oak, Q.uercw alba, in Augusta creek
in the U.S.A. Leaf-packs were constructed from air-dried leaves collected
just prior to abscission. The leaves were weighed into 5 g amounts,
moistened, fastened together with plastic buttons, tied with nylon thread
to a brick, and positioned with the top leaf facing into the current.
Standard packs were left in the stream over the course of eight months.
Packs were removed periodically for chemical analysis. Figures 4-u and
b show clearly how most of the soluble phenolic compounds and
carbohydrates were quickly leached out. The resistant, structural
cellulose and hemicellulose components disappeared much more slowly,
lipids at an intermediate rate (Figs 4-11c, d and e). Total nitrogen (Fig.
4-2fl actually increased initially, a trend also observed in other studies
and generally attributed to the microbial biomass incorporating nitrogen
40 GENERAL MODEL OF LOTIC COMMUNITY ENERGY FLOW § 4. I
from the water. At the end of the eight-month period only about 1696 of
initial dry weight remained. II is worth noting that allochthonous
material may be washed or blown into the stream long after its death, in
which case it will have been subjected to a period of leaching and
decomposition in the terrestrial environment. This will have modified its
chemistry and therefore its rate of decomposition in the su·eam.

800 la) Phenolic compounds 800 lb) Carbohydrates

400

0L-!:~:t:=::!::::::;t;::==tt::::~===-====;!!
0 8 16 24 32 8 16 24 32
.f.!
u
ea
a.
.,!.
ea
.!!
~
ea 800 Id) Hemicellulose
0

-~
i 400
.!:
C>
·c:C: 0 -=-o--±9---="15=---"""2""4,---~32
"iii 32
E
I!!
C>
E
60 (f) Total nitrogen

o.,,.__ __,,____,,..,,._----=..,____
0 8 16 24 32
Weeks
Fig. 4-2 Weights of various chemical components remaining in white oak leaf-
packs at various times during decomposition. (After SUBERKROPP, K., GODSHALK.
G. L. and KLUG,M.J. ( 1976). Ecology, 57, 720-7.)
Since chemical make-up is so influential in decomposition, inter-
specific differences in rates of disappearance are to be expected. At one
extreme, dead aquatic macrophytes are probably among the fastest to
decay while twigs may take several years. Species-specific rates of decay
have been noted for tree leaves with, for example, ash and alder
disappearing fastest (in excess of 1.5% per day), sycamore, maple and
willow at an intermediate rate, and beech and oak slowest (less than 0.596
perday:PETERSENandcuMMINS, 1974).
§ 4.1 GENERAL MODEL OF LOTIC COMMUNITY ENERGY FLOW 41

4.1. 2 The role of invertebraleJ


Microorganisms arc able to carry ou1 1hc 1ransfonna1ion of CPOM w
FPOM in isola1ion, bUl normally invertcbrales also play a role. The
influence of de1ri1ivorous shredders has been inves1igatcd by monitoring
weight loss of slandard leaf-packs in experimental s1rcams wi1h and
without 1he invertebralcs. Table 4 shows 1hat pignut hickory leaves lost
significamly more material in the presence of 1he shredder populations
(crancAy larvae, Tipula sp.; cased caddis larvae, PycnopJJche sp.; and
stoncAy nymphs, PleronarcyJ sp.). However, the invertebrates had a
negligible effect on white oak leaves during the six-day experiment.
Certainly, invertebrates invariably play some role both in comminuting
and consuming parts of large particles, but the extent of their influence
depends on the origin and composition of the organic maner.

Table 4 The influence of shredders on leaf weight loss


during a 92-day period in an experimental stream.
(After PETERSEN and CUMMINS, 1974.)

Perctntagt dry weight loH


without with
shredders shredders

Pignut hickory 33.04


White oak 21.58

The shredding role is important not only in providing larger surface


area for microbial activity and perhaps in preventing microbial
populations from becoming moribund, but also in contributing to
FPOM and thus making food available to the 'collectors'. This has been
demonstrated in an experiment using alder (Alnw rubra) leaves labelled
with a radioactive isotope of phosphorus (52 P) in an artificial stream. In
the presence of a shredder (the stonefly PteronarcyJ calijomica) two collector
species (larvae of the net-spinning caddis HydropJycht califomica and the
blackfly Simulium arclicum) were shown to accumulate a significantly larger
amount of 52 P in their bodies than when the shredder was absent (SHORT
and MASLIN, 1977).
There are two sub-categories of collector: the colleclor-galhererJ -
including many mayAy species, oligochaete worms and orthoclad
chironomids - inhabit areas where FPOM is abundant on the stream bed
and simply ingest whole particles; the collector-filterers, such as those used
by Short and Maslin (above), sieve FPOM from the flowing column of
water. The filtering technique is made possible by the flowing nature of
lotic systems and a great diversity of adaptations take advantage of this
feedingopportunity(HYNES, 1970).
4:l GESERAL ~tODEL OF LOTIC co~o-n·:--m· E:--ERG\" ROW § 4.1
./. r. J RtlaJit·r rolrs of microorgani.,m..• and ,i(·rn·rw in im·rrtd•r,1/r dirt
Of course. microorganisms continue 10 metaboli1.e dt·trital matt-rial
after it has entered the FPOM ca1ego~- and then the- rdationship lw1wcc11
dead organic mauer. microorganisms and dt·tdti\'tll'ous in\'crtdll",llt's is
at its most intimate.
The matedal consumed b" detriti\'ores i1wariabh· consists of two
componems - dead organic ma11er and mio-oorganist~lS. :\due{() their
relative importance in diet has been pro\'ided by studies on 1hc dlicicnn
with which ingested pure fungus or unconditioned tn~e leaf is ;1ssimila1ed.
GammantJ pJ.eudolimnneuJ, for example, assimilated onlv 17-19'.\; of the
energy contelll of discs of elm or maple leaf. the remaindn being lost in
the faeces. On the other hand, animals provided with pure cultures ol
aquatic fungi which are commonly associated with tree ka\'es wert· able 10
assimilate 74-8396 of their energy content. (BARLOCHER and HNDR1ct.,
197 5bl. Such results have led CUMMINS ( 1974) 10 liken 1hc microbial I issue,
to 'peanut butter' - a nutritious food, which occurs on a nutritionall\'
unsuitable de1ri1al 'cracker biJcuit' ! There is no doub1 tha1 a largl'
proportion of the diet of both shredders and collectors dt•rives from
microorganisms, and there is considerable evidence that de1ri1i\'on·,
prefer conditioned material and grow beuer when led on it.
However, the possibility remains that although dead organic ma11n i,
less readily assimilated it may still constitute a significant propon ion ol
energy income if large amounts arc passed through 1he alimcn1aq· Iran.
In other words, the un-nutri1ious 'crackers' may be more imponanl 1ha11
Cummins suggests if the 'peanut buuer' is spread 1hinly on 1hick
'crackers'. Cummins' analogy has one further shortcoming. 11 app<'ars
that the microorganisms associated with organic matter do 1101 simplv
provide a highly nutritious food supply, but their activity also rcnckrs dH·
detritus more suitable as food in its own right. GammaruJ. pJeudoli11mnew.
when given a choice between untreated sterile maple (Acer .1nccharum) lea!
discs and others which had been exposed to fungal excretions and
secretions, showed a significant preference for the latter even though no
fungi were actually present (BARLOCHER and KENDRICK, 1975a). It appears
1ha1 'fungal catalysis' can increase palatability by partly decomposing
dead material into subunits digestible by detritivores.
In the natural environment, the detritus-microbe complex is
assimilated by invertebrates with an efficiency of perhaps 4096. The
remainder is voided as faecal pellets which themselves cons1itute a
substrate for microbial colonization and processing. Where dense
invertebrate populations occur, much of the organic sediment will consist
of faeces with their associated microorganisms, and a continuous cycle
can be envisaged of ingestion and production of new faecal substrates.
The organisms are reproducing their own 'crackers' to be spread with
'peanut butter' time and again, but with a continuous topping up of dead
plant matter.
Much of the work on the microorganism/detritus/detritivore inter-
44 LONGITUDINAL PATTERN IN COMMUNITY ENERGY FLOW

The auached algal community would be expected to achieve its


maximum role nearer the headwaters than macrophytes, because many
species of the former are beuer able to withstand a more extreme flow
regime and, moreover, do not require a sediment of fine particles for
rooting. Proceeding downst ream, both attached algae and macrophytes
should continue to provide a significant energy base for the community
until the depth and turbidity of the water increase to the extent that
insufficient light reaches the bouom, when growth will be largely
restricted to the margins. Finally, in longer rivers phytoplankton should
contribute significantly to the community's energy base. It can be argued
that in undisturbed rivers an appreciable phytoplankton component will
only develop if the time taken for water to move from source to mouth i~
sufficiently long, otherwise populations could not bu ild up to any great
extent. While diatoms can often be taken in the water column in upstream
reaches these are derived from the attached algal community and do nor
constitute a true phytoplankton (whose complete life cycle takes place in
the water column).
Few studies have been made of total community functioning in loric
systems. Two quite thorough investigations which provide a useful
contrast are those on a small stream, Bear Brook, U.S.A., and on the Rivl'r
Thames in England , which on a world scale would be classed as a small -
to-medium-sized river.

4. 2. r The Bear Brook study


Bear Brook is a small stream running down a hillside in the foresrcd
Hubbard Brook catchment area of New Hampshire (see § 2.2 . 4). The
forest provided heavy shading so that no auached algae and f<.·w
macrophytes could survive. The only autochthonous production wa s
provided by mosses which contributed a bare 0.2% of the total energv
input. The allochthonous inputs were determined by analysis of water
samples and by trapping litter as it entered or Aowed down the stream . All
values were converted into energy equivalents and are shown in Fig. 4-4 .
Allochthonous matter was provided from three sources : 44% as leaves,

EXPORT M 111SO
Pan,cvl11e 5040
GROUNDWATER

AU.(ICKT-,sl

- -
•ROMUPSTREAM

•ORESTCAHOr<

Fig. 4-4 Diagrammatic representation of energy How in Bear Brook. All figures
arc in kj m- 1 yr• . (After FISHER, s. c. and LIKENS, c. E. ( 1973). Ecological monographs,
45, 4u-59.)
LONGITUDINAL PATTERN IN COMMUNITY ENERGY FLOW 45
fruits , twigs and branches from rhe forc·st canopy which fell or blew into
the stream, 2596 as dissolved organic mauer in groundwater drainage into
the srream, and the remaining 3196 in the Aowing water column from
upsrream (2296 dissolved and 996 particulate>. The total annual input of
energy to the stream was 25 360 kJ m- 2 •
Approximately 19 900 kJ m- 2 , almost equally divided between leaves
and branches. was stored within the 1700m study section. This detritus
reservoir did not vary much from year to yea~ and the system was
considered to be in a steady state. As would be expected. the rate of
decomposition of leaves was faster than that of branches, with turnover
times of I yr and 4.2 yrs respectively .
No detailed studies were performed on the heterotrophic components
of the community, but by calculating the difference between total energy
inpur and energy exported downstream as dissolved or particulate matter
ii was found that the heterotrophs accounted for 3496, with micro-
organism metabolism predominating.
The results of 1his study tally well with the hypothetical model shown in
Fig. 4-3. The system relies almost entirely on allochthonous matter as an
energy base. However, a similar study on New Hope Creek in North
Carolina revealed a peak of auached algal production on the stream bed
in early spring, before the forest canopy had developed (HALL. J. w., 1972,
quoted in MANN, 1975). In this case, the total yearly input ofleafmaterial
was estimared as 8800 kJ m- 2 yr-•, while another' 4-8000 was provided by
autochthonous production. Further research is required to determine
whether such seasonal switches in productivity are of general significance.

4.2.2 The Riuer Thames study


Much of 1he energy base in the River Thames, near Reading, was
produced autochthonously. in contrast to the situation in Bear Brook
(Fig. 4-5). Macrophyte production was restricted to the shallow margins
and accounted for only 185 kJ m- 2 yr-• . Attached algae were observed on

EXPORT

AllOCH-
TMONOUS
TREES

....,._,on 1uoo
{
Mtetopltyle1

An..-•- _ __:_:_--t,{]~~J--

Fig. 4-5 Diagrammatic representation of encrgy_flow in the River Thames, near


Reading. All figures are in kj m- 1 yr-•. (After MANN, 1975.)
46 LONGITUDINAL PATTERN IN COMMUNITI ENERGY FLOW § 4.2

the river bed and in the diets of certain invertebrates and fish. Production
by auached algae was not measured but was assumed 10 be small in
comparison to that of phytoplankton, which was 18 400 kJ nr 2 yr- 1 • This
assumption is probably correct. A study of another medium-sized river,
the Ouawa River in Canada, revealed that although the standing crop
biomass of attached algae was larger than 1ha1 of phytoplankton the
annual production of the lauerwas more than 2.5 times as great.
The alloch1honous input 10 the Thames was derived from the same
sources as Bear Brook, but only the tree leaves were moniwred. These
contributed only 330 kJ m- 2 yr 1 • The input of dissolved and FPOM from
upstream will certainly have been much more substantial than the
au1och1honous component. Some indication of their relative role~ is
given by the fact 1ha1 filter-feeders took approximately 2 .5 times as much
dead particulate mauer as phytoplankton cells. As was the case for Bear
Brook, however, the majority of the imported material will have been
exported downstream having played no part in community metabolism.
A certain amount of particulate organic mauer was stored on the river
bed and together with the auached algae this cons1i1u1ed an imponam
source of food for collector-gatherers such as 1ubificid worms and
gastropods. Unionid mussels and sponges were the dominant bemhir
collector-filterers, consuming large quantities of suspended FPOM and
phytoplankton, while filter-feeding zooplankwn were common among
vegetation and between pebbles on the river bed. Young fish derived most
of their food from zooplankwn, older ones from benthic detritus and
invertebrates. Pike and large perch were piscivorous.
This investigation of a downstream station where the river is wide.
quite deep and turbid, supports some general points in the model shown
in Fig. 4-3. For example, it has been shown 1ha1 the amotrophic
component of energy supply is relatively more important, that 1he
macrophytes are restricted 10 the river margins and that phytoplankton
assume significance. However, the lauer point needs 10 be in1erpre1ecl
with caution. The Thames has been greatly influenced by the ac1ivi1ies of
man, and it appears that the high population densities of phytoplankwn
are promoted by the presence of weirs and locks every few miles (w
maintain a constant depth for navigation). Thus, water has a sufficiently
long retention time for populations to build up. Retention time is also the
significant factor in relatively undisturbed rivers. Water in small streams
such as the Rheidol in Wales (48 km long) takes less than a day to run its
course. The average run of water for the Thames ( 240 km) is six or seven
days, while a long river such as the Volga in the U .S.S. R. before dams were
built in 1928 took 50 days to run its 2747 km course when in Hood, much
longer in summer. Since the generation time of many zooplankton is
three or four days, shorter for phytoplankton, it is hardly surprising that
short rivers, and the headwaters of all, are devoid of plankton. In general,
some planktonic organisms appear a few days from a river's source while
an abundant phytoplankton probably takes 20 days or so to develop. The
LONGITUDINAL PATTERN IN COMMUNITY ENERGY FLOW 47
plankton in longer rivers probably originate with innocula injected from
backwaters, adjacent lakes and swampy areas. The populations may pass
through several generations as they arc swept downstream, but the
majority must eventually perish in the sea.
These studies, and others like them, lend rnppon 10 the model of
longitudinal changes in community energy Row, but further work is
required.

4.2.J The ratio of CPOlH to FPOM


It is not only the sour-ce of the energy base that varies in a downstream
direction. The ratio of coarse to fine particulate matter would also be
predicted to change. CPOM is always a dominant input to forested
headwaters but its input per square metre of river surface diminishes
downstream. The CPOM entering headwaters is convened by both
abiotic and biotic processes to FPOM, and thus the latter should be
significantly more important downstream.
Such a fundamental difference in the form of available particulate
organic matter would be expected to be reAected in the composition of
the benthic invertebrate community, with shredders making up a larger
proportion upstream. This prediction was confirmed in a thorough study
of the trophic status of cad dis larvae in rivers running through the eastern
deciduous forests of North America. Figure 4-6 shows that shredder
genera predominated in upstream habitats, but that downstream the
proportion of shredders became markedly smaller in relation to
collectors. This change is to be expected because collectors consume

0-5 10 50 500
Width of stream (ml
Fig. 4-6 Pie diagrams showing the relative proportions of various categories of
caddis Ry genera in upstream and downstream sections of North American
streams. S - shredder, C-collector, P- predator, C -grazer. (After WIGGINS. c. e.
and MACKAY, R.J. (1978). Ecology, 59, 1211-110.) This is a test of the stream con-
tinuum concept ofK. w. CUMMINS.
48 CONCLUSIONS § 4.3
FPOM. Interestingly, shredders also decreased in relation 10 grazers ancl
1his reAec1s 1he increased production of auached algae in 1he unshaded
downstream reaches. The proportion of predatory genera did 1101 v,11)'-

4.3 Conclusions
In 1his chapter, downstream changes in several environmental fanors
have been used to explain some fundamental variations in community
energy How. The key factors were amount and form of alloch1honous
input, shading from above as influenced by river wid1h, shading
within 1he wa1er column as influenced by dep1h and 1urbidil)', flow
characteristics and 'age' of the water mass. Somt' of 1hese variables ,1lso
figured prominently in the discussionoflongitudinal dis1ribu1ion pauerns
of individual species, and, of course, this is 1101 surprising. After all, 1he
community is the sum of its parts, and facwrs 1ha1 influence dis1ribu1ion
of the pans also affect 1he whole. However, 1he energetics approach
provides a different perspective and throws new light on 1he dis1ribu1ion
of individual species. For instance, among the net-spinning hydropsychicl
caddis larvae, species in 1he subfamily Arctopsychinae spin ne1s wi1h a
coarse mesh, those of the Hydropsychinae spin intermediate mesh ancl
the Macronematinae the finest of all. The usual downstream dis1ribu1ion
of North American species is shown in Fig. 4-7. The coarse mesh of thl'
Arctopsychinae permits them to deal with the abundant CPOM (,1s
well as animal prey) which occurs in mountain streams, whl'reas lhl'
Hydropsychinae and 1he Macronematinae can fiher the minute FPOM in
the downstream reaches they inhabit.
The energetics approach also provides an insight imo a striking
life-history phenomenon observed in cool headwater streams. The·
detritivorous invertebrates are able to metabolize al temperaiures below
the normal optimum and this is seen as an evolutionary response to lhl'

Arctopsychinae .__ _ _ __,

Hydropsychinae

Macronematinae

Headwaters Downstream

Fig. 4-'7 Typical downstream distribution of three North American subfamilies


ofhydropsychid caddis.
COSCLL:SIONS 49

major autumnal input or particulate organic mauer. In addition, it is not


uncommon to find life cycles adjusted to synchronize with the autumnal
input of leaves (for a full treatment of life-history phenomena the reader
is referred to IIYNES ( 1970) Chapter 14).
Generalized schemes such as 1ha1 described in Fig. 4-3 alwavs have
numerous exceptions and these need to be pointed ou1. Some stre;ms rise
above the tree-line on mountains. These are comparatively unproductive
with most of their energy input deriving from anached algae and
mosses. Sometimes, man exerts an effect on the headwater environment
and streams running through deforested catchments arc also likely to
derive much of their energy from autochthonous sources. The major
influence of human activity, however, occurs further downstream where
the centres of population are to be found and where the requirements
associated with navigation, drinking water and sewage disposal are
paramount. Dams, lock gates and reservoirs increase the retention time of
water (producing what amounts to a string of lentic environments) and
thus the importance of phytoplankton. Sewage input constitutes a
significam addition 10 dissolved and particulate organic loads.
Finally. it is necessary to stress that the conceptual model presented
here is concerned with gualitative changes that occur down the length of
rivers. I I does not seek to predict which species of plants and animals will
be found, since this often depends on factors such as nutrient
concentration and climate which vary from river to river and from one
geographical location to another.
5 Microdistribution Patterns

5. 1 Introduction
Habitats in streams and rivers are not usually homogeneous. Local
heterogeneity exists in mineral particle size, the occurrence of detritus, in
local patterns of Aow velocity and in depth. This is reflected in the
distributions of benthic organisms which are, as a rule, highly clumped.
In other words, individuals of a given species tend to be over-represented
in certain microhabitat patches and under-represented or absent in
others. Most microdistribution studies have been performed in shallow
streams, which are relatively easy to study, and have concerned benthic
invertebrates.
Although the primary concern is with horizontal microdistribution on
the surface of the stream bed, recent investigations have shown that a
proportion of the populations of benthic invertebrates may be found
deep within the substratum. Work on vertical microdistribution will be
discussed first.

5.2 Vertical microdistribution in the substratum


The evidence from several recent studies, mainly by H. B. N. Hynes and
co-workers, has demonstrated that a large number of individuals of many
species occur quite deep within the bed of stony streams. These animals,
which live in the interstices, are termed collectively the hyporheoJ.
In an intensive study of the fauna of the Speed River in Canada,
monthly samples were taken with a core sampler at 1 o cm intervals down
to a depth of 70cm. Invertebrates rarely occurred at greater depth. Figure
5-1 shows the mean percentage depth distribution for all taxa combined.
The pattern is clear with maximum density just below the surface, and
decreasing with increasing depth. Density at the surface was estimated by
a different technique, 'kick' sampling within a roughly delineated area
into a net held downstream, and this unsophisticated approach probably
underestimated density by a factor of about 5. Even allowing for this,
however, no more than 1 796 of the total number of invertebrates occurred
on the surface.
The hyporheos can be subdivided into 'occasional' and 'permanent'
categories. The latter refers to forms such as micro-crustaceans and mites
which complete their whole life cycle within the substratum. Of the
insects, a few (e.g. the caddis larvae of HydropJyche and Rhyacophila) were
found exclusively on the surface. Most, however, were occasional
residents in the hyporheos and sought out this habitat during part of their
life cycle. Several taxa, including larvae of chironomids in the genus
MICRODISTRIBUTION ON THF. SUBSTRATUM SURFACE •
,') I

Cladotanytanus, overwintered al depth and migrated upwards in spring


when the surface water began w warm up. WILLIAMS and HYNES ( 1974)
postulated that these animals moved down in winter lO follow an optimal
temperature for growth. On a shorter time scale, there is evidence 1ha11he
fauna of a stream may move deeper inw the substratum during Aood
conditions, presumably gaining some protection from the current.
The discovery that in certain streams a large proportion of the
invertebrate community may lead a hyporheic existence, at least for pan
of the time, needs to be taken carefully into account when designing
sampling programmes.

Percentage of invertebrates
occurring at each depth
10 20 30
Surface

10

20

30
Depth
(cm)
40

50

60

70

Fig. 5-1 Mean percentage depth distribution within the stream bed for all
invertebrate taxa in the Speed River, Canada. (After WILLIAMS. D. D. and HYNES.
H. B.N.(1974). Fmhwal. Biol., 4, 2gg-56.)

5.g Microdistribution on the substratum surface


The identification of factors which influence microdistribution
involves first, description of the microdistribution, in a series of
standard samples, of the organisms being studied; secondly, description
of all the major microhabitat variables in each of the samples; and
thirdly, analysis of the results to discover relationships between these
environmental factors and the species distributions.
.52 MICRODISTRIBUTION ON THE SUBSTRATUM SURFACE § .5.3
~. J. r The i1zjluence of water depth and shadi11g
Depth appears to be of little direct importance in llll' dis1ribu1ion of
benthic invertebrates, but ii does influence macrophy1c pauerns. For
example, 1he leaves of the reed Pltragmites co,mmmis arc unable 10
photosynthesize under water so this species usually grows where al leas1 ,1
third of the shoot is above water. Depth is of indirect importance 10 plants
in 1erms of illumination. In general, pla111s which live in the deepest wa1cr
are those most tolerant of shading (e.g. rigid hornwort, Ceratopltyllum
demersum). Shading may also be caused by 1rees and shrubs which
overhang the water. Species which are shade-tolerant, such as strapwecd,
Sparganium demersum, do well under these conditions. Depth and shading
vary across the width of streams and rivers and this is often rcAcncd in
plant distribution.

5 .J. 2 Flow velocity


Flow velocity has been implicated in several studies as an important
determinant of invertebrate microdistribution. One such investigation
was that of EDINGTON ( 1968) on two species of net-spinning caddis in
streams in northern England. The larva of Hydropsyche siltalm is, for 1he
most part, detritivorous. It spins a streamlined, rigid net wi1h a regular
1"11esh ( 1 70 x 300 µm in the case of final instars) which filters our detrital
particles from the flowing water. Plectrocnemia conspersa larvae are
carnivores, trapping invertebrate prey as they walk or drili on 10 their
irregular and rather flimsy nets. In the study sites there was an alternation
of riffles and pools, as in most streams, and Edington found that H ..iiltalai
predominated in riffles and P. conspersa in pools. Using a miniature flow
meter, he was able to measure water velocity at the mouths of the nets or
the two species. This was facilitated by the fact that both kinds of caddis
larvae spun their nets on the tops of stones and boulders. H. siltalai nets
occurred in velocities between 15 and more than I oo cm sec•, being most
common in the fastest Aows. P. conspersa nets, on the other hand, were
found in velocities between o and 20 cm sec•, 9596 occurring in less than
10 cm sec•. Edington obtained convincing evidence that Aow velociry
was the determining factor by carrying out simple experiments in the
streams. He showed that the Aimsy nets of P. conspersa were unworkable a1
high velocity by rearranging stones upstream to increase the Aow over
established nets. The nets very quickly ruptured. In the laboratory he
demonstrated that 25 cm sec• was enough to destroy a net. The rigid ne1
of JI. siltalai was well adapted to withstand fast Aow and it still functioned
when flow velocity was reduced, although probably providing its
occupant with insufficient food. In low velocities, if the net was destroyed,
JI. siltalai failed to rebuild but moved to an area of higher flow. Finally,
Edington examined in the laboratory the relationship between ne1-
building success and flow velocity (Table 5).
MICRODISTRIBUTION ON THE SUBSTRATUM SURFACE 53
Table 5 Percentage of caddis larvae constructing nets
al different flow velocities.

Velocity <cm sec- 1 )


10 15 110

I lydropJycht Jiltalai
Plectrocnemia comptna

It is clear that flow velocity is a crucial factor for net-spinners, as it


probably is for filter-feeders in general. Further groups for which it must
be particularly significant are those that live exposed on the surface of
stones and others which require a continuous supply of oxygen-rich
water over their respiratory surfaces. For example, the tendency of
nymphs of the mayAy BaeliJ tricaudatus to leave an experimental
substratum tray after a reduction in Aow, has been attributed to their high
oxygen requirements (RABENI and MINSHALL, 19 7 7).

5.J.J Mineral mbJtratum particle Jiz.e


This factor has also been implicated in a number of studies, one of the
most thorough of which was performed on the freshwater amphipod
Gammarw pule.-..: U. H. R. Gee, unpublished). Using a multivariate
statistical analysis, Gee was able to show that factors such as depth and
Aow could not provide a convincing explanation for the clumped
distribution of the population in a stream in southern England. However,
their microdistribution was significantly related to mineral particle size
(as determined from photographs taken of the substratum surface). While
the smaller G. pulex appeared 10 be distributed across the whole range of
mineral particle size, avoiding only silt, the larger individuals were
progressively more restricted to larger particle sizes. Evidently, these
animals need interstices of the appropriate size within which to wedge
themselves and gain a measure of protection from the current and,
perhaps, enemies.

5.J.4 Food
It is 10 be expected that the distribution of animals will reflect the
distribution of their food, and very often this appears to be the underlying
factor. In the case of G. pulex, discussed above in the context of particle
size, the presence ofleaves was influential in explaining microdi.stribution
during the autumn and winter. Many detritivores are associated with leaf-
packs (the 'shredders' in general), while others (the 'collector-gatherers')
are more likely 10 be associated with finer detrital particles in the
substrate. Figure 5-2a gives a typical examele of the relationship for
several taxa between number of animals and-dry weight of fine detrital
particles. Other studies have shown that invertebrate predators often
54 MICRODISTRIBUTION ON THE SUBSTRATUM SURFACE § 5.3
exhibit aggregative responses, that is they are commonest where 1heir prey
arc abundant (Fig. 5-2b, c and d). Net-spinning lan-ae of 1he caddis
Plectrocnemia compersa (discussed in§ 5.3. 2 in the contex1 of flow velocity
preference), provide a good example. Their aggregative response comes
about because of a reduced probability of abandoning a net if prey
are taken regularly ( HILDREw and TOWNSEND, 1980).
Clearly, the presence of food is often the proximate cause of
invertebrate microdistribution patterns. Even in cases in which a fac1or
other than food is the proximate cause food may, nevertheless, be 1he
ul1imate factor. For example, the preference of net-spinning caddis for

"'
,:, Q)

lal ea. (b)


300 g ~400
e"'
'.E ~
a; u Q.
Q.
200 ~ai
"'
ni Q) 0"'
Ea.
·c E ~ ~0 200
-
Q)

"'"' 100 Es~ 0. ~


0 "' -c
0"'
ci - - leuctra inermis ci t:.
z
o.___,,---=-=,--:!:,-..,..-le-:--~ z o,..____~,.,,._,,-___..,,.,
·25 ·50 ·75 1·00 1·25 2000 4000
Plant detritus per sample No. of Chironomidae
(g dry weight) per sample

(cl 8 (di
10
"'"'
"'
E a, 8 6
.2 a.
.o E
g~
-
al"' 8.
ci E 4
6
4

fti ~
O! 2
- Q. 2
0

Prey biomass per sample (mg dry weight)

Fig. 5-1 Microdistribution of invertebrates in relation to food availability. (a)


Detritivores and plant detritus. (After ECCLISHAW. H . J. ( 1969). J. Anim . £col., 38,
19-33.) (b) Predatory chironomids in relation to herbivorous chironomids, their
prey. (After FAHY, E. (1975). Fmhwal. Biol., 5, 167-82.) (cl The alderfly Sia/is
fuliginosa and (d) PltctrOCMmia conspma in relation to chironomids and s!OneAies
combined. (After TOWNSEND, c. R. and HILDREW. A. c. ( 1979). J. Anim. £col., 48,
90~10.)
MICRODISTRIBUTION ON THE SUBSTRATUM SURFACE .5.5
certain flow velocities (§ 5.3.2) reflects the particular requirements for
efficient operation of their food-capturing nets. Furthermore, mineral
substratum preferences may underly a need for specific sizes of detrital
particles. RADEN I and MINSHALL ( 197 7) found that implanted trays of fine
mineral substratum collected smaller detrital particles than trays of
coarse subtratum. They suggested that larger numbers of invertebrates
colonized the former not directly because of mineral particle size but
because of the abundance of preferred detrital food.

5-.J. 5 Enemies
No direct evidence on the influence of predators and competitors is
available from invertebrate microdistribution studies. However, since
prey are likely to be more vulnerable in relatively homogeneous substrata
such as bedrock or silt because these offer fewer refuges from predation
than gravel, stones or leaf-packs, it can be postulated that the relative
densities of prey in various substratum patches will sometimes be
affected. In addition, a role for competitors can be inferred from
experimental work which has demonstrated a higher rate of dispersal
from densely populated patches. This will tend to reduce, to some
extent, the degree of clumping of a population.

5.3.6 Habitat ulection experiments


The discovery that the spatial pattern of a given species correlates
closely with a particular environmental factor should not be taken as
proof that the factor in question is the underlying cause of the pattern.
It is possible that the variation in both microdistribution and en-
vironmental factor might be caused by a second factor, especially as
several environmental factors usually vary simultaneously. A causative
link can be proved only in controlled experiments where one factor is
varied and the others held constant. EDINGTON (1968) showed that his
caddis larvae chose to spin nets in the preferred flow velocities(§ 5.3.2);
Gee showed that Gammarw pulex moved more slowly in mineral substrates
of the preferred size range and thus accumulated in the appropriate
patches (§ 5.3.3); HILDREW and TOWNSEND (1980) demonstrated that
Plectrocnemia compersa were less likely to leave areas where prey were
abundant (§ 5.3.4); and in many other cases, laboratory choice
experiments have confirmed that the correlation between distribution
and a particular environmental factor was due to a behavioural, 'habitat
selection' of appropriate patches of detritus or mineral substrate.
6 Benthic Community Dynamics

6.1 Introduction
The objective of most microclistribution studies has been to describe
the spatial organization of a benthic community at an instant. prnduring
a record analogous to a still photograph. The impression is that the
pattern so described is unchanging, and that its structure can be
explained simply in terms of the ability of organisms IO select appropriate
microhabitats. In fact, the assemblages arc in a continuous state or nux as
a result of shifting patterns of now and substratum and because or the
organisms' own activities. The results of a microdistribution study can be
likened to an action photograph, such as that or footballers 011 a
newspaper sport's page. To fully understand the dynamic 11;11u1-c ol the
system it is necessary to emulate a cine film analysis and disc-ovn the
movements which led to the configuration fixed in the still phowgraph.
as well as subsequent events.
When a large amount of rain or snow-melt results in increased flow the
shear stress exerted on the stream bed intensifies and the substratum is
scoured. Sanely beds and other readily erodable formations may lose
material to substantial depths, and organisms on and in the subsu·,11um
will also be removed. As the Aood recedes and shca1· stresses diminish,
material and organisms from upstream will be deposited. The mmTmclll
of invertebrates associated with flood conditions is termed cala.1trophic drift
and may involve the displacement of a large proportion of the l)('nthic
community.
Although during long periods of lower discharge the substratum is
relatively stable, many benthic organisms arc still on the mm·c·, either
along the stream bed or in the flowing column of watei-. If a net is placed
in a stream or river, facing upstream and suspended above the bottom, it
will rapidly collect an appreciable number ofbenthic invertebrates which
have lost contact with the substratum and arc being carried along in the
current. The vast majority are undamaged and would quickly have
regained a foothold had they not been trapped in the net. This
phenomenon is called invertebrate drift. ll occurs even in very slight flow
and is of major importance in the continuous redistribution of benthic
organisms from one patch to another.
Upstream, downstream and random movements in contact with the
stream bed, as well as movements through the substratum, add to the
complicated picture ofbenthic community dynamics.

6.2 Experiments on temporal changes in microdistribution


Whenever an area of substratum has been denuded of organisms,
whether through catastrophic drift or as an experimental procedure, it is
TEMPORAL CHANGES 1:,,; MICRODISTRIBLTION .57
quickly recolonized. The time taken to reac·h the characteristic density of
organisms is often only one or rwo weeks. This testifies 10 rhe large scale of
movement of organisms thar occurs all the time.
WILLIAMS and HYNES ( 1976) carried out an experiment in a slowl~·
flowing Canadian stream to determine rhe relative roles of various
categories of movement in coloni,.ation of implanted rrays of gravel and
rocks. Four major sources of colonists were assumed: invertebrate drifr,
upstream movement in contact with the bottom, upward vertical
movement from deep in the stream bed, and coloni,.ation from aerial
sources by oviposition. Four types of trap were designed to isolate each of
the fauna I fractions contributing to colonization. These are illustrated in
Fig. 6-1. Trap (al was constructed to prevent access to the6ox 30 cm tray of
substratum except by invertebrate drift. It consisted of a wooden box with
only the upstream end open. A perspex cover prevented ovipositing
insects entering from above. In fact, since this trap was placed on the
stream bed it must have collected animals moving downstream along the
bouom, as well as in the drift. Trap (b) was open at the downstream end. It
allowed only upstream moving invertebrates to colonize, having a mesh
screen in the upstream end. Trap (c) was screened at both ends but allowed
organisms to move vertically into the colonization tray from the
substratum below, while trap (d) was a floating trap with access only from
above. Its sole colonist source was thus ovipositing insects.
Two of each kind of trap were placed in the stream in June and left for
28 clays. Two controls consisted of the standard wooden frame but with
access possible from all sources. During the experimental period, the
control trays were colonized by an average of 34000 invertebrates,
principally chironomid larvae. The percentage contribution of each
source of colonists is shown in Fig. 6-1. All four sources were important,
at least in June, with drift contributing the largest number.
Williams and Hynes did not completely isolate drift as a colonist
source. This aspect was investigated by TOWNSEND and HILDREwb976), who
introduced two sets of trays to a sluggish stream in southern England, in
summer. One group of the trays (37 x 17.5 x 2.5cm) was placed on the
stream bed so that colonizing animals derived both from those in the drift
and those walking across the bottom. The other group was placed on
platforms suspended 5 cm above the bottom. These could only be
colonized by drifting organisms. At three-day intervals four 'bottom'
trays and four 'suspended' trays were selected at random, carefully
removed, and the invertebrates they contained identified and counted.
Figure 6-2 shows the pattern of colonization on the two kinds of tray
during the 12-day experiment. On each of the sampling occasions more
individuals had colonized bottom trays than suspended trays. This is
explained by the fact that the former received both drifting and walking
colonists. However, the average ratio of total numbers on the suspended
trays to numbers on the bottom trays was 0.81. Thus, more than 8096 of
the invertebrates moving from one area of stream bed to another
Perspex
cover

(d)
Aerial
28·2%

(b)

Fig. 6-1 Pie diagram showing percentage contribution to colonization of sub-stratum trays by (a) drift, (b) benthic
upstream migration, (cl vertical upward migration, (d) aerial sources. (After WILLIAMS and HYNES,,1976. )
INVERTEBRATE DRIFT 59
probably did so in the drift. The rates of accumulation of species were
similar on the two kinds of tray, demonstrating the overwhelming
importance of drift in colonization. Representatives of most species
arrived within three days.
It is clear from these experiments that invertebrate drift is of major
significance in benthic redistribution. The next section considers this
phenomenon in more detail.

,., (bi

.,, 60
s 10
I!
.0
CD 40 ftl 8
t: )(
CD
> !! 6

-
.!:
0
c:i
z
20
0
z
c:i 4
2

Day number
Fig. ~11 Colonization curves of(a) total individuals and (b) species, on bottom
trays (solid line) and suspended trays (dashed line). AfterTOWNSEND and HILDREW.
1976.)

6.3 Invertebrate drift


The phenomenon of invertebrate drift was discovered accidentally in
the 1920s during the course of studies primarily concerned with the drift
of terrestrial insects that fell onto the stream surface, a source offood for
fish. Since then, a substantial amount of research effort has been expended
on its investigation (HYNES, 1970; WATERS, 19711; MOLLER, 1974). The inci-
dence of non-catastrophic drift is universal in streams and rivers, and
may occur on a very large scale.
The a.mount of drifting varies considerably. Seasonal variations, with
larger numbers drifting in spring and summer, often simply reflect the
regular annual variation in density of the benthos. Another kind of
seasonal variation is demonstrated by certain species of blackffies,
mayRies and stoneflies, whose drift rate increases shortly before pupation
or emergence of the adults. This is probably correlated with a pre-
emergence change in behaviour which renders the individuals more
susceptible to dislodgement by the current. Conversely, small instars are
often commoner in the drift than larger animals.
6..J .1 Diurnal ptriodicily of drift
The most marked pattern of variation in numbers of drifting in-
vertebrates occurs over the 114-hour period of day and night. A large
60 INVERTEBRATE DRIFT

number of studies from different parts of the world have shown that the
amount of drift increases at night, and particularly during the period
soon after sunset. This is the case for a wide variety of taxa. Such pauerns
occur because of periodic changes in behaviour of the animals which
render them more susceptible to dislodgement. For instance, many sit
immobile during the day, sheltered within the substratum and under
stones, but emerge to forage at night. Presumably, this has the advamage
that it reduces the likelihood of being eaten by predators which rely on
vision . Several taxa, however, regularly exhibit higher drift rates during
daylight. Day-active patterns may in some cases result from a direct
water-temperature/activity relationship.
Several environmental factors exhibit diurnal periodicities and could
thus act to govern the daily pattern of drift. Obviously, incidem light
and temperature vary during day and night. However, pH and the
concentrations of oxygen, carbon dioxide and silicon have also been
shown to vary, in relation to the daily course of photosynthesis and algal
cell division . By carrying out experiments in which all factors were held
constant except light, HOLT and WATERS ( 1967 l were able to show that drift
was depressed during the night by keeping the lights on , and that
artificially advancing nightfall caused the drift peak to appear earlier ( Fig.
6-3). It is clear that light is usually the governing factor. Indeed. even
bright moonlight has been found to depress drift .
The solar eclipse witnessed in Australia on 113 October 1976 provided
an unusual opportunity to investigate the inlluence of light on drift. The

OPEN STREAM CONTROL-NATURAL LIGHT


(bi
0 ·2

0·1

01:::::::::::::::~-----
ENCLOSURE-EXPERIMENTAL LIGHT
·04

0·2

0 ½2!!1'!4~-oo~,.2-f4.~oo'JP--,2:.-:4~.oo~-=:24~.ll:1oo~H~

Fig. 6-s (a) The influence of continuous lighting on the drift race of Gammarus
pstudolim11111w. (b) The influence of experimentally advancing dusk on the drift
rate of Battis vagans. At the head of each graph the shaded portions represent
periods of darkness. (After HOLT and WATERS, 1967.)
§ 6.3 INVERTEBRATE DRIFT 61

eclipse was total for only three minutes commencing at 16.42 hours. The
periods of rapidly decreasing and increasing light intensity, 'dawn' and
'dusk', each lasted for 15 minutes. Hourly drift samples were taken in a
small stream on u, 23 and 24 October. Several taxa were found 10
be more abundant in the drift after sunset and these also exhibited
an eclipse-related increase in abundance !Fig. 6-4). Cricotopw
(Chironomidae) larvae were less abundant in the drift after sunset and
also exhibited an eclipse-related decrease in abundance.

LEPTOCERID CADOIS LARVAE HELODID BEETLE LARVAE CRICOTOPUS FLY LARVAE

Tome

Fig. 6-4 Occurrence in the drift on (a) u (b) 23 and (cl 24 October 1976 of
each of three taxa. A11he base of each histogram the black horizontal bar indicates
darkness. The solar eclipse on 23 October lasted for approximately gg minutes.
(After CADWALLADER, P. L. and EDEN, A. K. (1977). Awt.j. Mar. Frtshwaltr Rts., 28,
799-805.)

6.J. 2 Scale of movement in the drift


In order to appreciate fully the influence of invertebrate drift on
benthic community dynamics it is necessary to pose two questions about
the scale of movement involved: what proportion of benthic in-
vertebrates drift per unit time, and how far on average does each move?
In 1965J. M. Elliott (quoted in WATERS, 1972) devised an index express-
ing the percentage of the bottom fauna drifting over a unit area of stream
bed at an instant in time. The majority of values, calculated fora variety of
streams, have ranged from 0.01 to 0.5%. Although these appear low, they
indicate that many times the number of organisms present on an area of
substratum may drift over it in a day. During the colonization experiment
described in § 6.2, TOWNSEND and HILDREW (1976) also sampled the
undisturbed benthos. This allowed the number of drifting organisms
colonizing an area of substratum to be expressed as a percentage of the
total present on the same area of stream bed. The results showed that an
average of 2 .6% ofbenthic invertebrates shifted their position each day by
62 INVERTEBRATE DRIFT

entering the drift and regaining a foothold somewhere else. Some species,
notably those whose food resources were patchily distributed, drifted
more than the average, with, for example, a value of 1496 for the predatm-y
net-spinning caddis, Plectrocnemia cOT1Jpma. Other species drifted at much
lower rates, or indeed not at all, for example the pea mussel I, PiJidium sp.
Mel.AV (1970) attempted to discover how far individuals move while in
the drift. He did this by means of experimental disturbances of the stream
bottom to release organisms into the drift at increasing distances from a
drift sampling net. On the basis of his results (Fig. 6-5) it is clear that 60%
of organisms drift less than 10 m. However, this may have been an
underestimate because some animals may have been damaged or killed as
a result of their drastic treatment and may have drifted further than usual.
The consensus from several field studies is that the majority of drifting
organisms move less than 2 m at a time. Many regain the bottom
accidentally, when brought into contact with it, while others exhibit active
movements which reduce the period that they remain in the drift.

50

.., 40
'
E
d
Z 30
~
·;;;
~ 20
-0
:;;
0 10

Distance from sampler (ml

Fig. 6-5 Effect on drift density of disturbance of bottom fauna at increasing


distances from the sampler. (AfterMcLAY, 1970.)

6 ..J ..J The adaptive Jign.i..ficance of entering the drift


Recently, attention has been focused on the hypothesis that animals
do not all enter the drift as a result of accidental and undesirable
dislodgement. In some cases, invertebrates appear to 'use' the drift as a
means ofleaving an unsuitable patch.
Patches may be unfavourable because they do not possess acceptable
abiotic characteristics. Laboratory studies on the mayfly Ephemerella
Jubvaria (Fig. ~). for example, have shown that the proportion of
animals drifting is significantly higher when the water is turbid or when
20

g' 15 15
:i::
-0
Gi 10 10
.0
E
:::,
I
''
z 5 5 ''
0
'-/
'
0 3 6 0
Time (hrs)
(a) (bl
Fig. 6-6 Distribution of total numbers of Ephtmerella mbvaria nymphs drifting
in a darkened stream (a) at basal (dashed line) and increased (solid line) Row
velocity and (b) under clear Rowing (dashed line) and high sediment (solid line)
conditiom. (After CIBOROWSKI, J. J. H., POINTING, P. J. and CORKUM. L. D. ( 1977).
Fmhwat. Biol., 7, 567-72.)

flow velocity is too high. (Whether the latter results from an active
response or passive dislodgement is unclear.)
Alternatively, patches may not possess acceptable biotic characteristics.
The presence of appropriate food resources is often crucial in the
microdistribution of benthic invertebrates(§ 5.3.4). HILDEBRAND (1974)
has shown that the herbivorous nymphs of the mayfly Ephe~rella
needhami (a 'scraper') drift at a higher rate when fewer attached algae are
present on the stony substrate of an artificial stream, and HILDREW and
TOWNSEND ( 1980) demonstrated that the average time before the car-
nivorous larvae of the caddis Plectrocnemia conspersa drift from an experi-
mental stream is significantly shorter when prey are not available.
Experiments have shown that individuals are also more likely to drift in
the presence of competitors. This may occur simply as a result of the
'jostling' of animals feeding at high density (WALTON, REICF, and ANDREWS,
1977). However, in the case of P. conspersa, and other sessile net-spinning
caddis larvae, space in which to spin may be a directly limiting factor.
Highly aggressive behaviour has been reported for hydropsychid caddis
64 M0LLER'S COLONIZATION CYCLE

in uowclecl situations, and the larvae have been observed to stridula1e in


1he presence of conspecifics. This could se1ve as a spacing device. P.
compena, too, responds aggressively in 1he presence of an imruder.
Comact is usually momemary but occasionally it develops into a more
prolonged bout ol'biting. Evemually. one of the la1vae re1rea1s, often by
entering the drift. In the majority of cases the smaller one loses 1he
encounter and the winner remains in the net.
It is not always clear whether animals which enter the drill do so
actively, or passively as a result of increased suscep1ibili1y to dislodgeme111
(due to more active foraging when food is scarce, or grea1er activi"· in
encounters with competitors). It is important w note 1ha1 the adap1ive
advan1age associated with leaving an unfavourable patch may accrue ID
an individual whether it deliberately emers the drift or not.

6.3.4 A model of the role of drifl in benth1c microdij/rib11tion


The results presented in 1his chapter suggest that the probabili1v is
really quite high that during the course of a few clays any given individual
will significantly shift its position by entering the drift. Many will g,1in
some adaptive advantage by leaving an unfavourable patch ol'stream bed.
The move will probably consist of a short hop of a few lens of cenl inll'IIT~
to land randomly on another patch. This may prove ID be a sui1able
microhabitat with the appropriate abiotic condi1ions, food and few
competitors, in which case the probability will be low tha11he animal will
re-enter 1he drift. In this specific connection, WALTON ( t 978) has shown
that the likelihood of nymphs of the mayfly Stenacron i11terpw1r/a/11m IT-
entering the drift within a short time (5-30 min) of se1tling on a subs1ra1e
at nigh1 is significantly higher if epilithic food is not present.
It can be postulated that movements made in contact with the stream
bed will be determined in a similar way. Thus, for any given species.
despite the continuous redistribution that goes on, bottlenecks will occur
in the most suitable patches. It is these bottlenecks that are shown up by
microdistribution studies. That the pa11erns are not more clear cut, with
each species showing precise habitat preferences, is due to the strong
random element involved. Not only are an unmeasured proportion of
movements accidental and, from the point of view of the animal involved,
undesirable, but also small-scale substratum shifts must frequently occur.

6.4 Muller's colonization cycle


There is clearly a downstream displacement of organisms in the drift,
and the question can be posed: do upstream movements on the bottom
compensate for the downstream losses? The results of several studies
show that they do not. At most 50%, and usually far fewer, of downstream
losses are compensated by upstream migration of the larvae. Why, then,
do, upstream stretches remain populated? Part of the answer is, of course,
that a continuous net downstream displacement of organisms need not
M0LLER'S COLONIZATION CYCLE 65
denude areas upstream if there is also a continuous or periodic
replenishment through reproduction by those individuals remaining,
and one view has it that drift can be considered as the removal of animals
surplus to the environment's carrying capacity. Eggs of many species are
conce111ra1ecl in time and space. II is clear 1ha1 growing larvae could nor
all continue to coexist in a small patch where the eggs from which they
developed were laid, so a large-scale density-dependent dispersal is
inevitable. Density-dependent drift has been demonstrated both in
n,mire (rECKARSKY, 1979) and in the laboratory(§ 6.3.3).
Nevertheless, the 'surplus 10 requirements' view cannot be the whole
story. We have seen that the appropriate habitat of a species is of a limited
extent clown the length ofa given river, and that often quite rapid faunal
changes occur over short distances (Chapter 3). Thus, the downstream
displau:ment means that many animals drift out of the section of river ro
which they arc adapted. and sometimes into a lake or the sea where·
conditions arc usually totally unsuitable. I I is reasonable to postulate 1ha1
an evolutionary advantage will have accrued to adults which movc·d
upstream to lay eggs at the head of the appropriate section of stream,
sinn· their offspring would have stood the best chance of dcvelopin~
within an optimal environment. I
In 1954 Muller put forward his 'colonization cycle' hypothesis whid
relates to the insects of running water. The hypothesis consists of tht
concentration of eggs and young larvae in an upstream reach, the
downstream drift of immatures 10 colonize all suitable habitats, and an
upstream flight of adults to complete the cycle. That the immature stages
of stream insects drift at high rates and colonize suitable habitats in their
stream section is not in doubt. However, studies concerned with flight
direction have produced less clearcut results. As a rule, the adults are not
strong flyers and their flight direction may often simply reflect the
prevailing wind. This will tend to confound the results.
However, in a study of thirty-four species in which adults flying
upstream or downstream in a south Swedish stream were taken separately
in specially constructed traps, 55.896 of all individuals were moving
upstream. Some species, notably the caddisfly RhyacophilajasciaLa, showed
a much greater preponderance for upstream flight (svENSSON, 1974). ono
and svENSS0N ( 1976) approached the problem in a novel, if somewhat
drastic, manner. They carefully removed by hand all the pupae of the
cased caddis Polamophylax cingulalus from the uppennost 1.6 km of the
same stream in Sweden. The larval population in the following season was
approximately half as big as usual, suggesting that immigrant females
were responsible for about half of the nonnal recruitment.
One particularly graphic account of upstream migration refers to the
adults of a stonefly, Capnia atria, walking upstream along snow-covered
banks!
MOLLER ( 19 74) has himself pointed out that while the results of many
studies of flight direction support his 'colonization cycle' hypothesis,
66 M0LLER'S COLONIZATION CYCLE § 6.4
nevertheless some insects do 1101 display the normal pauern of upstn:am
flight and downstream drift. II has not been universally confim1ed and is
probably not universally applicable. There is no doubt, however, that it
has been very inAuential in the development of ideas about benthic
community dynamics.
Stream and river invertebrates which do noc have flying aduhs present a
rather different case. Gammarid crustaceans, for instance, are often
observed to exhibit high drift races. Presumably, the tendencv of
individuals lO move upstream coupled with a high rate of reproduction
serves lO keep upstream stretches populated. Members of another
invertebrate group are totally incapable of counteraCling the movement
of water in streams and rivers - the zooplankcon. These are only strongly
represented, as are the phytoplankton, in rivers where the time taken for
water 10 move from source to mouth is relatively long (§4.2.2). The
populations originate from backwaters and adjacent ponds, lakes and
swamps. They feed, grow and reproduce in the column of water as it
moves towards the sea, forming an important but 1empora11· componc-nt
of the river system al any point along its length. A small proportion m,1y
accidentally find its way into backwaters or lakes, further downs1ream,
where their descendants may persist, but the vast majoritv must
eventually perish in the sea.
It is clear 1ha1 the unidirectional Aow which charaClcrizes lotic
ecosystems is of crucial significance in the lives of both benthic and
plankconic organisms, reiterating a theme which has appeared in evny
chapter.
References

BARLOCHER, F. and KENDRICK, B. ( 1975a). Leaf conditioning by microorganisms.


Ouologia, 20, 359-62.
BARLOCHER, F. and KENDRICK, B. ( 1975b). Assimilation efficiency of Gammarw
puudol1mnatUJ (Amphipoda) feeding on fungal mycelium or autumn-shed
leaves. Oili.oJ, 26, 55-g.
BEREZAY, G. and GEE, J. H. ( 1978). Buoyancy response to changes in water velocity
and its function in Creek Chub (Snnotilw alromaculalw). J. Fuh. RtJ. Board
Can., 35, 1195-9.
BOVBJERG, R. v. (1970). Ecological isolation and competitive exclusion in two
crayfish (OrconuttJ viriliJ and OrcontcttJ immuniJ). Ecology, 51, 2115-36.
c1;MMINS, K. w. ( 1974). Structure and function of stream ecosystems. BioJcimct, 24,
631-41.
F.DINGTON, J.M. ( 1968). Habitat preferences in net-spinning caddis larvae with
special reference to the inHuence of water velocity.}. Anim. £col., 37, 67 5-g2.
HASLAM, s. M. ( 1978). Rivtr PlantJ. Cambridge University Press, Cambridge.
HILDEBRAND, s. G. ( 1974). The relation of drift to benthos density and food level in
an artificial stream. limnol. Octanogr., 19, 951-7.
Ill LDREW, A. G. and TOWNSEND, c. R. ( 1980). Aggregation, interference and foraging
by larvae of Pltetromtmia coruptrJa (Trichoptera: Polycen1ropodidae). Anim.
Behav., 28, 553-60.
HOLT, c. s. anci WATERS, T. F. ( 1967). Effect of light intensity on the drift of stream
invertebrates. Ecology, 48, 2115-34.
HYNF.S, H. B. N. (1970). Tht Ecology of Running WattrJ. Liverpool University Press,
Liverpool.
INGOLD, C. T. (1975). Aquatic and Wattr-bomt HyphomyctttJ. F.B.A. Scientific
Publication No. 30.
LIKENS, G. E. and BORMANN. F. H. (1971). Nutrient cycling in ecosystems. In:
EcoJJJltm Structurt and Function (Ed. J. A. Wiens). Oregon State University
Press, Oregon.
LOCK, M. A. ( 19 75). An experimental study of the role of gradient and substratum
in the distribution of two s1ream-dwelling criclads. FrtJhwal. Biol., 5, 111-16.
LOCK, M.A. and REYNOLDSON, T. B. (1976). The role of interspecific competition
in the distribution of two stream-dwelling triclads.J. Anim. £col., 45, 581-g1.
MANN, K. H. (1975). Patterns of energy Row. In: Rivtr Ecology (Ed. B. A. Whitton).
Blackwell, Oxford.
McLAV, c. L. ( 1970). A theory concerning the distance travelled by animals entering
the drift of a stream.}. Fuh. RtJ. Board Can., 17, g5g--70.
MOLLER, K. (1974). Stream drift as a chronobiological phenomenon in running
water ecosystems. Ann. Rtv. £col. SyJt., 5, gog-,g.
OTTO, c. and SVENSSON, B. w. ( 1976). Consequences of removal of pupae for a
population of Potamophylax cingulatw (Trichoptera) in a South Swedish
stream. Oili.oJ, 17, 40--3.
PATTEE. E., LASCOMBE. c. and DELOLME. R .• (197g). Effects of temperature on the
distribution of turbellarian triclads. In: Effects of Ttmpnalurt on Ectothtrmic
OrganiJmJ (Ed. W. Wieser) Springer-Verlag, Berlin.
68 REFERENCES

PECKARSKY, e. L. ( 1979). Biological interactions as determinants of distribu1ions of


benthic invenebra1es within the subs1rate of stony streams. limnol. Ouanogr.,
14, 59-68.
PETERSEN, R. c. and CUMMINS, K. w. ( 1974). Leaf processing in a woodland s1ream.
Fmhwat. Biol., 4, 343~8.
PHILLIPSON, J. (1966). Ecological Entrgttics. Studies in Biology no. 1. Edward
Arnold, London.
RABENI, c. F. and MINSHALL. G. w. ( 1977). Fac1ors affecting microdis1ribu1ion of
stream benthic insects. OiAos, 19, 33-43.
SHORT, R. A. and MASLIN, P. E. ( 1977 ). Processing of leaf litter by a SI ream
deu:itivore: effect on nutrient availability to collectors. Ecology, 58, 935-8.
SVENSSON, e. w. ( 1974). Population movements of aduh Trichoptera a1 a South
Swedish stream. OiAos, 15, 157-75.
TOWNSEND, C. R. and HILDREW, A. G. ( 1976). Field experiments on the drif1ing,
colonisation and continuous redistribution of stream bemhos.j. Anim. Ecol.,
45, 759-71.
VARLEY, M. E. ( 1967 ). British Frtshwattr Fishts. Fishing News Books, London.
WALTON, o. E. (1978). Substrate attachment by drifting aqua1ic insen larvae.
Ecology, 59, 1013-30.
WALTON, o. E.• REICE, s. R. and ANDREWS, R. w. (1977). The effects of densi1y,
sediment particle size and velocity on drift of Acronturia abnormis ( Plecop1era).
OiAos, t8, 191-8.
WATERS, T. F. (1971). Thcdrift of stream insects. Ann. Rtv. Entomol., 17, 253-7 2.
WILLIAMS, D. D. and HYNES, H. B. N. (1976). The recolonisation mechanisms of
stream benthos. OiAos, 17, 165-72.
lcont111ued from inside front cover)

78 Animals for Man John C Bowm


79 The Biology of Weeds Thom as,
11111111111111
ITM00007272
h

80 Colon izat ion of Industrial Wastel


Raymond P Gemmell .IS
81 The Social Organization of Honeybees Peter J Edwards and Stephen D Wrat en
John B Free 122 The Ecology of Streams and Rivera
82 An Introduction to Animal Tissue Cohn R Townsend
Culture J A Sharp
83 The Genetic Code Brian F C Clark Other volumes in preparation
84 Lysosomes Roger T. Dean
85 Science and the Fisheries D H Cushing The Biochemistry of Pollution
86 Plants and Temperature J H Ottaway
James F Sutcliffe Light and Plant Life
87 Evolution in Modern Biology F R and J M Whatley
J R Edwards Lipids and Polysaccharides in Biology
88 The Biology of Mucor and its Allies AnnaJ Furth
C T Ingold Bioenergetics of Autotrophs and
89 Life i n Sandy Shores Alan E Braf1eld Heterotrophs John W Anderson
90 Antibiotics and Antimicrobial Action Invertebrate Respiration Rufus M G Wells
Stephen M Hammond and Peter A. Lambert lmmunobiology Christopher J lnchley
91 Boreal 1: .... - • - - · , .,, , _,. - - """ ,.._ ·· '
92 Nitrr
93 Hae
94 Prin
Geo
95 Str~
R II\
96 P
97 II/
98 P
JI QH Townsend, Colin R.
99 TI The Rcology o~ Streams and
OI 5·41, 5
.S7 Rivers
100 C
J T6
101 S C.2
102 TI
Kl
103 S
Pl
104 B
105 Li
106 A
107 PI
108 A
109 LI
R
110 SI
D
111 II/

'o,"'
112 LI
113 TI
C
114 N
J
115 V
C
116 lr _
Development Roy A L Batt
117 Collagen : the Anatomy of a Protein
John W oodhead-Galloway
lo.gyno. 122

n ma an lmegmed and coherent


u of the ec:ott,gy of rumfng waters The princlpal
me ls patten, In the physico--chamic
nv ronment. ln the distribution of orgamema and In
community orgarnzation. Causes and CCJnSeQuences of
tt are 9lq)lored n both the large scale 88
ngilud, I v adon down the length of rivers and on
the micro scale. as moaaics wfthJn very short stretches.
y looking n ii at distributional phenomena within
1v1dual nvors It c mes possible to identify some
eneral pnnciptes and unifying conceptS Th1& approach
coritrasts with the largely descriptive naturi, of much of
the existing literature.
Streams a rivers are ,mportant features of all
la scapes and a thorou h understanding of their
ecology 1s a prerequisite Jqf:!Sffl,ctive m,nagement and
oo rvat:ifDllr,flil~l.oli.tiio
admirable
at any

IIWflBQ'ed by The
ng the
ng the
olog(sta.
tute lta
uld be
Oueena

You might also like