Dynamics of the Atlantic salmon (Salmosalar L. ) population of the Foyle, .

by

Patrick Boylan

A thesissubndtted for the degreeof Doctor of Philosophy.

Division of Environmental and Evolutionary Biology

Institute of Biomedical and Life Sciences

University of Glasgow

March 2004

OPatrick Boylan, 2004 Acknowledgments

I wish to acknowledgethe help and support of the Chief Executive, Mr. Derick

Anderson,the staff and the Board of the LoughsAgency who providedthe majority funding for this study. I also wish to recognisethe financial supportof the Marine

Institute and in particularDr. Niall O'Maoileidigh. Thanksare also due to the rest of

Agency's Scientific Advisors Dr. Walter Crozier, Dr. Philip McGinnity and Dr.

GershamKennedy for their manyhelpful comments. I wish also to acknowledgethe help and supportI receivedfrom manypeople at GlasgowUniversity and in particular my supervisorDr. Colin Adams. Thanks are also due to Prof. Felicity Huntingford

Midam Geurts,Sanne de Boer andthe staff at the University Field Station. Contents Page

Page no.

Summary i

Chapter 1. General Introduction 1

1.1 Atlantic salmonLife-cycle 1

1.2 ConservationStatus of Atlantic salmon 3

1.3 PhysicalDescription of Foyle Catchment 4

1.4 SalmonManagement in the Foyle System 4

1.5 PopulationRegulation 7

1.5.1 Density-DependentFactors 7

1.5.2 Density-IndependentFactors 8

1.6 PopulationRegulation in Salmonids 9

1.6.1 PopulationRegulation Processes in the JuvenileFreshwater 9 Phase

1.6.2 PopulationRegulation Processes in the Marine Phase 11

1.6.3 Population Regulation Processesin the Adult freshwater 13 Phase

1.6.4 MathematicalModels 14

1.7 Modelling of JuvenileSalmonids 17

Chapter 2. Competing modes of exploitation and their effects on 21 spawning successin Atlantic salmon (Salmo salar L. ) in the Foyle catchment, Ireland

2.1 Introduction 21

2.2 Materialsand Methods 23

2.2.1 StudyArea 23 2.2.2 Exploitation Catch Data 24

2.2.3 Salmon Population Size Estimates 26

2.2.4 Analysis 27

2.3 Results 27

2.4 Discussion 37

Chapter 3. Life-stage specific, stochastic environmental effects, 41 overlay density-dependent filial cohort strength effects in an Atlantic salmon (Salmo salar L. ) population from Ireland

3.1 Introduction 41

3.2 Materials and Methods 43

3.2.1 Study Area 43

3.2.2 Salmon Population Estimation 43

3.2.3 Population Structure 46

3.2.4 Data Analysis 47

3.3 Results 54

3.3.1 Density-Dependent Regulation 54

3.3.2 Life-Stage Specific Environmental Effects on Population 57 Size

3.4 Discussion 67

Chapter 4. The influence of broad scale climatic phenomena on long 72 term trends in Atlantic salmon population size: an example from the , Ireland

4.1 Introduction 72

4.2 Materials and Methods 75

4.3 Results 77

4.4 Discussion 81 Chapter 5. Local instream and catchment spatial scale habitat 83 characteristics determine 0+ fry density of Atlantic salmon (Salmo salar L. ) in the River Foyle

5.1 Introduction 83

5.2 Materialsand Methods 84

5.2.1 CatchmentScale Characteristics 86

5.2.2 Analysis 89

5.3 Results 90

5.4 Discussion 94

Chapter 6. General Discussion 98

References 104 List of Figures

Figure no. Legend Page no.

1.1 Life-cycle of the Atlantic salmon 1

1.2 Location map of the Foyle catchment 4

1.3 Five year running averageof commercial salmon catches 6 in the Foyle 1993-2001

1.4 Hypothetical stock-recruitment model 14

1.5 Three common stock-recruitment curves: 0--Linear; 16 2=Beverton & Holt (1957); 3=Ricker (1954)

2.1 Drift, Draft nets,uncorrected Sport Angling catches& 29 uncorrectedredd counts in the Foyle area1952-2000

2.2 Regressionof total annualcatch of salmonin the Foyle 35 areaand corrected redd counts

2.3 Commercialcatch residuals regressed on correctedangling 37 catch

2.4 Total catchresiduals regressed on correctedredd counts 38

2.5 Total catch residualsregressed on correctedredd counts 38 excludingtwo outlying points

3.1 Salmonlife history stagesfor onelife-cycle 49

3.2a A linear model fitted to parental population size (egg 55 equivalent *100,000) and filial population size (egg equivalent*100,000)

3.2b A Beverton& Holt model fitted to parentalpopulation size 56 (egg equivalent *100,000) and filial population size (egg equivalent*100,000)

3.2c A Ricker model fitted to parental population size (egg 56 equivalent *100,000) and filial population size (egg equivalent*100,000)

3.3 Regressionsof environmentalvariables during life-stage 58 specific eventsand populationsize residualsderived from the Ricker parental-offspringcurves 3.4 Mean residual population size values derived from the 64 Ricker parental-offspring curves for years with high (upper 30 percentile); low (lower 30 percentile) and intermediate (remaining 40 percentile) values for environmental variables during the life-stage specific periods.

4.1 Annual commercial catch of salmon in the Foyle area 77 1875-2000

4.2 The five year averagecommercial catches of migrant 78 Atlantic salmonfrom the Foyle catchment1875 to 2001 & the five yearaverage NAOI: 1875- 2000

4.3 The relationship between the winter NAO and commercial 79 fishery catches of Atlantic salmon returning to the River Foyle over 122 year. Breakpoint analysis shows a uncoupling of a negative relationship about a NAOI of 0.151

4.4 The change in North Atlantic Oscillation predicted by 80 sevenclimate changemodels. Dotted line showsthe mean of the sevenmodels, solid lines showthe 2 extremesof the predictedNAO change

6.1 Foyle area annual commercial salmon catches 1962-2003 102 with regression line List of Tables

Table no. Legend Page no.

Stock-recruitmentequations for Beverton & Holt (1957) 16 andRicker (1954)models

2.1 Correlations of commercial fishing catches and corrected 32 sport angling catches

2.2 Correlationsof commercialfishing catches,corrected sport 34 angling catches, corrected total catch, corrected redd countsand corrected total population

3.1 Comparisonof predictive linear; Beverton & Holt and 54 Ricker models

3.2 Regressionof environmentalvariables during life-stage 60 specific eventsand populationsize residualsderived from the Ricker parental-offspringcurves

3.3 ANOVA results of environmentalvariables during life- 62 stagespecific eventsand populationsize residualsderived from the Ricker parental-offspringcurves

5.1 Local scale site-specific variables collected for each 85 samplingfor 350 sitesin the Foyle catchment

5.2 Catchmentscale characteristicscalculated from ordnance 86 survey and geological survey maps for each of the 350 samplingsites in the Foyle catchment

5.3 Stepwise regression of local site specific stream 90 characteristics on 0+ salmon numbers for a) all sites (including sites with no 0+ salmon); b) only sites where 0+ salmon were present

5.4 Stepwise regressionof catchmentcharacteristics on 0+ 91 salmonnumbers for a) all sites (including siteswith no 0+ salmon);b) only siteswhere 0+ salmonwere present

5.5 Stepwise regression of local scale and catchment 92 characteristicscombined on 0+ salmonnumbers for a) all sites (including sites with no 0+ salmon); b) only sites where0+ salmonwere present Summary

This study examines the dynamics of the Atlantic salmon (Salmo salar L. ) population of the Foyle catchment in Ireland, through the analysis of long-term datasets and extensive field data. In Chapter I the current conservation status of the salmon is discussed with particular reference to the Foyle. A number of methods used in studying the juvenile and adult life-stages are reviewed.

Chapter 2 considersthe interaction betweencommercial netting and recreationalsport angling and the effect of total combined exploitation on an independentmeasure of populationsize (redd counts)using a 49 year dataset. While recognisingthat commercial netting had a relatively small negativeimpact on recreationalsport angling, the evidence suggeststhat year class strength is the principal modulator of variation in commercial catches, sport angling catchesand spawning escapement.

Chapter 3 examines the role of density-dependencein the Foyle salmon population. A

Ricker density-dependent model showed that spawning adult population size significantly predicted variation in the resultant filial generation, however, a significant amount of variation (ca. 68%) remained unexplained. Environmental factors were significant in explaining some of the remaining variance and these influences were linked to specific life stages. This finding strongly suggestspopulation bottlenecks in the complex life cycle of this species, during which specific environmental factors may have an impact, that they do not have during other periods. It was concluded that these life stage specific environmental effects are likely to contribute to the stochastic variation in population size resulting from the application of traditional stock-recruitmentmodels and that the identification and quantification of these effects should allow improved model accuracy.

Chapter4 investigatesthe effect of marineclimatic conditionsin the North Atlantic on the abundanceof returningmigrant Atlantic salmon,using a 126 year datasetof commercial catchesand an index of climate in the northernhemisphere, the North Atlantic Oscillation

(NAOI). The NAOI when below 0.151 explaineda significant proportionof variation in five year running meancatches of migrant Atlantic salmonreturning to the River Foyle.

This indicatesthat a significant proportion of the variancein populationsize in the past was the result of variability in conditionsin the marineenvironment. However, when the

NAOI was above0.151, this relationshipuncoupled. Models of climate changeindicate that the NAOI is likely to increasesignificantly with time. If thesemodels are correct,this study would lead to the conclusionthat a decouplingof broad scaleclimate effects on salmon population size will become the norm. Data presentedhere suggest two consequencesof this: that the value of the NAOI as a predictivetool for forecastingadult salmonpopulation size will be limited; and that the medianpopulation size will become lower in the future.

Chapter 5 tests the capacity of local instream and broadscale catchment characteristics to predict 0+ salmon abundance within the Foyle area. Using a combination of local site specific data, from 350 sites chosen for annual electrofishing surveys of 0+ salmonids, and semi-quantitative morphometric information and broadscale catchment characteristics derived from Ordnance and Geological survey maps, two models were constructed. Model

I used data for all available sites i. e. including those where 0+ salmon were absent, while model 2 used only those sites where salmon were present. Both of these models were

ii significant predictorsof juvenile abundanceusing site-specificvariables. However,both models were improved significantly by the inclusion of wider broadscalecatchment characteristics(model I r2=30%;model 2 r2=43%). When testedagainst an independent data set using paired Mests neither model differed significantly from the actual catch of salmon,although both had relatively high standarderror rates (84%±13 and 56%±15.2 respectively). It was concludedthat site-specifichabitat characteristicswere significant predictors of juvenile abundance,but with the inclusion of broadscalecatchment characteristicsthe models predictive power was greatly increased. This chapter also highlighted the potential detrimental impact of increasingurbanisation on the salmon stocksof the Foyle area.

iii Chapter 1. General Introduction

1.1 Atlantic salmon (Salmo salar L.) Life-cycle

The Atlantic salmon (Salmo salar L. ) is an anadromous species of fish whose range extends on both sides of the North Atlantic ocean. Its life cycle is outlined in Figure 1.1 and has been described by various authors (Jones, 1959; Shearer, 1992; Hansen, L. P. &

Quinn, T. P., 1998; Klemetsen et al., 2003).

Figure 1.1 Life cycle of the Atlantic salmon (Salmo salar L. )

'A ter Diet: Diet: F\ edOva

AL 7.77t'Air

Alevin

(-3 C111)

Sand Eel

A Parr (4- 1 cn 0 12-1

I Copyright Central Fisheries Board, The adults spawn in gravel areasof streamsand betweenNovember and January each year. The femaleslay eggs in shallow depressions(called redds) which are then fertilised by the male fish and covered by gravel. The eggs remain buried until

February/Marchwhen they hatchand the alevinsmigrate through the gravel into the river.

The juvenile fish will remain in the river for at least one year, but more usually two in

Ireland (Crozier & Kennedy, 1997)before going to seaas smolts. A numberof marine feeding areashave been identified, namely the Faroes,Iceland and the west coast of

Greenland. The fish will remain at sea for one or more winters. Those that return to freshwater after spending one winter at sea are commonly known as I seawinter salmon or grilse, but if they spendtwo or more winters at sea they are known as multi-seawinter salmon.

Adult fish may return to their natal streamat any time of year but most rivers have well defined periods of migration. For example,in the Foyle catchment,Ireland, the system which is the subjectof this study, the River Finn has a large numberof multi-seawinter salmon returning before the end of May supportinga recreationalspring fishery. This river also has an early I sea-winterrun of salmonwhich return at the end of May and into

Juneand July, with few fish generallyentering the river after mid July. In contrast,on the

River Mourne, a neighbouringcatchment, the majority of salmonenter the river during

June,July and August with salmonreturning, albeit in smaller numbers,until December

(LoughsAgency, unpublished data).

On their return migration from the sea salmon stop feeding and will live on their body energyreserves until spawning. After spawning,some salmon known as kelts maketheir way back downstreamand return to the marineenvironment. A small proportionof these

2 may return to spawn in subsequentyears. In Scottish Rivers previous spawners contributed between5-10% of spawning stock on average,with >95% of these being female(Shearer, 1992).

1.2 Conservation Status of Atlantic Salmon

In much of its rangethroughout the north Atlantic, the Atlantic salmonhas declined,both in termsof catchesand actual population size, since the 1960'sand early 1970's(Parrish et aL, 1998;Jonsson et aL, 2003).

In someparts of its range,such as the Bay of Fundy in Canada,and the Gulf of Maine in the USA, stockshave been listed as endangered(Chase, 2003). Commercialharvesting of salmonhas been stopped in manyareas of North America as a conservationmeasure, with the exceptionof somesubsistence catches (O'Connell et al., 1992;Dempson, et al., 2001b) and in some regions sport angling has been closed, or is subject to closure during the normal sportangling season(Dempson, et al., 2001a).

In a Europeancontext, commercialfisheries for salmon have been tightly regulatedfor many years. Despitethis however,southern stocks, in contrastto stocksin more northerly latitudes, still appearto be declining (Parrish et al., 1998). In order to protect current strongstocks of salmonthe EuropeanUnion (EU), underthe Directive on the conservation of natural and semi-naturalhabitats and of wild fauna andflora (92143EEC)(Habitats

Directive 1992)(Anon., 1992a), havedesignated many rivers throughoutEurope for extra protectionas SpecialAreas of Conservation(SAC). Within the Foyle systemthe Rivers

Finn, Mourneand Derg havebeen afforded such designation.

3 1.3 Physical Description of the Foyle System

The Foyle catchmentis situatedin the north west of the island of Ireland (Figure 1.2) and is approximately 4,5OOkM2in area. The catchment encompassestwo international jurisdictions; thoseof the and , which is part of the United Kingdom.

Figure 1.2 LocationMap of the Foyle catchment

N FoyleCatchment A

R. Roo 2ý1 C,

R. FWin

OWO*kw P,

Ft S" 10

05 10 20 30 4

1.4 Salmon Management in the Foyle System

The Loughs Agency is a statutorycross-border body, which aims to provide sustainable social, economic and environmental benefits through the effective conservation, management,promotion and developmentof the fisheries and marine resourcesof the

Foyle and Carlingford Areas. Ile Loughs Agency and its predessorthe Foyle Fisheries

4 Commissionhave collated a long dataset of catchfigures and populationestimates. These data indicatethat salmonabundance peaked in the late 1960's and subsequentlydeclined until the early to mid 1970's.

As a result of thesedeclines, a report was commissioned(Elson & Tuomi, 1975)which examinedthe salmonresource of the catchmentand maderecommendations as to how best managethe stock. Following this report the Foyle FisheriesCommission set up a real time managementregime, which, with somerefinements, is still in operation. Managementof the fishery is predicatedon meeting in-seasonpopulation size targetsfor the returning migrant populationabove a control site on the at measuredusing a Logie resistivity 2100C fish counterin-addition to a visual count of salmoncrossing the weir outwith the countingchannels. The first of thesetargets is the 3& June,that is, a little over two weeks into the commercialfishing season,which starts on the 15a'June.

This first target is setat 2,600returning migrant salmon(or in periodsof low water on the next flood). If the targetis not achieved,depending on the size of the flood, as measuredat a local Departmentof the Environment gauging station, both commercial fishing and recreationalangling will be curtailed by either 24 or 48 hours respectively. Similarly by the 10thJuly if a target of 4,200 returning migrant salmonhas not passedSion Mills, the fishing may againbe closedfor either 24 or 48 hours. However,the fishing may only be closed for a maximumof 48 hours in any one season. If by the 24thJuly the numberof salmonabove Sion Mills exceeds8,000, an extensionof 96 hoursfishing will be grantedto commercial netsmen. Finally, if by the 2& Septemberless than 7,000 salmon have crossedthe weir at Sion Mills recreationalangling will be curtailedby 10 days.

5 This end of season target of 8,000 salmon is based on the amount of nursery habitat

available upstream and its juvenile carrying capacity. Target egg deposition levels are set

-2 -2 for three juvenile habitat grades i.e. Grade I= 10 eggs in ; Grade 2=5 eggs M and

Grade 3=2.5 This habitat eggs M-2. combination of quality and quantity assessment

equates to approximately 6,500 salmon (male and female) spawning at the end of each

seasonwithin the catchment. Twenty five percent is added to this figure to allow for sport

angling, poaching and natural mortality bringing the final target population size to 8,000.

These targets are monitored using data from fish counters and adjusted for a visual

assessmentof the numbers of fish which cross the weir without passing through the

counting channels. The counting site isjust above the tidal limit.

Since this managementregime's introduction, stocks have increased and there is currently

a five-year running average commercial catch of salmon between 30,000-35,000 (Figure

1.3).

Figure 1.3 Five year running average of commercial salmon catches in the Foyle area 1993-2001.

50000

45000

35000 C 0 30000 E 2 25000

6 20000

15000

10000 5000 N 0 1997 1998 1999 2000 2001 Year 1.5 Population Regulation

1.5.1 General

As with all animals, mechanismsoperate to limit the size of individual populations.

Alisauskasand Arnold (1994)in a studyon Americancoots (Fulica americana)found that aI inearrelationship explained a high proportionof variance(12---0.54; p<0.00 1) betweenan estimateof populationsize (numbersshot eachyear) and the numberof temporaryponds availablefor breedingon the prairiesthe preceedingsummer. Models suchas theseallow predictionsof spatialpatterns of habitatuse and give an insight into mechanismsoperating on the population,thus enabling predictionsto be made which support population and habitat management.This is particularly importantin speciesof high conservationvalue or which are exploited.

Many studieshave attempted to identify thesemechanisms for Atlantic salmonbut most of these have been limited by the use of short-termdata sets or small sample size. One generalaim of this study is to uselong-term data sets on populationsize and exploitation to attempt to identify mechanismsand principal controlling factors influencing population size in the Foyle catchment.

Population regulatory mechanisms are frequently divided into two main types: density- dependentand density-independent factors.

1.5.2 Density-Dependent Factors

Density-dependentcauses of mortality are thosewhose mortality rate changeswith density e.g. competitionfor spaceand food (Elliott, 1985,1994;Haldane 1953; 1956;Kennedy &

7 Strange,1982; Kennedy& Strange,1986a; Kennedy & Strange,1986b). Both negative and positive effectsof populationdensity on survival havebeen shown for fish. Chapman

(1986) in a review of historical catch recordsof Pacific salmon species(Oncorhynchus tshawytscha0. kisutak 0. nerka 0. keta, 0. mykiss)from the ColumbiaRiver in North

America, identified reduced aboriginal fishing pressure,around 1800, as the cause of subsequentdecreasing runs. The author postulated that this was due to excessive escapementof salmonup river and density-dependentmortality occurringin the freshwater habitat. Heavy fishing pressureafter 1850probably increasedruns initially, but led to a sharp subsequentreduction as stocks were over exploited. Ross and Almeida (1986) showedthat silver hakes(Merluccius bilinearis) were subjectto density-dependentcontrol mechanisms,with lower growth ratesat high stockdensities.

Negative density-dependencemay also be known as concurrent (Solomon, 1949), compensatory(Neave, 1953)or regulatory(Nicholson, 1957)and simply meansthat as a population increases,the probability of individual survival decreases(Elliott, 2001).

However,population increases may not alwaysact negativelyon a species. For example, amonggoldfinches (Carduelis carduelis), birds in flocks of greaterthan eight consumedup to 2.3 times more seedsper unit time than singletons,thus increasingtheir survivorship potential(GRIck, 1986).

1.5.2 Density-independent Factors

Density-independentcauses of mortality are thoseexerting a similar effect on mortality, independentof the populationdensity. In general,density-independent factors are mainly thought to be environmentalin origin, with eventssuch as drought, temperature,storms

8 etc. being someof the main factorsidentified. Among Peregrinefalcons (Falco pergrinus) in Australia, for example,breeding successwas negatively related to rainfall with the effect that the total productionof the populationwas lowest in the wettestyears, mainly as a result of direct rainfall andrunoff flooding nestsites (Olsen & Olsen,1989).

In reality thereis no sharpdelineation between density-dependent and density-independent factors. For instance,deaths arising from biological causes,such as disease,parasitism, malnutrition and predation,which may be consideredto be initially density-independent will usually becomemore frequentas densityincreases. Density-dependent factors can act togetheror can compensatefor eachother. For exampleJenkins et al. (1963), in studying the effects of shootingon Red Grouse(Lagopus lagopus scoticus), found that lossesat

30% of the post-breedingnumbers were entirely compensatedfor by reducednatural loss, causingno depletionin breedingstocks (Jenkins et al., 1963).

1.6 Population Regulation In Salmonids

1.6.1 Population Regulation Processes In the Juvenile Freshwater Phase

There is evidencethat density-dependentmortality in the juvenile stagescan be a control of population size in salmonids(Ricker, 1954; Kennedy & Strange, 1982; Kennedy &

Strange1986a; Kennedy & Strange1986b; Egglishaw & Shackley,1985; Milner, et al,

2003; Ruggerone& Rogers,2003; Dumas & Prouzet,2003). Gee et aL (1978),in a study on the River Wye in Wales,found that maximumsmolt production,which they equatedto juvenile salmonsurviving to 2 yearsof age,was attainedat a fry densityof 0.75m'2on the l' of June. Above and below this stocking density, production decreasedfollowing a dome-shapedproduction curve. Gardinerand Shackley(1991), in a study of a six-year

9 dataset of salmonfry to their first and secondautumns on the ShelliganBum in Perthshire,

Scotland,found that productionfollowed a similar domeshaped Ricker model. However, in a review of the stock-recruitmentmodels of the time, Solomon(1985) suggestedthat, while Gee et al. (1978) werejustified in claiming that the curve was indeeddome-shaped, migration could explain differencesin the densitiesof fry in subsequentgenerations. In contrast,Gardiner and Shackley(1991) suggestedon the basis of work carried out by

Egglishawand Shackley(1977) that densityrelated losses in their study were most likely lossesof youngerparr dying in the stream. Elliott (1994,1989,1985),showed in his long- term studieson two small streams(Black Brows Beck and Wilfin Beck) in England,that density-dependentmortality is not necessarilyalways the most significant population regulatingeffect in salmonidpopulations. Black Brows Beck (holding a migratorybrown trout [Salmotrutta L. ] population)always had higherjuvenile densitiesthan Wilfin Beck

(containingresident brown trout) and was regulatedby density-dependentsurvival in the early stagesof the life cycle. There was no evidencefor this, however,in Wilfin Beck, wheresimple proportionatesurvival occurredwith fairly constantloss rates. Survival was reducedin both populationsby summerdroughts and also by spatesin Wilfin Beck. Black

Brows Beck trout were alwayslarger than thoseof Wilfin Beck of similar age,mainly due to fry size at the start of the growth season.Variations in water temperaturewere chiefly responsiblefor differencesin growth betweenyear classeswithin eachpopulation. Food intake was not a limiting factor except in the first winter of the life cycle and for adults over 3 years old in Wilfin Beck. Variation in individual size was inversely density- dependentin Black Brows Beck and decreasedwith age in Wilfin Beck, thesechanges being due to natural (stabilising)selection. There was also strongevidence for genotypic differencesbetween the populations.

10 Resultsfrom studiessuch as theseindicate a natural inter-systemvariation in the relative dominanceof density-dependentand density-independentcauses of mortality in population sizecontrol. This leadsto the conclusionthat, in orderto studythe dynamicsof individual systemsfor the purposesof populationmodelling, an assumptionof a genericrelationship appliedacross systems is likely to be inadequate.

1.6.2 Population Regulation Processes in the Marine Phase

Density-dependent mortality as a result of competition between Asian Pink salmon

(Oncorhynchus gorbuscha) and Alaskan sockeye salmon (a nerka) has been shown to occur at sea (Ruggerone et aL, 2003). It is not known if density-dependent mortality occurs with Atlantic salmon in the marine phase. As major differences exist between the

Pacific and Atlantic species it is impossible to apply conclusions based on Pacific salmon studies to Atlantic salmon. For example, Pacific salmon are the dominant fish species at sea while Atlantic salmon constitutes a very small percentageof the total marine biomass.

What is known, is that mortality in Atlantic salmon, once smolts leave the river, varies dependingboth on the river location and stock component(Anon., 2003c). Rod catches from the River Spey, Scotland indicate that migrating spring stocks of multi-seawinter salmon are falling (Anon., 1998a). However, numbersof grilse and summermigrating multi-seawintersalmon remain more stable. The mechanismsbehind Atlantic salmon survival at sea are not clearly understood. However, North American studies have observedgood correlationsbetween sea survival and oceantemperatures and the projected amountof winter habitatavailable to their stocksof fish at sea(Amiro, 1998).

11 In many Canadianrivers, juvenile productionis thoughtto be optimal. Despitereductions

in the commercialfishing effort and periodic increasesin the numberof salmonreturning to spawn,however, anticipated increases in subsequentrecruitment of spawningstock have

failed to materialise(Dempson et aL, 2001; Amiro, 1998;Ritter, 1997). This trend is also

shownby the decline in survival ratesof hatchery-releasedsmolts, which are independent

of the freshwaterhabitat. Amiro (1998) suggeststhat an increasein predatornumbers,

particularly harp seals(Phoca groenlandica), may explain the deficit betweenactual and

expectedreturns. In a Europeancontext, data from the in N. Ireland indicate

that salmonsurvival at seaoutwith direct anthropogenicinfluences has decreasedfrom an

averageof 30% pre 1998 (Crozier & Kennedy, 2001) to 10% in 2002 (Crozier et aL,

2003d). These declinesare also occurring at a time when smolt production from the freshwaterphase is decreasing(Kennedy & Crozier, 1991). Changesin the productivity

of Pacific salmon specieshave also been documentedand linked to climatic conditions

(Petermanet aL, 2003) and somestudies have developedpredictive models to track these

changes(Logerwell et aL, 2003). Chapter4 investigateslarge-scale climatic effects on

marine survival and the potential impact of global warming on future trends in salmon

abundancein the Foyle system.

With the decline in salmonabundance, attention has tendedto focus on the commercial netting sector and in particular on interceptorymixed-stock fisheries, such as those off

Greenland,the Faroes,the Northumberlandcoast of England, Northern Ireland and the

Republicof Ireland. In 2002 the Greenlandfishery was allocateda total allowablecatch of between20 and 55 tonneswhilst for the long line Faroesefishery no quota was set for

2002 or the 2003 season(Anon., 2003d). In Northern Ireland, within the Fisheries

ConservancyBoard area,a buy-out schemefor the commercialnetsmen was introducedin

12 2002. The drift net fishery limits off the Republic of Ireland's coasthave beenreduced from 12 to 6 miles and fishing is only allowedbetween 6am and 6pm Mondayto Thursday

(Anon., 1997a). In 1998the drift netscaught 58% of the total run of fish back to the Irish coast (O'Maoileidigh, pers. comm.). In 2001 carcasstagging was introducedon an all

Island of Ireland basis,whereby each individual salmon,whether caught commercially or by sport angling, must carry a tag with a unique identifying number.This has greatly reducedlarge scalepoaching and with the introductionof quotas(Republic of Ireland) to both commercialfisheries and sportangling, further restrictsexploitation on stocks.

Within the Foyle system,in line with theserestrictions, the commercialfishing seasonhas been reducedfrom March to Septemberin the 1960's and 1970's to a six-week period during Juneand July at present. In addition,there are only four days' fishing allowedeach week for both the drift and draft nets. The drift netshave also had the permissablefishing hours,reduced to 12 hoursper day.

1.6.3 2 Population Regulation Processes In the Adult Freshwater Phase

On return to fresh water, the main source of mortality in adult salmon is exploitation. It is estimated that estuarine or in-river draft nets catch approx. 28% to 30% of fish entering the

system on the River Eme. This figure is regarded as a national average rate of exploitation

for catchments where draft net exploitation occurs in Ireland (O'Maoileidigh, pers. comm.). Angling exploitationcan rangefrom 9.7% to 20.6%(Anon., 1992b)for grilse and

35-40% for stocks of Irish multi-seawinter salmon (O'Maoileidigh, pers. comm.).

Although, in the Foyle system, sport angling exploitation has reached50% for grilse

(Elson & Tuomi, 1975). As a responseto high exploitation levels, the Environment

13 Agency in Englandand Wales has introducedmandatory catch and releaseof salmonby anglers before 10 June (Anon., 1999). Many rivers in Scotland,while not having legislation to enforce similar controls, actively encourageanglers to return spring fish

(Atlantic SalmonTrust 1998). In the FisheriesConservancy Board areaof jurisdiction in

N Ireland,catch and release must be practicedon salmonprior to the I" June.

In this study (Chapter2) 1 attemptto examinethe role of competingexploitation methods on populationsize and also the interactionbetween these capture methods.

1.6.4 Mathematical Models

Severalmathematical models have been proposed to describedensity-dependence (Ricker,

1954;Beverton and Holt, 1957). When thesemodels are appliedto fish populations,they

are usually termedstock-recruitment models as they describethe relationshipbetween the

numberof recruits (R) to the populationand the parentstock (S) of fish. A hypothetical

stock-recruitmentrelationship is outlinedin Figure 1.4.

Figure 1.4 Hypotheticalstock-recruitment model.

Gn

I Teýgýgs) 14 Parents These models can be used to examinehow changesin, for example,fishing mortality,

affect yield. For thesemodels to be effective it is necessaryto know the productionof

parentstock over a wide rangeof recruit numbers. The parentstock can be expressedas

numbersof adults, biomassor egg production (Elliott, 1994; Chadwick, 1985; Elson &

Tuomi 1975).

A numberof stock-recruitmentmodels are shownin Figure 1.5 and their derivedequations

outlined in Table I. I. Model 0 shows a population that is increasing without any

restriction. While in model 1, as egg density increases,a dome-shapedcurve emerges.

Thus the initial effect of increasingpopulation size is density-dependenceoperating in a

positive manner. However,as the populationreaches maximum return it startsto decline

as negative density-dependenteffects become dominant. In model 2, the relationship

reachesan asymptoteas egg densityincreases (Elliott, 1994),but thereafterremains stable.

Figure 1.5 Three common stock-recruitmentcurves: 0--Linear; 1= Ricker (1954); 2=Beverton& Holt (1957).

12-

Cý% 10- 02 r: 8-

6- 0 1. 4- E 0 2- z 0-. 0 1111 20 40 60 80 100 120 140 Number of eggs (S m-2)

15 Table 1.1 Stock-recruitmentequations for Beverton& Holt (1957)and Ricker (1954) modelswhere R representsthe recruitsand S theparental stock. Proportionate survival Mortality rate [In (SIR)] Replacement Key (RIS) abundance fisheries Equation Function Relationship Function Relationship (S* for R =;S) references with S (all with S (all decreasing increasing

monotonic) monotonic) (I)R=aS exp(- alexp(bS) Concave bS-In a Linear (In a)lb Ricker bS) upward (1954) (2)R=aS1(l+bS) Concave In Convex (a-l)lb Beverton al(l+bS) upward (I+bS)-In upward & Holt a (1957) In most studies on salmonids to date, model 1, the Ricker (1954) stock-recruitment model,

hasbeen found to be the most applicable.Elliott (1994)found that this modelgave the best

fit to his datafor sea-trout(Salmo trutta L. ). Elson and Tuomi (1975), Gee et aL, (1978),

Gardinerand Shackley(1991), Kennedyand Crozier (1993) fitted Ricker curvesto their

data on Atlantic salmon and many North American studies of Pacific salmon have utilised

this model (Ricker, 1954,1989). Buck and Hay (1984), however,found that a Beverton

and Holt curve fitted their population data on Atlantic salmon, with survivor density

increasing to an asymptote (curve 2, Figure 1.5) rather than following a dome-shaped

curve.

All of thesepopulation models incorporatea number of inherentproblems. Regression

methods for estimating the parametersof these simple stock-recruitmentequations

generallylead to upwardbias in assessmentsof productivity at low spawningstock sizes,

and underestimatesof the spawningstock that would producemaximum averagesurplus

for harvest(Hilbom, 1997;Myers, et aL, 1995;Myers, et al., 1994).Bias is due to effors-

16 in-variables effects when spawning stock measurementhas been inaccurate(Elson &

Tuomi, 1975). It can also be difficult to generatedata points for an over-abundanceof parent stock or egg deposition(Kennedy & Crozier, 1991; Smith & Walters 1981),and thereforeit may be difficult to determineif the data set takesthe form of a dome-shaped

Ricker curve or a flat-toppedBeverton and Holt curve. There is also bias in time-series effects when harvestrates have been relatively stableso that recruitmentanomalies result in subsequentchange in spawningabundance (Walters & Ludwig 1981). There are no satisfactorycorrection methodsfor thesebiases, since they involve fundamentalloss of information about averageresponse, especially at low spawningabundance (Myers 1997;

Schnute& Kronlund 1996.). Beyond such bias problems,there is also uncertaintyas to how to measurethe reliability of parameterestimates and define policy prescriptions especiallyin circumstanceswhere the recruitmentrelationship may be non-stationarydue to persistent directionalecological or environmentalchanges (Amiro, 1998;Ritter, 1997).

As a result of these uncertainties,Walters (1981) recommendedthat regular probing experimentsbe carriedout on modelsand that in generalresults from thesetypes of model be treatedwith caution.

In this study,Chapter 3 attemptsto addressthe role of density-dependencein determining adult returnsand also the potentialfor environmentalfactors in influencingthese returns, with the useof long term dataavailable for the Foyle system.

1.7 Modelling of Juvenile Salmonids

Density-dependentmortality is known to occur in juvenile salmonids(Elliott, 1994,1989,

1985;Kennedy & Crozier 1991). Ibis may be as a result of competitionfor resourcesand

17 territoriality (Heggenes& Saltveit, 1990; Titus & Mosegaard,1992; Kennedy 1983;

Kennedy & Strange 1981; Marshall, 1995). The measurementand classification of instrearnhabitat areasis thereforecritical in estimatingthe potentialcarrying capacityof streamsand thereby deriving data for stock-recruitmentrelationships. In assessing instrearnhabitat, some studies have used wetted area of the river accessibleto fish (Elson

& Tuomi, 1975)while othershave quantified the availablenursery area (Inoue et al., 1997;

FHR, 1991). In theselatter studiesattempts were madeto quantify the river area,based on its morphology,and then relate the stock (as determinedfrom electrofishingsurveys) of each habitat reach or type to give an overall estimateof productivity. However, these modelsquantify actualproductivity at any given time and makeno attemptto predict fish numbers. In dynamicriverine environments,change, whether natural or man-induced,is always occurring, with concomitant effects on fish and other animal assemblages.

Therefore,predictive modelshave been developedin an attemptto monitor and explain these impacts (Wright et al., 1989; Whitehead, 1992; Milner et al. 1998). The River

InvertebratePrediction And ClassificationSystem (RIVPACS), developedby the Institute of FreshwaterEcology, is one suchexample (Wright et al., 1989; 1996). It encompassesa computermodel built from a databaseholding biological and physicaldata collected from over 8,000near pristine samplingsites in Englandand Wales. Siteswere selectedto cover as wide a range of environmental variables as possible, with the exclusion of unrepresentativepolluted sites. Invertebrateswere collectedfrom a wide variety of habitat types at each samplingsite and samplingwas repeated over different seasons.Samples were collected,sorted and identified to provide speciespresence/absence, abundance and diversity estimatesfor each site. The Biological Monitoring Working Party (BMWP)

scoreand AverageScore Per Taxa (ASPT) were computedfor each site as a measureof

the richness and diversity of the invertebrate community. Using the TWINSPAN

18 technique(Hill, 1979),RIVPACS invertebratedata were analysedto provide a systemfor the classificationof river siteson the basisof the distributionof invertebratefauna present.

In addition to the invertebratesamples collected, a large number of physico-chemical characteristicswere estimatedor measured,such as channelwidth, gradient,flow velocity and substratetype. On the basisof (a) theseenvironmental observations, (b) the observed invertebratepopulation data and (c) the TWINSPAN classifications,a predictivetechnique basedon multiple discriminantanalysis (MDA) was developed.This is usedto predict the probability of occurrenceof individual taxa, and BMWP and ASPT scoresfrom valuesof physico-chemicalvariables alone. At a river site of interest,RIVPACS can be used to predict targetvalues of BMWP, ASPT and the expectedcomposition of invertebratefauna on the basis of a limited number of observations/measurementsof physico-chemical parameters.Comparison of observedand predicted scoresusing results from the direct samplingof invertebratesat the site and RIVPACS predictions(based on physico-chemical parametersat the site), gives a measure of any loss of biological quality due to environmentalstress.

Another model which usesa similar approachto RIVPACS is HABSCORE(Milner et aL,

1993). This is a fisheriesmanagement model and is designedspecifically for salmonids.

Direct sampling by electrofishing is conducted at a number of sites, alongside measurementsand observations of physicalhabitat features, such as:

1) > Conductivity(pS cuf

> Cover

> Flow type

> Substrate

19 > Mean width

> Mean depth

Theseare combined with datafrom maps,including:

> Gradient(m km")

> Catchmentarea (km -2)

> Altitude (m OD)

> Distanceto mouth(km)

The model is calibratedon the basisof theseobserved data and can then generateestimates of fish populations from measurementsof physical habitat variables alone. As with

RIVPACS, a comparison of model predictions with population estimates made by electrofishingmay be usedto detectanomalies in observedecological data which may be attributableto impactingfactors. The modelswere producedfrom a data set of 224 sites on II catchmentsthroughout Wales. The siteswere screenedcarefully to avoid any which were subject to identifiable environmentalimpacts. Models were developedfor four categoriesof salmonid:trout 0+, trout >0+, salmon0+ and salmon>0+. The predicted densitiesrepresent an index of site expectation. This is not to be confusedwith carrying capacitywhich may be rarely reachedin sites receivingonly naturalrecruitment, because density-independenteffects keep the populationat lower levels (Milner et aL, 1998).The likelihood that many sitesused in habitatmodelling have populationdensities below their carryingcapacity is oneof the principal constraintson the performanceof habitatmodels.

Chapter5 attemptsto build upon existing models by using site specific and broadscale catchmentcharacteristics to investigatetheir relative importancein determiningAtlantic salmonfry numberswithin the Foyle system.

20 Chapter 2. Competing modes of exploitation and their effects on spawning success in Atlantic salmon (Salmo salar L.) in the Foyle Catchment, Ireland

Z1 Introduction

Conflict arising between resource user groups exists in many differing situations around the world today (Quizilbash, 2001). With regard to animal populations, this can be an argument over resource allocation or concerns over the conservation status of the exploited population (Claytor, 2000). Many of these have in the past been pushed to the edge of extinction (blue whale (Baleanoptera musculus) (Clark, 1973)) or over it (Great Auk

(Pinguinus (Alca) impennis); PassengerPigeon (Ectopistes migratorius) (Newton, 1998)) through over-exploitation. It is the balancing of these arguments by managers which science strives to inform.

Amongst the animal species deemed to be of high conservation value and named in statutory instruments requiring protection, the Atlantic salmon (Salmo salar L. ) is unique

in the degree to which it is exploited across its range in . Although designated as a

species whose conservation requires the establishment of Special Areas of Conservation

(SACs) under the European Union Habitats and Species Directive (Anon., 1992), exploitation through commercial high seas netting, coastal and estuarine netting and sport

angling is high. For example, the global harvest of this species in 1999 was ca 2,200

tonnes, the vast majority of this catch being made in North Eastern Atlantic waters (ca

2,000t) (Anon., 2001).

21 There have beengrowing efforts to reducethe impact of interceptorycommercial fishing on stocksso that individual populationswhich may be over exploitedare protected(Anon.,

1996).As well as conflict betweenexploitation and conservationthere is also scopefor conflict betweenthe multiple usersof a resource.There is a generallyaccepted belief that

commercial interceptionnetting of migrant salmon returning to the natal streams,by

trapping,gill or seinenetting in coastalwaters, estuaries and rivers, reducesthe potential

for exploitation by riverine sport angling (Johnston,2002; Anon., 2003b). Basedon the

presumptionthat the greatesteconomic value is derivedfrom sport angling,there has been

considerablepressure to close, or at least reduce, commercial fisheries to allow sport

angling to flourish (Whitehead,2003; Anon., 2003). As a result, interceptionnet and

trapping fisheries have been reducedconsiderably over the last 20 years or so (Anon.

2001). However,despite this widely held belief, there is a paucity of data which shows

significant benefits deriving for recreationalangling from the closure of a commercial

fishery (Jensenet aL 1999)and what datado exist may not provide a clear cut picture.For

example,Bowker et aL (1998) found someevidence to suggestthat diminishedestuarine

netting exploitationhad a beneficial impact on salmonangling successon the River Usk,

England, while with the closure of the marine Norwegian drift net fishery Jensenet al

(1999) concludedthat this contributedto increasedcatches of grilse and smaller 2 sea-

winter salmon in freshwater in Norway. McKinnell & Karlstrom. (1999) and

Romakkaniemi et aL (2003) found some evidence that reduction in commercial

exploitationcontributed to increasedstocks in manyBaltic rivers. A buyoutof commercial

netting stationsin Iceland by angling interestswas shownto have positive effects on the

sport angling catchesin the local rivers (Einarsson& Gudbergsson,2003). In contrast

however,Shearer (1992) found that following net closures,rod catchesdid not necessarily

improveon the AberdeenshireDee in Scotland.

22 This chapterexamines the potentialconflict betweenthe requirementsof managementfor protection of this species and exploitation (commercial net and trap fisheries and recreationalangling for sport) in one population,the Foyle system. Within this catchment area there has been long-term monitoring of catches of salmon from all forms of exploitation, in addition to the collection of an exploitation-independentmeasure of populationsize.

Specifically two hypothesesare tested:

1. Commercialnetting and trapping of Atlantic salmon in the Foyle systemsignificantly

impactsupon sportangling, by depressingangling catches.

2. Total exploitation (from all sources)has a significant impact upon the ultimate size of

the spawningpopulation.

Z2 Materials and Methods

2.2.1 Study Area

The River Foyle system drains ca 4,500km2of the northwest of the island of Ireland

(Figure 1.1), dischargingnorthwards through the estuary into the northern

Atlantic Ocean.

This systemsupports a significantpopulation of Atlantic salmonfor which there is a long

history of commercialnet and trap fishing and sport angling stretchingback over several

hundred years (Foyle Fisheries Commission, 1953; Elson and Tuomi, 1975; Foyle

23 FisheriesCommission, 1996). Total catch in the Foyle areain 1999was approximately60 tonnes,representing 11 % of the total catchfor the islandof Ireland(Anon. 2000).

The LoughsAgency of the Foyle, Carlingford and Irish Lights Commissionis responsible for the protection,conservation, improvement and developmentof the salmonidfisheries within the Foyle and Carlingford catchmentareas. As part of theseresponsibilities, the agencyand its predecessorthe Foyle FisheriesCommission, have collected data on catches and populationsize on an annualbasis since 1952. Thesedata are describedbelow.

2.2.2 Exploitation catch data

Drift nets

Drift nets are gill nets deployedclose to the water surfacein the sealough and up to 12

miles out from the north coast. At presenta maximum of 112 licencescan be issued

annually. In the past,the numbersissued ranged from 39 in 1952to a peakof 139in 1962.

The nets utilised at seaare 1500in long, with a depth of 45 meshsand a meshsize of 63

mm knot-to-knot. Those used in the Lough are 900 in long but otherwise of similar

dimensions. The drift-netting seasoncurrently commenceson 15th Juneand endson 3 1"

July and nets are fished for four days each week betweenthe hours of 6 am and 6 pm.

Licenceholders have been required to makeannual returns as a condition of the grant of a

licencesince 1952.

Draft Nets

Draft netsare seinenets which am deployedin-river in the main River Foyle and the River

Roe at defined netting stations.They are currently used from 15'h Juneto 31" July, for

four dayseach week over the 24-hourperiod. Tidal movementand water conditionslimit

24 their use. On average50 licenceswere issuedannually over the last 5 years(Wysner, pers. comm.) However,this hasfluctuated between a maximumof 524 in 1959to a minimum of

43 in 1999. There is a statutoryrequirement for the usersof thesenets to make catch retums.

Stake Nets

Stake netting was a commercialfishery run by the Foyle FisheriesCommission between

1952and 1988. A barrier net stretchedat right anglesto the shorelineled to a box trap

wherethe salmonwere caught. They werefished in the lower River Foyle and the estuary.

Thesewere temporary fishing stations,which were installedannually. The Foyle Fisheries

Commissioncollected annual catch records for the durationof this fishery.

Bag Nets

Bagnets were of a similarconstruction to thestake nets but wereleft in placepermanently.

Onelicence was issued annually to be fishedin theLough Foyle estuary, although id 1988

anadditional licence was issued. Data for thesenets are available from 1964to 1990when

operationsceased. The Foyle FisheriesCommission collected annual catch records for the

full periodof their use.

Catch Effort of Commercial Net and Trap Fisheries

The effort employedin all commercialfisheries in the Foyle systemhas changed over the

years.The seasonis considerablyshorter now than previously (currently 15 June to 31

July, cf. March to Septemberin 1952)and the numberof hours where fishing is allowed

each week has been reduced. There has also been a shift in emphasisbetween the

fisheries, with the closure of some methodsof exploitation and a shift in the relative

25 importance of draft and drift net use. Over the years the actual returns from the commercialnets have beengood and would appearto tie in with field staff observations

(R. Wysner,pers. comm.). For the purposesof analysispresented here no correctionhas beenmade for changesin fishing effort as it is the absolutenumber of fish removedthat is most likely to impactupon recreationalcatches and population size.

SportAngling ExploitationReturns

It is a statutory requirement that sport angling licence holders make an annual return of

their catch in the Foyle and Carlingford areas.The returns from recreational angling are, in

general, low (Small, 1991); this is certainly also true of the Foyle system. Thus, the sport

angling catch return cannot be regarded as an accurate measure of total exploitation by

recreational anglers. Therefore, to estimate the actual sport angling catches a correction

factor was applied to the data (Small, 1991). This correction allowed for the non-reporting

of fish caught when anglers made a return and is as follows ((0.3/(percentage return of

licences/100))+0.7) this produces a raising factor which is applied to the actual number of

fish declared.

2.2.3 Salmon population size estimates

Redd Counts

Since 1952the Foyle FisheriesCommission (and subsequentlythe Loughs Agency) field

staff have made counts of redds (salmon nests) during and immediately following the

salmonspawning period in approximately260 zoneswithin the catchmentannually. This

is only a partial count of all redds within the Foyle catchment.The accuracyof counting

can be subjectto environmentalconditions such as high water flows, which can obstruct

26 proper assessmentthrough poor visibility, or flattening of reddsmaking them difficult to see. However, as a method of long-term population monitoring, it is recognisedas a useful tool (Elson & Tuomi, 1975;Hay, 1984;Isaak, et aL, 2003).Highly accuratedata on spawning population size for one year, 1999 showed that counts at these 260 sites representedI I% of the total reddsbased on the percentageof spawningareas available to fish within the Foyle catchmentcalculated from instrearnhabitat surveys (Loughs Agency, unpublisheddata). Thus to estimatethe absolutenumber of reddseach year in the Foyle catchment,a correctionfactor of 9.1 was appliedto redd countsfrom all years.Each redd was takenas representinga singlefemale salmon (Hay, 1987). In order to thereforeget an overall population size an additional 40% was addedto incorporatethe male conponent

(Wray, pers.comm. ).

2.2.4 Analysis

Catchesand redd count estimateswere assembledby calender year. The data were analysedusing the statistical software packageSPSS 10.0. Pearson'scorrelations and linear regressionswere usedto look for trends in relationshipsbetween exploitation and subsequentspawning population size. To avoid Type I statisticalerrors, the acceptable probability when using multiple tests was correctedusing a Bonferroni correction.For clarity the correctedprobability equivalentsare presentedhere where appropriate.

2.3 Results

Analysis of the 49-yeardataset showed considerable variation in the yield from the various fonns of exploitation.

27 The 49-year meancatch in the drift nets was 26,753 individual salmonbut varied from

2,347 in 1952to 65,654 in 1983(Figure 2.1). Reporteddraft net catcheswere generally lower than thoseof drift net catchesafter the mid 1970's,but the draft nets havea higher overall mean catch of 30,928. The highest reported annual catch was also by draft net,being 82,106 in 1962 while the lowest was 5,434 in 1991. The Foyle Fisheries

Commissionstake nets operatedfrom 1952to 1988,during which period they comprised on average13% of the total commercialnetting and trappingcatches. The bag nets,which were operatedbetween 1964 and 1990,averaged 1% of the total commercialnet and trap catches.

The sport angling fishery reportedyield over the 49 year run averaged1,443 with a range of 379 (1984)to 5,100(1965).

28 a) C14 Sport Angling catches (no. of salmon) / Redds 0000 0 C) 0 LO 0 LO C) N C\j 0

C) C)

rj) 0) CM Q) 1. Nd 4a) If oa CL U) ...... r,- OD

CM 00 (D

N- N- 0) i-

>-

04 t- o)

r,-

IV

CM (0

U I ozz:::- .. 4

C) N U) CY)

C\j LO CY)

CD CD 000 C:)

00 r,- (0 U) CY) EI (uowjes jo-ou) saq31eo lau UejCV4! jC1 Actual redd (nest)counts between 1952 and 2000 averaged5,383 with the lowestcount in

1999 of 640 and the highest in 1966 of 22,606. Thus the estimatedcorrected mean, minimum and maximum redd numbersfor the whole catchmentwere 81,647,9,607 and

342,858respectively.

2.3.5 Hypothesis 1: The interceptory commercial fishery for returning migrant salmon impacts upon the recreational angling fishery

If one form of exploitation were negatively impacting upon another then clearly the exploitationmethods conducted at earlier points of interceptionwill impact on successive forms of exploitation. Thus, one would predict that drift netting, being the first of the interceptionfisheries exploiting the returning migrant population,would impact on draft netting and the combination of drift and draft netting would impact upon recreational angling. If there were an impact of initial exploitation methodson later ones then one would expecta negativecorrelation between the catches.

Correlationanalysis was used to test for a negativerelationship in catchesbetween drift and draft netting in the Foyle system. Catchesfrom the drift net fishery were not correlatedwith equivalentannual returns from the draft net fishery over the period 1952to

2000 (Table2.1).

Similarly to test for an effect of all the commercialnet and trap fishery catchescombined upon the sport fishery, the annual draft and drift net returns from 1952 to 2000 were combinedwith the catchesfrom other commercialnet sources(stake nets from 1952 to

30 1988and bag nets from 1964to 1990),to producea measureof the total catchesfrom all commercialfisheries over the 49 year period. Thesedid not predict sport fishery catches (rable 2.1).

When the commercial (drift, draft, stake and bag net fishery catches)were compared individually with the sport fishery catches, no significant correlation relationships with the sportfishery returnswere found (rable 2.1).

31 (1q Ln (» (D Ln r- f2 Co N 0 C) Ln [ý C) 0) M 1- t M " . qt 1- (0 CM le CI) (0 r- Ict Ln le lq 0 Co v- LO 0 Co cm 0 CM 0 0

*t o 0 0 o- e

(7) 9 8 r- Lf) 0 N Le) 8 CD P CY) e CY r- 0 cr) (» lit > li t .u uý cý ce cý r- C) Co 0 0 (D E E 0

CX3 ciý CMl e vLO - N ei s> N 8 .2 c\I r-.:cý cm 2 ei e

C» cy r- r_ 0 to 0 * N 0) Co 0 cr) to 0 cli 0 Cf) r_ m cý Ce

CY 0 rli (0 0 0 .

(0 cý

(D , -k '0 (D

CY

0 0 0 0 0 (D .0 iu 4ma) (D 4Z90) 0 0 0 (D 0) c) 0 (D C) (D 0 0) 0 = (i Q = ci CD 0 0 0 0 0 P) cý U) U) c\I 2 CM ci 0A (0 ce cý (0 Co a z9 a) a) 6 0tu a tu(D a Q)cu 6 w0 Z! CL U5 Z CL U5 Z CL U5 Z CLu5 Z CL u5 Z CL ý5 Z Cl cm . . . rn

0 ZU E E 0 <

E-4 Co EL 0 c] m Coig 0 C) F- Co , 2.3.6 Hypothesis 2: Exploitation negatively Impacts upon spawning population size

To test the hypothesis that exploitation is having a significant negative impact on that population, a measure of the annual exploitation was derived by combining the catch returns from all fisheries (commercial netting, trapping and sport angling). This was used to test for a relationship with an indicator of breeding population size that was independent of exploitation; the number of redds (nests).

Annual corrected redd counts were significantly but positively correlated with the total annual catch returns of all forms of commercial exploitation combined (Table 2.2). The total annual catch (including the sport fishery) also significantly predicted corrected annual redd counts in a linear regression model (FI,47=43.85; p<0.001; r2=0.47 - Figure 2.2).

Looking at each exploitation method for which there are a full 49 years data, salmon population size, as determined by corrected redd counts, was also significantly and positively correlated with drift netting, draft netting, stake nets, bag nets and total catch

(Table 2.2). When corrected using Bonferroni drift netting and bag netting were not significant.

One possible explanation for the significant positive relationships between different exploitation types, and between exploitation and population size, is that years with relatively large returning migrant populations are swamping any negative impacts. To explore this further, the same relationships between exploitation rates with population size were examined, at high and low population levels, separately.

33 It en

10 9

cc

'o ,02

0 0 cq 0 !ý 9)

"' Ui § r" q§ -M§- (), Ici 2 k§-. g§t,- E E 0

N (7) ý 9§ § CO - It . : C4 rIt

ID

V) I I

GOD - "0 I.,

ca

e

251c'1 r c C r_ C c c r_ 0 0 0 0 0 .0 .2 .2 .0 . . .

0 0 i a cts C ca C c r_ C C (D c cm 4 4 rz 6 caa) d, cis0) Ch 0Cli ca Ch (1) ca 6 cc(D b (1)ma ca CLco z a. 05 z M Fn Z C'Li5z CL05 Z COL)Co z (L rb z CL.Fn Z ('L ih z

0 i2 cc, b E 0 t 8 8 "0 8.0 s 9 20 g0a I Co' . , -

E-4 Figure 2.2 Regression of total annual catch of salmon in the Foyle area and corrected redd counts.

250000

200000

150000

100000

0 0 20,000 40,000 60,000 80,000 100,000 120,000 140,000 160,000 Total Commercial salmon catch (no. )

In years when the sport fishing catch returns were below the mean for the period 1952 to

2000, there was no significant correlation between these catches and any of the commercial

catches or corrected redd counts. In years when sport-angling catches were high, i. e.

exceeding the long-term mean, no significant correlations were found.

When the corrected redd counts were less or greater than the mean of the 49-year period

(81,647) there was no significant correlation with total commercial catch or sport angling

or all catchescombined.

35 To derive a measure of commercial netting catches which was free from the effects of population size, commercial catch was linearly regressed on total population size and the

residuals extracted. As auto-correlation artificially increased the significance of the

regression model and adjusted its direction it is therefore impossible to use this to make

any predictions. No interpretation is therefore put on the significance of this model, rather

it is a means to develop a straight line through the data and derive the residuals for further

analysis. These residuals significantly predicted variation in sport angling catches in a

linear regression (FI,47--210.69; p<0.002; r2= 0.168 - Figure 2.3) with sport angling catch

declining with increasing commercial catches independent of population size. Probability

analysis showed, that when commercial netting catches were less than expected for the

returning population size (i. e. below a0 residual), there was a 2.06 times chance that the

long anglers' catch would exceed the term mean (4,321 fish) (Probability analysis -

Relative RiskJRRJ = 2.06 with 95% confidence limits 1.24 to 3.41: X2 = 6.89; p_0.009

with a Yates Correction) (Lachin, 2000).

A similar approachwas used to look in greaterdetail at the impactof total catch(including

commercialand sportfishing) on the subsequentspawning population or redd counts. The

total catch was regressedagainst total populationsize and the residualsextracted. These

were thenplotted against the reddcounts. The relationshipwas not significant).

36 Figure 2.3 Commercial catch residuals regressedon corrected sport angling catch.

0c

Q ro

C

C I

-60000 . 50000 -40000 -30000 -20000 -10000 0 10000 20000 300DO 40000 Commercial catch. residuals

2.4 Discussion

It is axiomatic that, were there no coastal and tidal water interceptory fisheries for migrant

salmon returning to the Foyle catchment, there would be more salmon available for

exploitation by sport anglers. As the commercial fisheries intercept the returning

population before it becomes available to the sport anglers it is commonly perceived that

such an impact does exist and results in depressedangling catches. The availability of high

quality data from a very long-terrn monitoring programme has enabled a pragmatic

approach to assessmentof the impact of one exploitation method on another. Thus, here

the hypothesis was tested that commercial netting and trapping catches in the Foyle system

have a significant negative impact upon recreational angling catches. Analysis of the data

37 would suggestthat this doesnot occur in the Foyle area. Data from the drift net fishery, the first interceptionexploitation of returning migrants, showsno evidenceof negative impact on the in-river draft net fishery or upon upstreamsport angling leading to the tentativeconclusion that it is populationsize that controlscatches by anglers. Significant positive relationshipsalso exist betweencommercial netting and trapping catches. The analysisof the residualswith sportangling catches allowed a more detailedexamination of the interactionbetween two competingforms of exploitation. The result of this was,when commercialcatches deviate from thosepredicted by populationsize, theseresiduals were significant predictorsof the sport fishery, explaining 19% of the observedvariance. In addition,there was a reasonableprobability in yearsof low commercialcatches when there were strong cohorts of returning adult migrants,that the sport fishery would benefit by higher catchesof salmon. Thus the hypothesisthat commercialinterception netting and trapping of Atlantic salmon reducessporting angling catchesin the Foyle catchmentis supportedby the data. However, this should be qualified, as the amount of variance explainedby the commercialnetting returns is relatively low leaving 83% of variance unexplained.

The secondhypothesis that exploitation from all sources(sport angling and commercial netting and trapping) significantly impact upon the ult imate size of the spawning population,is not supportedby the data. As all exploitationof salmontakes place prior to the spawning period in any year, logically, exploitation rates in any one year may be expectedto depressthe spawningpopulation size. Strong positive correlationsbetween annual exploitation and ultimate spawning population size over a 49 year period superficially suggeststhat, at the levels of exploitation practiced in this catchment, exploitation does not have a significant impact on spawning population size. Further

38 analysis of the data did not highlight any significant effects. and the most likely explanationfor positivecorrelations between exploitation and spawningescapement is that yearclass strength is the determiningfactor.

Other studies have observed that many animal populations can sustain very high levels of exploitation without any apparent adverse impact on overall population size, e.g. Laurian et aL (2000) in their work on moose populations in Quebec, Canada found that the species could sustain high levels of exploitation and had a certain adaptability which allowed it to maintain high productivity despite the intensive harvesting. Also, Bosch et aL (2000) who looked at the effects of culling on gulls and Frederiksen et aL (2001) who examined culling and its impact on cormorants both found that the populations could sustain very high levels of harvesting.

With regard to Foyle salmon one explanationfor these findings is that years of high abundanceare maskingthe effects of exploitation in yearsof lower numbersof returning salmon. To test this yearsof low sport angling catch and spawningpopulation size (redd counts) were examined.No significant negative relationshipswere found, this would suggestthat even in yearsof lower abundance,the commercialnet fisherieswere still not having a negative effect on rod catches and that total exploitation (commercial and sporting)rates were not havinga negativeeffect on subsequentspawning escapement.

Thus it is concludedthat in the Foyle catchment,although commercial fishing has been shown to have a significant effect on sport angling catchesthere still remains 83% of variance unexplained,therefore the year class strength of returning migrants is the principal modulatorof variation in the commercialnet catches,sport angling catchesand

39 escapement,and that this effect over-rides any potential conflicting impacts of one exploitation type on anotherand of total exploitation on spawningpopulation size. It is unclear how these findings might apply to populationselsewhere. The Foyle system supportsa highly abundantpopulation comparedwith other EuropeanAtlantic salmon rivers (five year running averagecommercial catch between30,000-35,000 salmon) so it may be able to support multiple exploitation methodsthat other systemsmay not. In addition the strongmanagement structure in the Foyle systemhas resultedin tight control of exploitationin yearsof low abundanceso protectingstocks during theseperiods. Thus, in the Foyle area,the questionregarding cessation of netting may be regardedlargely not as a conservationissue, but ratherone of resourceallocation.

40 Chapter 3. Life stage specific, stochastic environmental effects, overlay density-dependent filial cohort strength effects in an Atlantic salmon (Salmo salar L.) population from Ireland

3.1 Introduction

All animal populationsare subject to controlling mechanisms,which limit their size.

However the relative roles of density-dependentand density-independentfactors in determiningultimate population size have long been controversial(Elliott, 1985,1994;

Sinclair, 1989; Newton, 1998).It is now widely acceptedthat for populationsto persist over time, at least one negativedensity dependentmechanism (where the probability of survivorshipdecreases as populationsize increases)must operate(Elliott, 2001).In nature, ultimate populationsize is unlikely to be determinedsolely by negativedensity dependent regulatorymechanisms but is most likely to be the result of complex interactionbetween density dependentand density independentfactors (reviews in Newton, 1998; Elliott

2001). Elliott (1994) in long-term studieson trout (Salmo trutta L. ) showedthat both density-dependentand density-independentfactors contributed to ultimate populationsize but their relative effects varied betweenhabitats. Studies such as those by Whittaker

(1971) and Newton and Marquiss (1986) have demonstratedthat populations may be regulatedby density independentfactors in stableenvironments but under lessfavourable conditions,density independent factors can assumemuch greater importance.

The evidencefrom a significant number of studiesover a broad range of speciesis that factorsmodulating animal population size do not operateat all times over the life-cycle but may act disproportionatelyat one or more ontogeneticor life-cycle events(Achord et aL,

2003; Langeland & Pedersen,2000). In red grouse (Lagopus lagopus) for example,

41 negativedensity dependance has been shown to occur during autumnwhen territoriality is at its height (Jenkins,1963). Similarly in oystercatchers(Haematopus ostralegus) density dependentregulation occurred during spring when densitiesincreased during the breeding period(Harris, 1970).

In a review of the literature Sinclair (1989) showedevidence of "populationbottlenecks", i.e. life stagespecific, negativedensity-dependent effects, on populationsize in a number of populationsof insects,fish, birds and mammals. Identification of life-stage-specific density independenteffects on ultimate population size has receivedless attention than density dependentfactors, possibly becauseby their nature they are more difficult to detect. However, some studies have shown such effects. In a study of an Australian peregrine falcon (Falco peregrinus) population, Olsen & Olsen (1989) demonstrated environmental(flooding) density-independenteffects during nesting on egg mortality. In sea-trout(Salmo trutta) in Black Brows Beck, Elliott and co-workers(Elliott et al. 1997) showedsignificantly depressedpopulation size as the result of summerdroughts and the effect was life-stagedependent while a numberof studieshave highlighted the potentialfor exogenousenvironmental impacts on a number of salmonidspecies (Einuin et al., 2003;

Hill et al., 2003;Azumaya & Ishida,2004).

As a result of the existanceof distinct phaseshifts in the life cycle and periods during

which abundancechanges, the Atlantic salmon(Salmo salar) is a specieswith the potential for populationbottlenecks (Elliott 1994;2001).

Here for the Foyle system, Ireland, several long-term datasetson returning migrant

Atlantic salmon population size are combined with environmentaldatasets to test two

42 hypothesesrelated to the control of ultimate population size. It is postulatedthat this population is primarily regulatedby density dependentfactors, but that environmentally

inducedeffects operatingat specific life stages,which affect the magnitudeof changein populationsize during populationbottlenecks, overlie this effect.

3.2 Materials and Methods

3.2.1 Study Area

The River Foyle systemdrains ca 4,500krnýof the northwestof the islandof Ireland (Fig.

1.1),discharging northwards through the Lough Foyle estuaryinto the northernAtlantic.

This system supports a significant population of Atlantic salmon for which there is a long

history of commercial net and trap fishing and sport angling extending back over several

hundred years (Foyle Fisheries Commission, 1953; Elson and Tuomi, 1975; Foyle

Fisheries Commission, 1996). Total catch in the Foyle area in 2002 was approximately

I 10 tonnes, representingca. 11% of the total catch for the island of Ireland (Anon. 2002).

The Loughs Agency of the Foyle, Carlingford and Irish Lights Commission and its

predecessorthe Foyle Fisheries Commission, have collected data on catches and

populationsize on an annualbasis since 1952. Thesedata are describedbelow.

3.2.2 Salmon Population Estimation

As the Foyle salmon population is exploited, an estimate of the returning migrant

population size has been maintained by the statutory body by combining data on

commercialnet catcheswith sport anglingcatches and the fishery escapement(that portion

43 of the populationthat remainsfollowing exploitation).The origins of a numberof elements of thesedata are described below.

Commercial Salmon Catches

Four forms of commercialexploitation of returningmigrant Atlantic salmonhave operated since1952 on the Foyle system.

Drift nets are gill nets deployedclose to the water surfacein the Foyle Estuary and in inshorewaters up to 12 miles out from the north coast. Drift nets have been in constant operationin the areasince 1952.

Draft netsare in-river seinenets which are fished in the main River Foyle andRiver Roe at definedpoints. Draft netshave been in constantoperation since 1952. Thereis a statutory requirement for the users of both drift and draft nets to make catch returns.

Stake nets are net traps run as a commercial fishery by the Foyle Fisheries Commission between 1952 and 1988. Annual catch records were collected by the Foyle Fisheries

Commissionfor the durationof this fishery.

Bag Nets are trap netswere of a similar constructionto the stakenets but were left in place permanently. Data are availablefor thesefrom 1964to 1990,the period spannedby the

fishery.

44 Catch Effort of Commercial Net Fisheries

The effort employed in these fisheries has changed over the years. The season is considerably shorter now (currently 15th June to 3l't July, cf March to September in 1952 the first year of the data set) (Loughs Agency, unpublished data) than previously and the number of hours fished each week has been reduced. There has also been a shift in emphasisbetween the fisheries,with the closure of some methodsof exploitation and a shift in the relative importanceof draft and drift net use. The drift nets particularly have become more efficient at catching salmon with increasing mechanisationand the availability of more effective netting materials. Over the yearsthe numberof returnsof catchdata from the commercialnets have been high and catchdata appear to broadlytie in with field staff observations(Wysner, pers. comm.). For the purposesof this paper no correction to catch data has been made for effort, as it is the absolutenumber of fish removedin the fisheriesthat reflect returningmigrant population size.

SportAngling ExploitationReturns

Despiteit beinga statutoryrequirement that sportangling licence holders make an annual catchreturn, the numberof returns,as a percentageof total licencessold was highly variablebetween 1952 and 2000. This is typicalof previouslypublished studies (Small,

1991). To determineannual catch rate by therecreational fishery a correctionfactor was usedwhich wascalculated using a techniquedescribed by Small(1991) to determinea realisticmeasure of anglingcatch. This correctionallowed for the non-reportingof fish caught when anglersmade a return and is as follows ((0.3/(percentagereturn of licences/100))+0.7) this producesa raisingfactor which is appliedto theactual number of fish declared.

45 Spawning Population Estimates - Redd Counts

Since 1952the Foyle FisheriesCommission (and subsequentlythe Loughs Agency) field staff have made counts of redds (salmon nests) during and immediately following the salmonspawning period in approximately260 zoneswithin the catchmentannually. This is only a partial count of all reddswithin the Foyle catchment.The accuracyof counting can be subjectto environmentalconditions such as high flows, which can obstructproper counting through poor visibility, or flattening of redds making them difficult to see.

However,as a methodof long-termpopulation monitoring, it is recognisedas a useful tool

(Elson & Tuomi, 1975;Hay, 1984;Isaak, et aL, 2003). Highly accuratedata on spawning populationsize for one year (1999) showedthat countsat these260 sitesrepresented I I% of the total reddsbased on the percentageof spawningareas available to fish within the

Foyle catchmentcalculated from instrearn habitat surveys (unpublisheddata, Loughs

Agency).Thus to estimatethe absolutenumber of reddseach year in the Foyle catchment, a correctionfactor of 9.1 was appliedto redd countsfrom all years.Each redd wastaken as representinga singlefemale salmon (Hay, 1987).

3.2.3 Population Structure

Scaleanalysis of 813 migrant adult salmonreturning to the Foyle from 1968,1969 and

1970 showed that the population comprised93.8% 1-sea-wintersalmon (fish that had remained at sea for only one winter) (Anon., 1969; 1970; 1971). The age of metamorphosisfrom the freshwaterto the seawaterphase (smolt age)was also determined from thesefish. On averageover theseyears 1% left the river at one year old, 92% at two

46 yearsof age and 7% after three yearsin fresh water. Scaleanalysis on a sampleof 81 salmon in 1998 (W. Crozier, pers. comm) found that 97.5% were I-sea-winter and 13% hada smolt ageof I+yrs; 84% hada smolt ageof 2+, and 3% of 3+.

For subsequentanalysis, all fish were assumedto be I sea-winterfish but variation in smolt ageswere appliedto the data so that the recruitswere correctly apportionedto their respectiveparental year classby the following method. From adult scaleanalysis the ratio of smolts ageswere known for the years 1952to present(Loughs Agency, unpublished data). This information was used to apportion the recruits of spawning fish to the populationin their year of return i.e. fish spawnedin 1952would return as adults in 1955,

1956or 1957and wereapportioned accordingly.

To comparelike-with-like all populationdata (commercial catch, recreational fisheries and redd counts) were convertedto an "egg number equivalene'. As fecundity is related to body size, a fecundity of 1,430 eggs kg-1 of fish (Shearer 1992) was used. The mean weight of salmon was taken from the commercial fisheries for each year used in the analysis. In addition, as the sex ratio is typically skewed in adult migrant salmon populations,a sex ratio for the population of 60:40 (female to male) (Loughs Agency, unpublisheddata) was usedto adjustegg depositionnumber for datafrom the commercial andrecreational fisheries.

3.2.4 Data Analysis

To examine the hypothesesthat the Foyle Atlantic salmon population is regulatedby density-dependentfactors, the relationshipbetween parent and progeny population size,

(total eggequivalent derived as above)was examined.

47 Linear regressionsand curvi-linear stock-recruitmentmodels (derived from the literature)

were testedto find the model which bestexplained the variancein recruit populationsize.

To test for additional environmentallyinfluenced modulatorsof population size, during

key life cycle stages (potential population bottlenecks),a number of environmental

parameterswere usedto predict variation in residualsextracted from the stock-recruitment

model.

Ufe-HistoryStages

Eight specificlife history stageswere. recognised (Figure 3.1):

48 0 z

C 0

421

(D tb to

0 E V)as ff)

ýA4 Spawning and egg incubation

The first stage in the salmon's life cycle examined was from November in year x to

January year x+l, when adult fish are engaged in spawning and eggs are developing in the redds. Four variables were used to test for a relationship between environmental conditions during this period (Nov-Jan) and the size of the returning migrant cohort. These were: rainfall, water flow, air temperature, and the North Atlantic Oscillation Index

(NAOI) (see below).

LarvaelAlevin Emergence

Subsequent to this, variation in environmental conditions during the period of alevin emergencefrom the redd; - February year x+1 to March year x+l, was tested for its effect on population size. Environmental variables used were: rainfall, air temperature and water flow, during this period.

Fry establishment - Year 0+ Summer

Following emergence, fry establish territories in the streams during their first summer

(May year x+I to August year x+l). Three variables were tested for significant effects on population size during this period, these were rainfall, air temperature and water flow.

Year 0+ Overwintering

Overwinteringof fry was examinedby looking for significant effects during the period

Novemberin year x+1 to Januaryyear x+2. The variablestested for effectson population sizeduring this periodwere rainfall, air temperature,water flow and the NAOI.

50 YearI+ Summersurvival

The effectsof the environmenton one year old parr summersurvival, coveringthe period

May year x+2 to August year x+2 were tested using three variables: rainfall, air temperatureand waterflow.

Year 1+ Overwintering

Overwintering parr during the period November year x+2 to January year x+3, were tested for effects on the overwintering juvenile salmon population size using the environmental variables rainfall, air temperature, water flow and the NAOI

Smolt migration

The smolt migration period May - June year x+3 was testedagainst the environmental variablesrainfall, air temperatureand waterflow.

Marine Survival year x+4

The NAOI and sea-surface temperature anomalies north of Iceland were tested for significant survival effects during the period when the salmon were in their first winter at sea, November year x+3 to February x+4.

51 EnvironmentalVariables

The environmentalvariables used are explained in greaterdetail below: -

Air Temperature,Local Sea-surface Temperature and Precipitation Data

This information was obtained from the Meteorological Office in Dublin. These

environmental data were collected at Malin Head, Co. , which is at the entrance to

Lough Foyle. Air temperature and precipitation data were available from May 1955 to

present, while sea-surfacetemperature data were available from May 1958. The data used

were an average for each month available.

Waterflow

Thesedata were derived from a hydrometricstation on the CamowenRiver at ,Co.

Tyrone, which is operatedby the Departmentof the Environmentfor N Ireland. For the

analysisan averagemonthly flow rate was used. The data were availablefrom January

1975to December2000.

The North Atlantic Oscillation Index (NAOI)

The NAOI is an atmospheric phenomenon, which is measured as the difference in air

pressure between the Azores and Iceland (Hurrell, 1995; Hurrell, 2003; Gillett et al.,

2002). The winter index, used in this analysis, is calculated by taking the mean of the

index between December and the following March. This Winter NAOI Index then

referencesto the January of that year.

52 Winter sea-surface temperature anomaliesfOr Grimsey Island, Iceland

The period November x+3 to January x+4 was used to examine possible environmental effects of the marine environment. These data were acquired from the British Atmospheric

Data Centre (BADC) and are representativeof temperature change.

Analysis

Two statisticalapproaches were adoptedto examinethe variationin populationsize independentof densitydependent effects. Three parental-offspring population size models were tested(linear; Beverton & Holt; Ricker) using the softwarepackages SPSS and

Statistica.The model which best fitted the data and provided the greatest predictive ability waschosen for furtheranalysis. Environmental variables were then used as independent predictorsof residualsderived from this modelin regressionanalysis. However, as it is highlypossible that any environmental effect on populationvariation may not be linearor may showsignificant threshold effects (see Elliott, 1997),population residuals in years with high for examplewater flow (upper30 percentile),low water flow (lower 30 percentile)and intermediate water flow (mid40 percentile)were compared using ANOVA with a posthoc Tukey test pair-wise group comparison. This techniquewas used with all of the environmentalvariables. To avoid Type I statisticalerrors, the acceptable probabilitywhen using multiple tests was correctedusing a Bonferronicorrection. For claritythe corrected probability equivalents are presented here where appropriate.

53 3.3 Results

3.3.1 Density Dependent Regulation

Of the three models examined relating parental size to the number of offspring recruited into the population, linear regression explained the lowest amount of variance (17.1%),

Beverton and Holt 27.2% and the Ricker model explained most variance at 31.2% (Table

3.1; Figure 3.2a; 3.2b; 3.2c).

Table 3.1 Comparison of predictive linear; Beverton & Holt and Ricker models.

Model r2 p n F

Linear 0.171 0.003 43 9.88

Model r, n constant parameter

Beverton & Holt 0.272 43 5.94 0.0005

Ricker 0.312 43 4.385 0.00017

This model suggests that maximum filial population size resulted from parental egg

deposits of around 500,000,000 eggs. Egg deposition in excess of this resulted in a low

rate of decline in filial recruitment to the population, deposition rates below optimal filial

recruitment declined more rapidly with decreasingegg deposition. Thus, both positive and

negative density dependent factors appear to modulate recruitment in this population. To

derive a measure of survivorship, i.e. population size independent of parental population,

residuals from the Ricker curve were derived. These were then used to examine the role of

environmental factors controlling variation in survivorship.

54 Figure 3.2a A linear model fitted to parental population size (egg equivalent *100,000) and filial population size (egg equivalent* 100,000).

18,000

16,000

14,000

12,000 g 10, ODO

8,000

LL 6,000

4,000

2,000

01 ODO 2000 3000 4000 50DO 6000 7000 8000 9DOO Parental (eggs*100,000)

55 Figure 3.2b A Beverton & Holt model fitted to parental population size (egg equivalent 100,000) and filial population size (egg equivalent* 100,000).

Model: v3=(constant*v2Y(1+(param*v2)) y=((5.9411483)*xy(1 +((0.0005041)lx)) 18000

16000

14000

12000

10000

8000 ÖJ3

6000

4000 .1 a'! 2000 11«

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 I 1 Parental (eggs*] 00,000)

Figure 3.2c A Ricker model fitted to parental population size (egg equivalent *100,000) and filial population size (egg equivalent* 100,000). Model: v3=constant*v2*exp(-param*v2) y=(4.3854143)*x*exp(-(0.0001741)*x)

16000

. . 12000 .

8000 (A

4000

n 0 1000 2000 3000 4000 5000 6000 7000 8000 9000 I I Parental (eggs* I 00, OW) 56 3.3.2 Life-stage specific - environmental effects on population size.

Spawning and egg incubation

Four variables,air temperature,water flow, rainfall and the NAOI, were usedto test for

correlationsbetween survivorship and environment. Water flow (positively) (Figure

3.3a) and air temperature(negatively) significantly predicted population size residuals

(Figure 3.3b). The two remaining variables did not significantly predict survivorship

(Table3.2).

Years with low air temperatures during the incubation period had significantly higher

survivorship than years with high air temperatures but there was no significant difference

with years of intermediate temperature and high or low temperature years in ANOVA

analysis (Table 3.3; Figure 3.4a). Years of low water flow had significantly lower

survivorship than years with high water flow during this period (Table 3.3; Figure 3.4b).

Survivorship in intermediate water flow years did not differ significantly from that of

either high or low water flow years. However, when corrected using Bonferroni these

were no longer significant.

To determinethe environmentalvariable most influencing spawningand egg incubation

survivorship two variables water flow, air temperature were entered into a stepwise

forward multiple regression.

Water flow was the most importantdeterminant of variability in survivorship,however, air

temperaturealso contributedsignificantly to explainingvariance, with thesetwo variables

aloneexplaining 54.9% of total variance(F2,19 = 13.177:p<0.001).

57 Figure 3.3 Regressionsof environmentalvariables during life-stage specific eventsand populationsize residualsderived from the Ricker parental-offspringcurves. a) water flow during spawning and incubation, p<0.001; b) air temperatureduring spawning and incubation,p<0.009; c) rainfall during larval/alevinemergence, P<0.029; d) North Atlantic Oscillation during 0+ winter, p<0.022;e) water flow during 0+ winter, p<0.012;f) North Atlantic Oscillationduring I+ winter, p<0.011; g) North Atlantic Oscillationduring winter marine feeding, p<0.008; h) sea surfacetemperature anomalies north of Iceland during winter marinefeeding, p<0.002. 39 2000- 30

0- low - t 7 67.OS le 67 670 6723 i000- 100047 67.05 6671 -eliz-74.672 67ý 5 -2000

3D00 -m

-4000 -m (cun»co) Wo" flow (eunwu) wäter now

10000 i om)- - 3b ww oow - 60oD 6=- 40W 4000 1 2W0 1 et 6 t ýo ;4 01 -ýI 45 :: 4W4 6 ep- 0,0 -10 - ý 0.1 o _ -400

Air tompendm (Fa Im hWt) NAO

10000 iooo- 8000 30 am- 39 6= *Wo

4W0 20W 0 ) 10 -20M1 -4000. -40W

-ein -ein NA O

io oo o-- 3h oow - 3d 8m. 6000- 6=- 0 40W 2 0 4000- .e16 - , 2000. 60 - ý* 1 t- 00-w -ýO jr4_wýOD; j loo 2w 300, -40M- iw 6000- ---- . - KAO BST(anonn394

58 Larvae/AlevinEmergence

Of thevariables tested rainfall, air temperatureand water flow in theupper catchment, only rainfall was shownto be significantduring this period,however, following correction usingBonferroni this was no longersignificant (Table 3.2; Figure 3.3c).

When population size residuals were compared in years with low, intermediate and high flow, the data hinted at an inverse relationshipbetween flow and survival (Table 3.2) althoughthe resultwas not significant.

The regressionof rainfall on survivorshipexplained 9.3% of the total variance(FI, 39

5.11; p<0.029).

Fry establishment- Year 0+ Summer

Threevariables (summerrainfall; air temperatureand waterflow) were regressedon survivorship. None of thesewere found to be significantpredictors of varianceor correlatedto the populationresiduals, although water flow duringthis periodwas nearly significant(Table 3.2).

ANOVA analysisof years of high, intermediateand low rainfall showedno significant differencesin survivorshipbetween these years (Table 3.3).

59 I s ON 0 Cf) a 0 0 -4 -4 -0 -. 4 0 - cf) N C14C14 N cn Nr 1* N IRT NT C4 't "I N NT IRT Itt Cq It IRT (14

C,

0 0 r.

4.42 "t N 00 N N %0 cn r- 0 - r- "0 rf) r- tn .. 4 N'o -4 ". 4 m Cf) C% 00 0 9 6 6 01 6 6 d d 6 6 6 al61 C26 06 . 8 r-: 8 1cc-1 C5 GA ce A ý A (7, 1ý0 en 00 -4 r, ýo 10 It Itt a ý cn C-4 cn Ci Cn C) C) q -4 . 0 cn . 6 6 c; C5 r.5 c? (3 o

- Cý M r- B r- ýo 10 (14 en cq cf) , 8 0 W) oo ýo W) W) ýo t W4 0 W)4 IR cl (7N I r-: C,; , C; ci wi r-: 6 C-i6 "0

110 C14 Cý ON * & 0 0ý tn r- I en :t 00 ON ;s ON -4 r- ON q ":cn eq "" ,rý: 7, ý6 00 r- IRT 00 IRcq cn C5 N - C14 ýocn -, 0 C? r, i en

T I I

"I ZIO

0 0 CD. 0 CI. 0 cl 0 E 2d E co 4 0 4-4 0 r- ý

rA

4.4 0 9) 4.4

1.4 b4

Qn cc Cd 1-4 Cd 1 l I cq en en en en en cq (1) W) IRT It Itt C14 It It C*4 C*4 RT cn

W)

jW ej 4r.1, Ll pw 14 14i4 0 a U

00 tf) i 0 ýoc f ) 1.0- 4 - 4 8 tn ýc00 1W) eq Ici " m r. l: %ro, d d (::ý 0 6

t4r) N 00 t- IT eq cl I., en -4 N -4 - en cf) -4 0 0 - 0 -. 4 c.4

00 %.o 00 00 ON tn C-1w! -1 C-1 0 0 W) dC'4 00 cq ". 4 r- C5 r-: C)I" "0 I I I I I I I I I I

0 rý 4R oo ý:t , 'n 8 A c 4 00 en 4 c 't C, ' 06 00 ýc1-4 C5cý 11 00 Cý0? 'T 00 -, 1-4

"Ci cl En 8E r. .19C = E 0 = 0 s P-.4 o .2 -S C) -< 4)

ý 101.

Cd

0

W

0

C14 0 I I 1 ON 'o 'o "D r- ýo C140 C" C', m Itt cn r- m r en M 00 -- '4 a 00 00 00 C#) ýo ýo 00 tlý ýo -4 ON

4 "t 0 t, ý.q t (7Ncn ýo m 0 N Ch V) en 00 "0 .2 -, 00 r- - Cý - 00N - 0 W'l Cý NT r- en en - -

1 C,"04 C',cn r- C,4 eq en ;S r- wl, 00 ýý ýo R - 00 0. 00 W) X? C "1*"-0 " 10 00-4 cf) ON 00 Cý

= &W Cl. r.

"0 1 I l I ON CD ""t --, -4 "' I wl o 110 eq 00 N 00 WI) ýo 00 W) 19t W ) r - r- CIO 0 C\ 6 r- c) 00 r- 4 C5 6 6 d 6 6 d d C) (=5 c o o

1- 00 -0 00 00 00 ON CN ON C., 00 CN Cý CD 0 V) Ci 't -t n n cf) cf) - cf) cf) -t - 't "t cli C14C14 N NI N C141clt I Clf cli I Clf CIT Cý N N r- 00 C\ 00 %,o 't en - en V*j 'r, W" cn C) - r r a 00 s tq (ý m - Cý (ON ý14 i cf) * ON00 W) 00 c 00 en - Cn 00 00 c cli Cý C5d d 6 Cý 6 6 ý6 cli d

-0 cp o 0 7; E E E tz = E C) o 4" C) jwo 0 o

P4 z 00"

0

r. 0

12

elý I N4 m cu \O

0 C14 00 WI) ýo ýc t- W, 00 en (7% ýo oo %.o C-A r- '9 ' ") 00 C,4 2ýý2 ,r- C14N . n. .

00ýo MC,4 ýo- enr- r-ýo en cnC) enIt W)I- C14 N It 10 - cc Go

10 l tn 00 It ýr tn"' oo;i 00 rn 0 in 00 0 en a ON 00 14 800 1"01 1* -. , a0 00 00

Ici ýc r- cq 't 00 r- en -. 4 tn - In cn ý9 Mo cn ýo 0 N N r- 6 C5C5 C5 6 6 6 C5 C;6

N cl C,4 C-4 C-4 cT Cf C4 cf Cq

Cý "t cn - M r- r- en r- r- W) CN 00 -4 c, N 0 N 1

3 L) 2. E 0 a P- ýu lu 12 12 d) CEo 46In ý C) N .21ý I ro 0 . < 2 w . is .ý4 04 < z ý: P.,:ý ý: ý 04-V) "01-41

cl 1.4

CIO)

0

L_ L_-- L :3 Figure 3.4 Mean residual population size values derived from the Ricker parental- offspring curvesfor years with high (upper 30 percentile);low (lower 30 percentile)and intermediate(remaining 40 percentile)values for environmentalvariables during the life- stagespecific periods separately for: a) air temperatureduring spawningand incubation;b) water flow during spawning and incubation; c) North Atlantic Oscillation during 0+ winter; d) water flow during 0+ winter; e) North Atlantic Oscillation during 1+ winter; f) rainfall during smolt migration;g) North Atlantic Oscillationduring winter marinefeeding. All figures are significant at p<0.05 but columns with identical lettering are not significantly different in post-hoctesting. Figures 4a and 4b were not significant after Bonferronicorrection.

4a 4e

15w- 2000 1000- 1128 1500 1804 6w- 1000 0--a 500 c AC 500 «Am 0 -0 . 1000 - -610 -5w A , 1500-. AB -1130 -468 . . 181361 -1000 -2000 -1500' low Intermediate high low Inteffneclate hIgh Air temperature NAOI

4b 41 2,366 25W 0 U2174 M-746 9.500 1500 11,500 600 000 -1125 B BW AB -326 c0A ; 1600 2260000, 25W I low intermediate high low ktermeclate high wetwIflow Raintaft

4C 4g

2500 25M 2000 Ai 2000 1500 1500 1000 1000 500 RAB 5w Ba 0 0 E-1328 -W c -500 A '500 A -387 -1000 -1000 -W -1500 1500 low intermediate high low Intermediate Ngh NAOI NAOI

4d

0- 9 »500- -1000. -719 -1115 ABC -1999 ABC 20D0 2500 low Intermediate high waliff now

64 Year 0+ OverWntering

Of the four variables examined (rainfall, air temperature, NAOI & water flow) only the

NAOI and water flow were found to significantly predict survivorship. The NAOI had an

inverse relationship (Figure 3.3d) while water flow was positively correlated with

survivorship (Table 3.2; Figure 3.3e). Correcting for multiple tests using Bonferroni

however meant that the NAOI was not a significant predictor.

Years with low, intermediateor high values for each of theseenvironmental variables

showedsignificant differencesin survivorshipfor the NAOI (Table 3.3; Figure 3.4c) but

not quite significant for water flow, althoughthe data hinted at a positive relationshipto

higherflow ratesin the uppercatchment (Table 3.3; Figure 3.4d).

To determinethe environmentalvariable most influencing 0+ winter survivorship,NAOI

and water flow were enteredinto a stepwiseforward multiple regression.Water flow was

found to be the most importantdeterminant explaining 23% of the total variance(F1,21

7.562;p<0.012). NAOI did not add significantlyto the model.

Year 1+ Summer survival

Summerrainfall, air temperatureand water flow effectson survival were ascertainedusing

regressionand ANOVA. None of thesevariables were found to havea significanteffect.

65 Year 1+ overwintering

Of four environmentalvariables (rainfall, air temperature,NAOI & water flow) examined

only the NAOI was found to be significantly inverselyrelated to survivorship(Table 3.2;

Figure 3.3f).

Comparing survivorship in years of low, intermediate and high values of the four

environmentalvariables showed that only survivorship in high, intermediateand low

NAOI differed significantly (Figure 3.4e). Post-hoctesting showedthat survivorship in

low NAOI years was significantly higher than that of intermediate years, but not

significantfrom yearsof high NAOI.

The NAOI explained12.5% of the total variance(FI, 42 = 7.157;p<0.01 1).

Smolt migration

Rainfall, air temperature, water flow and the sea surface temperature at Malin Head did not

significantly predict survivorship.

However, comparing years with low, intermediateand high values for rainfall using

ANOVA showedsignificant differencesin survivorship(Table 3.3 & Figure 3.4f). Post-

hoc testing showedthat survivorshipwas higher in yearswith a high rainfall at this time

comparedwith yearsof low or intermediaterainfall.

66 Marine Feeding- year x+4

The NAOI significantly predicted survivorship(Table 3.2; Figure 3.3g) (inverse relationship),as did the seasurface temperature anomalies north of Iceland (positive relationship)(Figure 3.3h).

Post-hoc testing showed that low NAOI years had significantly higher survivorship than yearswith intermediateor low NAOI (Figure3.4g).

To determinewhich environmentalvariables (NAOI & SST) best predictedsurvivorship they were enteredinto a stepwiseforward multiple regression. Sea surfacetemperature was the most importantdeterminant of variability in population size residualsexplaining

21.7% of total variance(FI, 34 = 10.718:p<0.002). NAOI did not add significantly to the model.

3.4 Discussion

The evidencepresented here supports the hypothesisthat density-dependentmortality is an importantunderlying mechanism controlling Atlantic salmonpopulation size in the River

Foyle catchment. Modelling the effect of parentalpopulation size on the filial population showeda significantrelationship, with the Ricker curve providing the best model fit to the data. The domednature of this curve suggeststhat negativedependence occurs at high densities.Elliott (1994) showedfor brown trout that this modelfits datafor other salmonid populationswhile other authorshave shownthis dome-shapedmodel fits Atlantic salmon

(Chadwick, 1982; 1985a; 1985b; Kennedy & Crozier, 1993;). However, in the study presentedhere these density-dependent effects only accountedfor 31% of the total annual

67 variability in population size. Clearly other factors also influence the population size.

Here, a life stage specific approachwas adopted to aid the identification of factors modulating the remaining variation in population size. Biologically important critical survivorshipperiods ("bottlenecks"sensu Elliott, 2002) were identified and broad scale environmentaldata usedas predictorsof survivorshipvariance to determinetheir relative influence on a comprehensivesuite of the critical life stage specific events faced by

Atlantic salmonduring ontogeny. This approachshows clear evidenceof environmentally inducedpopulation regulating effects at a numberof critical life stages.

During the spawningand incubationperiod, standardisedin this study betweenNovember and Januaryof the following year, it has been shownthat air temperatureand water flow are good predictorsin explainingresidual variation. Higher survivorshipis correlatedwith lower air temperaturesand higher water flows in the upper catchment. Approximately

54.9%of survivorshipvariation was explained at this life stage.It is unclearif theseresults are linked to the spawningperiod itself or the incubationperiod of the eggs. However,it is assumedthat water temperaturewill be a function of air temperaturealthough probably with some time delay. In previous studies it has been shown that a drop in water temperatureis requiredbefore spawningoccurs (Shearer,1992) and it is well recognised that egg incubationis controlleddirectly by water temperature,higher temperaturesleading to fasterdevelopment and emergence from the redd(Heggeberget, 1988).

On emerging from the redd, the alevin must establish a territory and this period is recognisedas a populationbottleneck in other salmonidspecies (Elliott, 1994;Jones et aL,

2003). High water during this time, when fish are first exposedto full stream flow conditions,can be particularly important in determiningultimate population size (Elliott,

68 1994). Consistentwith this, it was shown that rainfall explained9.3% of survivorship variationat this life cycle stage. Although water flow was not significantat this stage,this could be as a result of low samplesize as low rainfall predictedhigher survivorship.

Previous studies (Elliott, 1994; Elliott, 1997) have linked high periods of mortality to summerwhen water conditionscan havea severeimpact on productivity. This may occur through direct mortality or indirectly, for exampleby affecting growth ratesand therefore potential for survival at other crucial periodssuch as first entranceto the marine phaseof the salmonids'life cycle. In the study presentedhere, there was no direct evidencefor environmentalcontrol of survivorshipduring summerperiods, although survival in either year 1 (0+) or year2 (1+) was nearlypositively significantly relatedto waterflow.

In contrast,the periodswhich appearto influencesurvivorship in the Foyle catchmentare more prominentin winter. Survival for both 0+ and 1+ winter periodsis stronglylinked to the negativephase of the NAOI and in the first winter to increasedwater flow in the upper catchment. Thus environmentalfactors accountedfor 23% and 12.5% of survivorship during first and secondwinter periodsrespectively.

The final critical phase in the freshwater element of the salmons' life-cycle is that of its entrance to the marine environment as smolts. Here it has been shown that high rainfall greatly benefits survivorship during this period but interestingly the worst case scenario appears to be that of intermediate flows, very low flows having a much smaller negative effect. It has been shown previously in other studies that high water flows greatly facilitate movement of these fish downstream (Shearer, 1992; McCormick et aL, 1998; Byrne et aL,

2003) and that low flows may be detrimental to survival. However the impact of these

69 intermediateflows havenot been observedpreviously. One possibleexplanation for this finding is that in very low flows the fish are less likely to migrate actively downstream while in intermediatewater conditionsfish will attempt migration and thereforemay be moreprone to predation.

Survival once the salmonreach the seais known to be affectedby marine conditionsand thesehave been linked previouslyto the NAOI and also seasurface conditions for North

American stocks of salmon (Friedland et al., 1993; 2003a; 2003b). Very little similar information is available for Europeanpopulations although Martin and Mitchell (1985) linked sea surfacetemperatures north of Iceland to differing abundancesbetween single and multi-seawinter salmonand Friedlandet al (1998) linked a thermal habitat index of

10-13'C in the North Sea.to survival of Scottishand Norwegian stocks.

The NAOI may influencesurvival in a numberof ways, suchas affectingcurrents, salinity profiles and seasurface temperatures. These conditions may impact on salmondirectly, for instance it is known that salmon cease feeding at temperaturesbelow 3'C

(Shearer,1992), or indirectly by affecting prey or predator abundance. Sea surface temperaturenorth of Iceland appearsto be of particular importanceas this variable was found to explain 23% of variance. There may of coursebe other mechanismsor factors which are linked to theseconditions of which little is known at presentand which are affectingthe abundanceof the salmonduring this phaseof their life-cycle.

The Ricker density-dependentmodel explaineda proportionof the variancein recruitment, however, this is not to state that the mode of operation of environmentalfactors is completelydensity-independent. For example,in looking at the effects of water flow on

70 0+ and 1+ over wintering, higher water flows lead to greaternumbers of fish surviving.

This is possiblydue to the effect of greaterareas of habitatbeing available as a direct result

of the increasedflow. If so, this implies that theseenvironmental factors are operatingin a

density-dependentmanner as well as having somedensity-independent effects.

This study highlights the importanceof density-dependentregulation as a controlling

mechanismfor the salmonpopulation in the Foyle area. However,it has beenshown that

environmentalfactors also play an important role in determining ultimate returning

populationsize driving life-stageevents that may potentially act as populationbottlenecks.

It is concludedthat the life-stage specific environmentaleffects which are shown to be

operatingin this populationare likely to havesimilar effects in other salmonidpopulations,

and will contributeto the apparentstochastic variation in populationsize resultingfrom the

applicationof traditional stock-recruitmentmodels. The identification and quantification

of these effects, from this long-term dataset,should improve the predictive ability of

models by enabling the constructionof more sophisticatedmodels that combine parent

populationsize and environmental factor variance.

71 Chapter 4. The influence of broad scaleclimatic phenomenaon long- term trends in Atlantic salmon population size: an example from the River Foyle, Ireland. -

4.1 Introduction

Many animal populations are known to experience large-scale fluctuations in population size e.g. lemmings (Lemmus lemmus (L)) and voles (Clethrionomys rufocanus) in Northern Norway (Ekerholm et aL, 2001) and red grouse (Lagopus lagopus)throughout their range(Jenkins et aL, 1963;Newton, 1998). Many of these cycles are related to natural phenomena, operating independently of direct anthropogenicinfluence, for example,changes in the distribution and abundanceof the Pacific sardine(Sardinops caeruleus) and northern anchovy(Engraulis mordax) havebeen shown to be relatedto climate shifts in the north-eastPacific over a 70 year period (Rodrfguez-SAnchezet aL, 2002).

With regard to Atlantic salmon Martin and Mitchell (1985) showed correlations betweensea surface temperatures north of Icelandand the abundanceof I -sea-winter and multi-sea-winterfish returningto the , Scotland. Increasingsea surface temperaturesin the sub-Arctic were associatedwith larger numbers of adults returning as multi-sea-wintersalmon and fewer returning as I-sea-winter salmon.

They also showed that the averageweight of I-sea-winter salmon increasedwith populationsize of that group. Friedlandet aL (2000) also found linkagesbetween survival of salmonand sea-surfacetemperatures in the North seaarea.

72 Similar linkageshave been found for Pacific salmon(McFarlane et aL, 2000; Mueter et aL, 2002) and North American stocks of Atlantic salmon, where studies have correlatedmarine survival of Atlantic salmonto winter seasurface temperatures in the

Labradorsea (Friedland, 1998; Drinkwater, 2000; Reddin et at., 2004).

Previous studies have shown that the atmospheric phenomenon the North Atlantic

Oscillation is also related to stock abundance in Atlantic salmon (Friedland et aL,

2003; Beaugrand & Reid, 2003). This index can be used as a proxy for climate variation on a large scale. The North Atlantic Oscillation in winter (NAOI) has been shown to provide a good index of the dominant mode of winter climate variability in the North Atlantic region ranging from central North America to Europe and into much of Northern Asia. It is calculated as a ratio of mean atmospheric pressure in the

Azores to that of Iceland between December and the following March. Thus extreme high mean values result from intense low pressure centred over Iceland, corresponding with high pressure centred over the Azores; when this situation is reversed, the corresponding values of the NAOI are low (Hurrell et aL, 2003). A high value NAOI is likely to be indicative of higher frequency and more violent winter storms crossing the Atlantic Ocean on a more northerly track. This results in warmer and wetter winters in Europe and in colder and dryer winters in northern Canada and

Greenland, while the eastern United States is more likely to experience milder and wetter winter conditions (Hurrell et aL, 2003). As a consequence,sea surface winter temperatures over much of the Arctic will generally be colder at a higher NAOI

(Dickson & Turrell, 2000; Hurrell et aL, 2003). Low NAOI values are likely to indicate fewer and weaker winter storms bringing moist air into the Mediterranean and cold air to northern Europe. However, areas such as Greenland will have milder

73 winter temperaturesand as a result, Arctic seasurface temperatures will be generally higher (Huffell et aL, 2003).

The NAOI has previously been shown to correlate with changesin behaviourand abundanceof a wide range of natural animal populations, from the timing of migration in birds (Forchhammeret al., 2002),plankton abundance (Reid & Planque,

2000),abundance of squid (Loligoforbesi) in Scottishwaters (Pierce & Boyle, 2003) to emergenceof sea-troutfry from spawningbeds (Elliott et aL, 2000).

The spatialscale and geographicrange over which the NAOI appearsis indicativeof climate patterns and makes it a potentially powerful predictor of population fluctuations in European Atlantic salmon foraging at sea in winter.

Determiningclimatic influenceson the dynamicsof wild populationsrequires long- term measuresof populationabundance; such datasets are very rare. Howeverwithin the Foyle catchment,Ireland, there is a long history of commercialfishing for Atlantic salmon extendingback severalhundred years (Anon. 1953; Elson & Tuomi, 1975;

Anon., 1996). Commercialcatch records have been maintained since 1875with only a short break beforeand during the First World War (1909 to 1919). Theserecords show periods of abundanceand periods when the catcheswere substantiallylower.

Many reasonsfor these changeshave been proposedsuch as Ulcerative Dermal

Necrosis(UDN), drainageschemes, pollution and land use change(Elson & Tuomi,

1975; Shearer,1991; Magee et aL, 2003; Waring & Moore, 2004). However, here theserecords are usedas an index of population size, to examinelong-term change and the potential effects of climate. Specifically, the hypothesisis testedthat the

74 NAOI can be used to explain variance in the abundanceof salmon in the Foyle catchment.In addition, published data from climate change models predicting the effect of global warmingon the NAOI areused to makepredictions about the possible effectson salmonpopulations in the future.

4.2 Materials and Methods

Catch data are not absolutemeasures of populationsize (Shearer,1992), but there is evidencethat thesedata do representa significant relative measureof the dynamic change in returning migrant population size for the Foyle system. Firstly, catch returns apparentlyaccurately reflect actual catchesfrom the commercial fisheries

(Wysner,pers. comm. ). Secondly,over a 49-year period since 1952,there has been an independentmeasure of populationsize. The numberof redds(nests) have been monitored annually at 260 sites spreadthroughout the catchment.Over the period

1952 - 2000, commercialfishery catch was a very good predictor of redd number ?=0.45; (Linear regression:FI, 47= 39.9; p<0.001).These data strongly supportthe principal assumptionof the study presentedhere; namely that commercialnet catches reflect change in the relative population size of the returning migrant salmon populationto the River Foyle.

Data from commercialcatch returnsare collatedby the LoughsAgency of the Foyle,

Carlingford and Irish Lights Commission. Prior to 1999, its predecessorsthe Foyle

FisheriesCommission (1952 - 1999)and the Foyle andBann SystemsLtd. (pre 1952) gatheredand collatedthis information. Overall commercialcatch returnsdating from

1875 are availableon a yearly basis. The commercialcatch returns arise from four

75 types of fisheries.Drift nets are gill nets deployedclose to the water surfacein the lough seawardsof Lough Foyle. Draft netsare seinenets, which are deployed in-river in the main River Foyle and the at defined netting stations.Stake netting was a commercialtrap fishery operatedby the Foyle FisheriesCommission between

1952and 1988.Bag netsemployed a similar constructionto stakenets but wereleft in place permanently.One licencewas issuedannually to be fished in the Lough Foyle estuary,although in 1988 an additional licence was issued. Data for thesenets are availablefrom 1964to 1990when operationsceased. Post 1952when more detailed records are available, the data show catchesof adult salmon peaking during the monthsJune to July. Fish countingstations operating at presentindicate that the main run of returning adult salmon occurs during these months with lower numbersof salmon returning in the spring and autumn/winter periods (Loughs Agency, unpublisheddata).

The catchdata used here were not correctedfor effort, as prior to 1952no measureof effort wasrecorded.

Data on the NAOI are availablefrom 1856 onwards(Hurrell, 1995; Hurrell, et aL,

2003). The winter index usedin this analysiswas calculatedby taking the meanof the index betweenDecember and the following March. This winter NAOI then referencesto the Januaryof that year. NAOI varies from year to year, but also exhibits a tendencyto remain in one phasefor intervalslasting severalyears. In order to smoothout inter-annualvariation and examineperiods of abundance,a five-year runningaverage of both the NAOI andcommercial catches was used in the analysis.

76 The datawere analysedusing the statisticalsoftware package SPSS 10.0. Linear and

breakpointregression analyses were used to determinethe relationshipbetween the

long-termcatch data and the NAOI. RelativeRisk Analysis was usedto determinethe

probability of catchesexceeding a definedthreshold (Lachin, 2000)

4.3 Results

The commercialcatches of Atlantic salmonin the Foyle systemvaried significantly

between 1875 and 2001 from a maximum of 149,635 to a minimum of 12,500

salmon.Annual meanreported catch was 52,082+/- 2,541 (mean+/- S.E. ). Although

there is considerableinter-annual variation evident, some underlying year-on-year

trendsare apparent(Figure 4.1). A five year running meanshowed that catcheswere

high at the end of the 1800s;in the mid 1930sto 1940s;in the 1960sand early 1970s

and againin the 1980sand early 1990s(Figure 4.2).

Figure 4.1 Annual commercialcatch of salmonin the Foyle area1875-2000.

160000 140000- 120000- 0 E 100000- 1 80000- is 60000- c; 40000 - 20000- 0i 1875 1900 1925 1950 1ý75 2000 Year

77 Figure 4.2 The five year average commercial catches of migrant Atlantic salmon from the Foyle catchment 1875 to 2001 & the five year average NAOI: 1875 - 2000.

140000 4

no. of salmon 120000-- --3 ...... NAOI 100000-- 2, ý

0 :% S 80000-- It ev Oi 60000-- --t

40000--

20000-- ---2

0 -3 1875 1900 1925 1950 1975 2000 Year

To examinethe effect of climatic changeon catches,the five-year running meanwas

regressedon the five year running mean for the NAOI. This showed a highly

significant negative relationship (Fj,jo6 = 79.07, r2 = 0.43, P<0.001, b= -0.654,

intercept= 55,790,test of gradientt= -8.892, P<0.0001). However, there is some

evidencethat the relationshipbetween NAOI and catch was not constantacross the

full range of NAOI values. Breakpoint analysis was used to determine the point of

any potential change of influence of the NAOL This analysis suggesteda substantial

NAOI relationship change above a NAOI value of 0.151. Years when the was above

this point showed no significant relationship between the NAOI and catches (FI, 50=

1; 1.08). However, for 1.16; P=0.286; b= -0.15; intercept = 45,66 test of gradient: t= -

years with a NAOI below 0.15 1, there was a highly significant negative relationship

b= intercept 41,005.8; (FI, 54 = 129.97; r2 = 0.701; P<0.001; -0.84; = test of

78 gradient: t=- 11.4; P<0.00 I) (Fig. 4.3). It was noted that the years of the lowest

recorded NAOI (1960's) coincided with the highest catches (Fig. 4.1).

Figure 4.3 The relationship between the winter NAO and commercial fishery catches of Atlantic salmon returning to the River Foyle over 122 year. Breakpoint analysis shows a uncoupling of a negative relationship about a NAOI of 0.15 1.

140000

"" 120000 "" w 100000 .' . 80000 " "" Z 0 60000 fj 40000 $ (0 20000

02 34 NAO

Seven models of the effects of climate change on the NAOI resulting from a 1% per

annum compound change in atmospheric carbon dioxide levels have been summarised

by Gillett et al. (2003) (Figure 4.4). All models predicted an increasing NAOI with

time. The mean of the seven models suggestsNAOI will increase approximately five

fold over the next 75 years. Assuming that the NAOI has a similar responseto change

over time as does the NAOI for the rest of the year, NAOI values of under 0.151 will

become less frequent than in the past, the frequency of occurrence decreasing with

time. There is some evidence that this may already be occurring as in only 2 of the

last 10 years has the NAOI dropped below 0.151. As a result, data presented here

would suggest that the linkage between the NAOI and population size would

frequently remain uncoupled in the future. In addition, these data predict a change in 79 the frequencyof abundant,returning migrant populationsof salmon.Relative Risk

Analysis showsthat the probability of achievinga catch of salmongreater than the median(47,800 between 1879 and 2001) is 2.34 times lower for a NAOI greaterthan

0.151 than for a NAOI of less than 0.151 (Relative Risk Analysis: RR=2.34; 95% confidence limits = 1.52-2.34; Yates corrected P<0.0005). Thus, assuming that climate changemodels do reflect likely patternsof changein the NAOI, then the long-termfuture medianpopulation size of returningmigrant salmon is likely to be lower thanthat of the past.

Figure 4.4 The changein North Atlantic Oscillation predicted by seven climate changemodels. Dotted line showsthe meanof the sevenmodels, solid lines show the 2 extremesof the predictedNAO change.

12 10

"" "_"_4 ." _. -...

. 4- . "" CL 0 4.4 4c z

2000 2010 2020 2030 2040 2050 2060 2070 2080 2090 2100 Year

I-Predicted Max -Predicted Min ...... Average

80 4.4 Discussion

The population size of adult Atlantic salmon in the Foyle system as indicated by commercialcatch returns has been highly variable over the years since 1875, but catchesare generallyregarded as large when comparedwith other systems(Shearer,

1992). Within this variation there are generaltrends of changingabundance linked with variation in the North Atlantic Oscillation. This feature may affect salmonid abundancein severalways. In the freshwaterphase, Elliott et aL (2000) found that the dateof fry emergenceover 30 yearsin a seatrout nurserystream, predicted by an

individual basedmodel (Elliott & Hurley, 1998),correlated highly significantly with

the NAOL Theseauthors also found a significant relationshipwith instreamwater

temperatureand suggestedthat this may be the actualmechanism driving variation in

the model. As a measureof the dominant mode of winter climate in the northern

hemisphere,the NAOI has also beenlinked directly to seatemperatures which affect

salmonwhen feeding at sea(Martin & Mitchell, 1985;Reddin et aL, 2000;Dickson &

Turrell, 2000). Here, a negativerelationship is shown betweenthe NAOI and the

abundanceof migrantsalmon returning to the River Foyle in Ireland for indicesbelow

0.151. Above 0.151 this relationshipbreaks down and at higher values there is no

significant relationshipbetween NAOI and catch abundance.The proximateclimatic

mechanismmodulating salmon abundanceis unclear, but there are a number of

possibilities.Dickson and Turrell (2000) outlined the relationshipbetween ice flux

and a positiveNAOL which increasedinto the 1990s.This potentiallyaffects salinity

and watertemperature. The NAOI may also indicatechanging thermal habitat optima

available to salmon through the cooling or warming of either migration routes or

feeding areassuch as thosenorth of Iceland or off Greenland. Martin and Mitchell

81 (1985) found evidenceto suggestthat the abundanceof multi-sea-winter salmon returning to the River Dee in Scotland, was linked to an increasein sea-surface temperaturein the waters north of Iceland, which has in turn been linked to a decreasingNAOI (Hurrell et aL, 2003). This mechanismsuggests that a strongly negativeNAOI resultsin an increasedquantity or quality of thermalhabitat available to Atlantic salmon,resulting from increasingwater temperaturein theseareas. In the northwest Atlantic, a thermal habitat index has been used effectively to predict salmon abundance.This index shows a significant relationshipwith the NAOI and may thereforebe considereda useful tool in determiningsalmon abundance (Dickson

& Turrell 2000; Beaugrand& Reid, 2003). The study presentedhere indicatesthat this may not be valid now or in the future.

Although when hindcastingover the last 126 yearsthe NAOI is clearly a significant predictorof returningsalmon migrant numbers, it is clear that the value of the NAOI as a forecastingtool to predictpopulation size is highly suspect.Here it is shownthat when the NAOI exceeds 0.151, then predictivecapacity breaks down. Over the last decadethe NAOI has been greaterthan 0.151 in 8 out of the 10 years. Basedon current climate changemodels, this is a trend that is likely to becomemore extreme.

Thus it is concludedthat it is unlikely the NAOI will be a useful tool to forecast

Atlantic salmon population abundanceand that its inclusion in predictive models shouldbe viewed with caution.

82 Chapter 5. Local instrearn and catchment spatial scale habitat characteristics determine 0+ fry density of Atlantic salmon (Salmo salar L.) in the River Foyle.

5.1 Introduction

Temporal variations in population size arise as a result of density-dependentor density-independentfactors and the interplay betweenthem (Elliott, 1994; Newton

1999; Harwood et aL, 2003). In contrast,spatial variation in density is primarily a function of the habitat in which animals live (Jenkins, 1964; Kennedy & Strange,

1980;Kennedy & Strange,1982; ElliotL 1994). For a numberof speciesenough is known about the relationshipbetween habitat use and population size to be able to construct predictive models to estimatecarrying capacity of a given habitat. For example Alisauskas and Arnold (1994) in a study on American coots (Fulica americana)found that a linear relationshipexplained a high proportion of variance

(r2---0.54;p<0.001) between an estimateof populationsize (numbersshot eachyear) and the numbersof ponds on the prairies the preceedingsummer. Models such as these allow predictionsof spatial patternsof habitat use and give an insight into mechanismsoperating on the populationthus enablingpredictions to be madewhich supportpopulation and habitatmanagement. This is particularly importantin species of high conservationvalue or which are exploited. One such speciesis the Atlantic salmon (Salmo salar L. ). A number of habitat type, population regressionmodels linking local habitat type and density already exist for stream dwelling salmonids

(B inns & Eiserman,1979; Milner et al, 1995,1998;Grant et al., 1998;Poff & Huryn,

1998).

83 As some elementsof habitat for stream dwelling salmonids, for example water quality, are partly a function of activity higher in the catchmemthis study attemptsto extend existing models by considering environmental effects at both local and catchmentscales. Specifically, it attemptsto model effectsof the uppercatchment on the deviationsin densityestimated by local habitattype.

5.2 Materials and Methods

Five minutetimed electrofishingsurveys (Crozier & Kennedy,1994) were carried out using an Electracatchbackpack electrofisher at 307 and 279 sites, during July and

Augustof 2001 and2002 respectively,throughout the River Foyle catchment,Ireland.

This work is part of the LoughsAgency's ongoingdevelopment programme and was carriedout by field staff. The Agencyhas a maximumof 350 suitablesites identified, a target of 200 of theseare to be fished at a minimum each year. For each site all underyearling(0+) salmonfry and brown trout (Salmotrutta L. ) were counted.Fish lessthan 7 cm werecounted as 0+. Other fish speciessuch as stoneloach(Barbulata barbulatus),roach (Rutilus rutilus), eels (Anguilla anguilla), encounteredwere noted as present.

Sites were chosenspecifically for their generalsuitability for juvenile salmonids,i. e. relatively shallow (10-30 cm deep)and relatively coarsesubstrate. Sixteen measures of in-streamand banksidehydrology and geomorphologywere gatheredconcurrently with the electrofishingdata and used in the analysis(Table 5.1). The widest site fished was 33 in with a maximum depth of 30 cm. The greatestmaximum depth fishedwas 120cm with a site width of 24 in.

84 Table 5.1 Local scalesite-specific variables collected for eachsampling for 350 sites in the Foyle catchment.

Width - this was the average width at the site and was measured in metres. Minimum depth - minimum depth of water measured in centimetres. Average depth - average depth of water at the site measured in centimetres. Maximum depth - maximum depth of water measured in centimetres at the site. Flow (m-) -A net float was timed over a 10 m stretch and recorded in seconds. Bankside cover - right hand bank - this was recorded as a percentagefor both banks. Bankside cover - left hand bank Over hanging cover - right hand bank - this was recorded for presenceor absence. Over hanging cover - left hand bank Percentagebedrock Percentageboulder Percentagecobble Percentagegravel Percentagefines Percentagesand Percentagemud

For the purposesof analysis,the data sets were combined, and every secondsite extracted. One set of datawas usedto constructthe model while the secondset was used to test its predictive power.

To categorisethe broader scale catchment on salmon fry density, data for the catchmentupstream of each survey site were extractedfrom 1:50,000 Discoverer

SeriesOS mapsand GeologicalSurvey of N Ireland maps(1: 50,000 Drift Edition and

1:250,000 Quarternary Edition). Catchmentdata collected included (Table 5.2):

85 Table 5.2 Catchment scale characteristicscalculated from ordnance survey and geologicalsurvey maps for eachof the 350 samplingsites in the Foyle catchment. Catchment area (krrý) Stream order Distance of the site from the stream source (km) Distance of the site from the streamsconfluence with the main River Foyle (km) Altitude (m) Slope at the site Presenceof lake upstream No. of housesper km' upstream of site Distance to nearesthouse (km) Distance to nearestfarm (km) Percentageof urban areasupstream Percentagearea of grassland upstream Percentage rea of woodland upstream Percentagearea of peat upstream Percentagearea of glacial alluvium upstreaiý Percentagearea of glacial lacustrine upstream Percentagearea of glacial calcareoustufa Percentagearea of glacial boulder clay Percentagearea of glacial sand and gravel Percentagearea of bed rock Percentagearea of gneiss Percentagearea of felsite Percentagearea of diorite Percentagearea of sandstone

5.2.1 Catchment Scale Characteristics

Catchment area

The catchment area above each sampling site was calculated by determining the river networkboundary above the samplingsite andcalculating catchmentarea in kM2.

Streamorder

Streamorder was derivedusing the methoddeveloped by Horton (1945).

86 Distance from source

The distanceof the electrofishingsite from the upper-mostreaches of that arm of the river system.

Distance from confluence Wth the River Foyle

This is the distanceas measuredfrom the map of the survey site to the streams confluencewith the main-stemof the River Foyle.

Altitude

This was taken by interpolationfrom height contourson the map either side of the site.

Stream Gradient

The stream'sgradient was derivedby measuringthe horizontal distancebetween the nearesttwo contourlines and dividing this by the changein altitude.

Presence or absence of a lake upstream

Number of houses present

This wascalculated by countingthe individual housesfrom the map.

87 Housing density

A densityfor the catchmentwas then derivedby dividing the numberof housesby the catchmentarea (km2)above the samplingsite.

Distance to nearest house

The distanceto the nearesthouse from eachsite was measuredin metres.

Land use

Three categoriesof landuse:urban; grasslandand woodland were defined from the

1:50,000 series 'Discoverer Series' maps. The areaof eachwas measuredin km2for eachcatchment, above each sampling site. Thesewere then convertedto percentages of the sub-catchmentarea and arc-sine transformed before fuller analysis.

Geology

Seven geologicalcharacteristics of the catchmentoutlined below were also categorisedfor eachsite. These were obtained from 'Drift andQuaternary editions of

GeologicalSurvey of N Ireland' maps. The areas,in kmý,covered by the various

depositswere measured on a digitisingpad and converted to percentageswhich were

thenarc-sine transformed for analysis.

88 The characteristicsused were:

Drift

9 Peat

* Glacial alluvium - Waterbornematter depositedby rivers

9 Glacial lacrustine alluvium - Similar to alluvium only deposited in a

lacustrineenvironment

" Glacial calcareoustufa

" Glacial boulderclay

" Glacial sandand gravel

Solid

" Bedrock

" Gneiss

" Felsite

" Diorite

" Sandstone

5.2.2 Analysis

The data was modelled in SPSSV. 10 using single and stepwiseforward multiple-

regressionshaving 0+ salmonas the dependentvariable.

89 5.3 Results

Regressingthe catchof 0+ salmonfry from the 5 minute electrofishingsamples using all sites including thosewhich had no 0+ salmonpresent, on local site specific data

(Table 5.1) in a forward stepwisemultiple regressionshowed that maximum water depth at the site was the best predictor of 0+ fry number,deeper water having lower

0+ abundance(Fi, 139=4.91; r2---0.027; p--0.028). The model was significantly improved however, by the addition of the width at the site (F2,138=4.65;r2=0.05; p---0.011)and the percentageof bedrock present(F3,137=4.317; r2---0.066; p<0.006)

(Table5.3).

Table 5.3 Stepwise regression of local site specific stream characteristics on G+ salmon numbers for a) all sites (including sites with no 0+ salmon); b) only sites whi-rP ftA.cnlmnn were nrerzent. a) F df rz p< Maximum water depth at site 4.91 1,139 0.027 0.03 Maximum water depth at site Width 4.65 2,138 0.050 0.01 Maximum water depth at site (-); Width (+); 4.32 3,137 0.066 0.006 Bedrock b) Bedrock 3.844 198 0.028 0.053 Bedrock(-); Average water depth 3.821 12,97 0.054 0.025

Increasingriver width had a positive effect on numbersof 0+ salmonpresent but the

percentageof bedrockat the site was negativelyrelated to salmonnumbers. When

only sites with 0+ salmonpresent were included it was found that the percentageof

bedrockand averagedepth providedthe best model (F2,97=3.82;r2---0.054; p=0.025),

both of thesewere negatively related to increasingsalmon numbers.

Local instrearnhabitat characteristicswere highly significant predictorsof between

site variation in salmonnumbers, however the percentageof variation explainedwas

low (6.6% for all sites;5.4% for sites without 0+ salmon). To test if broaderscale

90 catchmentcharacteristics were affecting salmon abundance,wider catchmentscale variables(Table 5.2) were used as predictors against 0+ salmon numbers.Firstly, thesewere examinedagainst all of the available sites including those which did not have 0+ salmon. The percentageof glacial sandand gravel in the catchmentabove the site was found to be the best predictor (FI,73ý--11.83; rý--0.128; p<0.001), the addition of felsite (F2,72--9.63;r2=0.189; p<0.001) and altitude at the site greatly

improvedits predictive ability (F3,71=8.421;r2=0.231; p<0.001) (Table 5.4a). When

only sites where salmon were found were included the size of the catchmentarea

abovethe site wasthe bestpredictor (FI, 48=10.426; r2---0.164; p<0.002), the addition of

the percentageof sandstone(F2,47=7.846; r2--0.218; p<0.001) and the distanceof the

site from the source of the river improved the predictive capacity of the model

(F3,46=7.045;r2=0.270; p<0.001) (Table 5Ab).

Table 5A Stepwiseregression of catchmentcharacteristics on 0+ salmonnumbers for a) all sites (including sites with no 0+ salmon);b) only sites where 0+ salmonwere present.

a) F df r' P< Glacial sand & gravel 11.83 1,73 0.128 0.001 Glacial sand & gravel Felsite 9.63 2,72 0.189 0.001 Glacial sand & gravel Felsite Altitude 8.42 3,71 0.231 0.001 b) I I Catchment area 10.43 1,48 0.164 0.002 Catchment area Sandstone 7.85 2,47 0.218 0.001 Catchment area Sandstone (-); Distance from 7.05 3,46 0.27 0.001 mainstern Foyle I I I

Both of thesespatial scaleswere then combinedin a single analysisto determineif

catchmentfeatures were more importantthan site specific instrearncharacteristics. It

was found that when all the sites(including thosewith salmonabsent) were included

glacial sandand gravel,percentage of felsite, altitude, minimum depthof water at the

site and stream width provided the best predictive model (F5,69=7.347;r2=0.300;

91 p<0.001) (Table 5.5a). Of these, altitude and the minimum depth of water were negatively related while the rest were positive. When the catchment scale characteristicsand instrearnvariables were included for only those sites with 0+ salmonpresent, the bestpredictors were found to be catchmentarea above the site, the maximumdepth of water found at the site, the numberof houseskrn-2 upstream, the amount of urban area upstream,the percentageof boulder clay upstreamand the percentageof sandstonein the catchment above the site (F6,43=7.190;r2=0.43 1; p<0.001)(Table 5.5b). Maximum depthof water, the amountof urban areaupstream and the percentageof sandstonewere negativelyrelated to 0+ numberswhile the rest of the variableswere positive.

Table 5.5 Stepwiseregression of local scaleand catchmentcharacteristics combined on 0+ salmonnumbers for a) all sites (including sites with no 0+ salmon);b) only siteswhere 0+ salmonwere present.

Equation 1: Y= (Glacial sand/gravel*267.777)+ (Felsite*810.421)+ (Altitude*- 0-0676)+ (Minimum waterdeDth at site*-0.300)+ (Width at site*0.543)-0.0439 a) F df r' P< Glacial sand & gravel 11.83 1,73 0.128 0.001 Glacial sand & gravel Felsite 9.63 2,72 0.189 0.001 Glacial sand & gravel Felsite Altitude 8.42 3,71 0.231 0.001 Glacial sand & gravel Felsite Altitude 7.60 4,70 0.263 I 0.001 Minimum water depth (-) I Glacial sand & gravel Felsite, Altitude 7.35 I 5,9 0.300 10.001 Minimum water deptli (-); Stream width (+)

92 Equation2: Y= (Catchmentarea*0.124) + (Maximum water depth at site*-0.331)+ No. houseskrd I *2.182) + (% urban area upstream*-1197.368)+ (Glacial boulder clav*252.548)+ (Sandstone*-2099.587)-5.801 b) F df r' p< Catchment area 10.43 1,48 0.161 0.002 Catchment area Maximum water depth 8.54 2,47 0.235 0.001 Catchment area Maximum water depth No. 7.31 3,46 0.279 0.001 2(+) houseskm 1 1 Catchment area Maximum water depth No. 6.92 4,45 0.326 0.001 houseskrný (+); Urban area upstream (-) Catchment area (+); Maximum water depth No. 7.04 5,44 0.381 0.001 houses krlf2 Urban area upstream (-); Glacial boulder clay Catchment area (+); Maximum water depth (-); No. 7.19 6,43 0.431 0.00 houses kM, 2 (+); Urban area upstream (-); Glacial boulder clay (+); Sandstone(-)

In order to test the predictive ability of the combined local and catchmentwide characteristicsthe seconddata set of electrofishingsites was used. Salmonnumbers predictedfrom the sites not used in model development(equation 1: using all sites) significantlypredicted actual numbers caught, although with a low degreeof variance explained (FI,80=4.849; r2=0.045; p---0,03 1). Using equation 2 i.e. only those sites where 0+ salmon were present, also significantly predicted the actual numbers of fish caught,although again a very small amountof variancewas explained(Fl, go= 7.51; ? ---0.045;p=0.007). In order to investigate this further, paired Wests were used, it was found that using salmonnumbers predicted by equationI did not differ significantly from actual number (t=-0.609; 8ldf; p---0.544). The mean error, the differential betweenactual and predictedfish being 84% ±13S.E. Similarly, using the model developedfor only siteswhere salmon were detected(equation 2) actualfish numbers did not differ significantly from predictednumbers (t=0.354,91df; p=0.724). The meaneffor rate being 56%±15.23S. E.

93 5.4 Discussion

Local habitat structure is known to influence the density of juvenile salmonids

(Kennedy & Strange, 1982,1986a, 1986b; Elliott, 1994; Kocik & Ferreri, 1998).

However,wider catchment,characteristics may be importantfactors in explainingboth temporaland spatialsalmonid variability (Milner, et aL 1998;Pess et aL, 2002). The model developedin this study proposessome novel additions to previous models, which are shownto be significantin explaining0+ salmonnumbers. Two variantsof the model were developed,variant I included local instreamand catchmentscale characteristicsand all of the available sites (including those where 0+ salmonwere absent) while the second variant (including local and broadscalecharacteristics) included only sites at which 0+ salmon were present. It was found that this combination of broader scale catchmentcharacteristics (glacial sand and gravel, felsite and altitude) and instreamvariables (minimum water depth and river width) explained30% of the observedvariation in variant 1. In this model catchmentwide features clearly explained more of the variation in salmon numbers than local instreamfeatures. In the secondmodel variant, differing broadscalecharacteristics were also found to be important(catchment area above the site, the numberof houses km-2upstream, the amount of urban area upstream,the percentageof boulder clay upstreamand the percentageof sandstonein the catchment)and also different site specificdata (maximum depth of waterfound at the site). This variant of the modelas might be expectedalso explained a higher proportion of the explainablevariance

(r2---45%)

The relationshipsin both of these models exposea strong relationship between broadscalecatchment wide featuresand 0+ abundanceand are consistentwith current

94 understandingof how theseunderlying physical attributescan influence fish habitat potential. Geology and geomorphicprocesses dictate the range of morphological characteristicsa streamreach can exhibit, thus partially determiningthe physicaland biological characteristicsof fish habitat.In model I which includedall sites,the area upstreamof glacial sand and gravel had a positive effect on salmon abundance.

Juvenilesalmon generally inhabit areasof shallowfast flowing water with a substrate, which is gravel and cobble dominated4in addition, adult salmon require gravel depositsfor spawning(Jones, 1956; Shearer, 1992). Thereforeit is likely that greater

areasof sandand gravel depositswill contribute to providing such suitablehabitat.

The secondvariable, felsite, is an igneousrock. Hicks and Hall (2003) found that in

the presenceof anotherigneous rock, basalt, production of juvenile salmonidswas

positively relatedto its abundance,and they found that streamswhich flowed through

areas of this rock type had generally steepergradients than those which flowed

throughsedimentary sandstone rocks and that riffle habitatpredominated which is the

favouredhabitat forjuvenile salmonids.Salmon abundance decreased with increasing

altitude,higher reaches of streamsare generally regarded as being lessproductive as a

result of lower temperatureswhich are known to adverselyaffect growth rates and

densitiesof salmonids(Alabaster, and Lloyd, 1987;Elliott et aL, 1998). The first of

the two local habitat features included was minimum water depth which had a

negative effect on salmonid abundance. Symons and Heland (1979) found in

laboratoryexperiments that 0+ salmonpreferred depths in the range6-9cm deep. It is

postulatedthat, as minimum depth at the surveyedsite decreased,it was dropping

below this preferredrange for 0+ fish, and thereforedensities decreased as a result.

The secondand final local habitat feature was river width which was positively

related to 0+ abundance.This variable is possibly related to fry dispersal post

95 hatchingasadults tend to spawnin the upper reacheswith fry dispersingdownstream

(Shearer,1992).

In model variant 2, broaderscale catchment characteristics were more importantthan site specific variables in explaining variation. Larger catchmentshad a positive influenceon salmonabundance, as wasfound by Milner et al. (1995). Interestinglyas the numberof housesincreased salmon numbersalso increasedalthough increasing urban areashad a negativeaffect. This suggeststhat up to a certain level increasing housenumbers may improve productivity, possibly through increaseof nutrientsbut above this threshold they may become detrimental to stocks, with increasing proportionsof urban areasin catchmentshaving a negativeeffect on fish abundance.

Sandstone,being a sedimentaryrock is the corollary of felsite, and as notedby Hicks and Hall (2003) showed negative relationshipsbetween salmonid abundanceand increasingpresence of of this rock on rivers in north America. They found that rivers which flowed through sandstonewere generally low gradient with long reachesof deepslow flowing water,areas such as theseare not favourableto 0+ Atlantic salmon production. The only site specific variablewhich was included in the secondmodel variant was maximum water depth. Kennedy and Strange (1985) also found a negativerelationship between increasing water depthand salmonidabundance in their study on streamsin N Ireland, with preferred0+ salmonhabitat in water generally lessthan 20cm in depth.

The results show that there is no significant statisticaldifference between the actual abundanceof salmonand those predictedby the models,but the averagedeviation betweenactual and predictedcatches was 84 and 56% respectively. Thus the second

96 variant of the model was the more robust of the two. This is to be expectedas only siteswith 0+ salmonwere included. The relatively low predictivecapacity and high error rate may in part be attributableto the choice of site. The sites chosenwere identified specifically as good quality sites for 0+ salmon and so are likely to be similar within a narrow range of attributes, thus eliminating large variation. The predictivecapacity would in all likelihood be improvedthrough the selectionof sites which arenot as suitedto 0+ salmonproduction.

In summary, it is shown in this study that habitat on two scales, site specific and catchment wide, affects the juvenile density of salmon in the River Foyle catchment.

Instream habitat characteristics are shown to be significant predictors of juvenile abundance, but with the inclusion of broader catchment scale characteristics the predictive power of the model is greatly increased. Increasing urbanisation is highlightedas a potentialfuture threatto the salmonof the Foyle area.

97 Chapter 6. General Discussion

The Atlantic salmon is in decline in many parts of its range. However, the Foyle population, having declined from a period of high catchesin the 1960's and 1970's

(Figure4.1), remainsrelatively stableat a lower level of productionwhich is still high in comparisonwith manyother systems(Anon., 2001).

In periods when abundanceis high, conflicts between users (sport anglers and commercialharvesters) of the resourcemay diminish. However, during periodsof low

abundancethis is liable to change. Theseconflicts generallyarise out of concernover

resourceallocation and issuesof conservation. Chapter2 illustrated,using a long-time

seriesof data the impactsof each of the user groups on each other and on successful

spawningescapement. A 49 year time seriesdataset comprising commercial net and trap

catches,recreational angling returnsand an independentmeasure of breedingpopulation

size for the Atlantic salmonin the Foyle catchmentwas usedto investigatethe impactof

competing methods of exploitation on each other and the effect of exploitation on

populationsize. Positive correlationsbetween commercial netting and trapping catches

over the 49 years of this study did not support the hypothesis that commercial

interceptionfisheries impactednegatively on each other. In order to test the hypothesis

that large returning migrant population sizes mask underlying negative relationships

betweenexploitation, or betweenexploitation and spawningpopulation size, years with

smaller than average returns were examined separately. No significant negative

correlations between catchesfrom differing forms of exploitation nor between total

exploitationand spawningpopulation size were detectedin this subsetof the data. With

98 regardto sport angling catchesit was shownthat a reductionin commercialexploitation may result in an increasein the sport fishery catch. However with only 19% variance explainedby commercialcatches, this effect was weak. Positivecorrelative relationships betweenall forms of exploitation combinedand an independentmeasure of spawning populationsize did not supportthe hypothesisthat exploitationhad a significantnegative impactupon ultimatespawning population size between 1952 and 2000.

It was concludedthat the Atlantic salmon population supportedby this catchmentis sufficiently large to mask any potential for negative impacts of exploitation on sustainabilityof stocks.In addition, althoughthere is somesmall statisticallydetectable impact from commercial harvesting on the sport fishery, the different modes of commercial exploitation do not significantly impact detrimentally upon each other.

Therefore, in the Foyle area, cessationof commercial harvesting should be largely regardedas an issueof resourceallocation rather than conservation.

In Chapter3 using the samelong-term data set, the role of density dependentand life- stagespecific environmentalfactors in controlling population size was investigated.A

Ricker density-dependentmodel showedthat spawningadult populationsize significantly explained variation in the resultantfilial generation,however a significant amount of variation (ca. 68%) remainedunexplained. It was shownthat environmentalfactors were significant in explaining someof the remainingvariance and that theseinfluences were linked to specific life stages. This finding strongly suggestedpopulation bottlenecks in the complex life cycle of this species,during which, specific environmentaleffects may have had an impact they did not have during other periods.It was concludedthat these life stage specific environmentaleffects were likely to contribute to the stochastic

99 variation in population size remaining after the application of traditional stock- recruitmentmodels and that the identification and quantificationof theseeffects should allow improvedmodel accuracy.

Chapter3 highlightedthe importanceof variousenvironmental parameters on salmonand in Chapter4 this was investigatedfurther by the testing of the effect of marineclimatic conditionsin the North Atlantic on the abundanceof returning migrant Atlantic salmon, using a 126 year datasetof commercialcatches. ' Catches of salmon from commercial netting stations significantly predicted a measureof population size independentof catchesover a 49 year period; hence commercialnet catcheswere assumedto be an adequatemeasure of relative population size. The North Atlantic Oscillation index in winter (NAOI) provides a generalisedmeasure of climate variability for the northern hemisphere. Between 1875 and 2001, the NAOI explaineda significant proportion of variation in five year running meancatches of migrant Atlantic salmonreturning to the

River Foyle. When the index was below 0.151, the NAOI correlatednegatively with salmon catches 70%) indicating that a significant proportion of the variance in population size in the past was the result of variability in conditions in the marine environment.However, when the NAOI was above 0.151, this relationshipuncoupled.

The probability of catchesexceeding the long-term median, was 2.34 times lower in yearswhere the NAOI was abovethe 0.151breakpoint than for yearswhen it was below

0.151. The NAOI had exceededthe 0.151 thresholdon 8 out of the 10 years prior to

2001. Models of climatechange indicate that the NAOI is likely to increasesignificantly with time. If these models are correct, this study would lead to the conclusionthat a decouplingof broad scale climate effects on salmon population size will becomethe norm. Data presentedhere suggest two consequencesof this: that the value of the NAOI

100 as a predictivetool for forecastingadult salmonpopulation size will be limited; and that the median population size will become lower in the future.

Chapter5 testedthe capacityof local instreamand broadscalecatchment characteristics to predict 0+ salmonabundance within the Foyle area.Data was collectedfrom 307 and

279 sites for annual electrofishing surveys from 2001 and 2002 respectivelyof 0+ salmonidsand semi-quantitativestream morphometric information. Using a combination of these local site-specificvariables and broadscalecatchment characteristics derived from Ordnanceand Geological survey maps, two models were constructed. Model I useddata for all availablesites i.e. including thosewhere 0+ salmonwere absent,while

model 2 wasdeveloped using only thosesites where salmon where present. Both of these

models were significant predictors of juvenile abundanceusing site-specific variables

althoughthe varianceexplained was small. However, both were improved significantly ? by the inclusion of wider broadscalecatchment characteristics (model 1 =30%; model 2

x2=43%).When testedagainst an independentdata set using pairedt-tests, neither model

differed significantly from the actualcatch of salmon,although both had relatively high

standarderror rates (84%±13 and 56%±15.2respectively). It was concludedthat site-

specifichabitat characteristics were significantpredictors of juvenile abundance,but with

the inclusionof broadscalecatchment characteristics the models' predictivepowers were

greatly increased.

Although the Foyle population of salmonhas declined from the 1960's and 1970's in

commonwith many other systemsaround the world, if the longer term catchesare taken

into considerationthis period of lower level abundancemay just be part of a longer term

101 natural cycle. If the period from the mid 1960's is taken as an example (Figure 6.1) it

may be seenthat the population appearsto be dramatically declining.

Figure 6.1 Foyle area annual commercial salmon catches 1962-2003 with regression line.

160,000 -

140,000- y= -2578x + 115209 120,000- R2=0.6468 r- 100,000- 0 E 80,000 -

C 60,000-

40,000-

20,WO - 0 -, IýIIIIIIýIýIIIIIIIIIIIýIIIIII-, IýIIII do ýý KI, Aý # -x'o ,pI FJK -lq Ile -lq Nq -IR,-IC', NOI -lq NOI-lq NOI-l0j ":b -1010 CNNoj ,q-, oj Year

Most of the intensive research work on Atlantic salmon commenced during this period.

However, if the longer term picture is looked at (Figure 4.1) it may be seen that the

catches during the 1960's coincide with the highest recorded abundance in the long-term

datasetand the period we are in now is almost like a background level of abundance.This

poses the question if there were similar long-term datasets for other stocks of salmon

would they show a similar picture?

Fluctuations such as these have been observed in many different animal populations as

previously seen with lemmings, voles (Ekerholm et al., 2001) and red grouse (Jenkins et

aL, 1963; Newton, 1998). Many of these cycles can be related to natural phenomena,

102 which operateindependently of direct anthropogenicinfluence, as for examplechanges in the distribution and abundanceof the Pacific sardineand northernanchovy (Rodrfguez-

SAnchezet aL, 2002). The resultsof this study have thereforewider implicationsthan just for the managementof salmonand it is hopedthat it will contributeto the debateon the managementof commerically exploited speciesand the relative roles of density driven populationmechanisms and environmental factors.

Specifically with regard to the Foyle salmon this study has highlighted a number of

issueswhich may impact on the return of the populationto the higher levels previously

experienced.Of particularconcern are thosewhich may not have beenan issueduring

previous periods of lower abundance,namely direct and indirect anthropogenic

influences. Furtherwork is requiredto understandhow theseinfluences interact to affect

the Atlantic salmon in particular further developmenton the combination of stock

recruitmentand environmentalmodels where availabledata permit. It is hoped that this

thesis will form the basis of future study, in particular the developmentof pre-fishery

abundancemodels which can be linked by genetic analysisto specific stock structures

within the Foyle catchment.

103 References

Achord, S. Levin, P.S. & Zabel, R.W. (2003). Density-dependentmortality in Pacific

salmon:the ghostof impactspast? Ecology Letters 6,335-342.

Alabasterand Lloyd (1987).Water Quality Criteria for FreshwaterFish. London: Food and

Agriculture Organisation.

Alisauskas,R. T. and Arnold, T.W. (1994). American Coot. Migratory shoreand upland

game bird managementin North America (T.C. Tacha & C.E. Bmun eds.).

Kansas:Allen Press.pp. 127-143.

Amiro, P.G., (1998a).Recruitment of the North Americanstock of Atlantic salmon(Salmo

salar) relative to annual indices of smolt production and winter habitat in the

northwest Atlantic. Canadian Stock Assessment Secretariat Research Document

98/45.

Amiro, P.G., (1998b). The abundanceof Harp sealsin the North Atlantic and recruitment

of the North American stock of Atlantic salmon (Salmo salar). Canadianstock

AssessmentSecretariat Research Document 98/84.

Anon. (1953). Foyle Fisheries Annual Report for 1953, Londonderry: Foyle Fishery

Commission.

104 Anon. (1969). Foyle Fisheries Annual Report for 1953, Londonderry: Foyle Fishery

Commission.

Anon. (1970). Foyle Fisheries Annual Report for 1953, Londonderry: Foyle Fishery

Commission.

Anon. (1971). Foyle Fisheries Annual Report for 1953, Londonderry: Foyle Fishery

Commission.

Anon. (1992a). Directive on the conservationof naturaland semi-naturalhabitats and wild

faunaand flora. HabitatsDirective 92/43EEC.

Anon. (1992b). Salmon ResearchAgency of Ireland Annual Report 1992, Newport:

SalmonResearch Institute of Ireland,Farran Laboratory.

Anon. (1996a). Foyle Fisheries Annual Report for 1996, Londonderry: Foyle Fishery

Commission.

Anon. (1996b). Report to the Minister of the Salmon Task Force, Dublin: Government

PublicationsOffice.

Anon. (1997).Habitat SurveyClassification. Departmentof Agriculture for N. Ireland.

Anon. (1998a). SalmonTask Force, Implementation Document. Dublin: Departmentof the

Marine andNatural Resources.

105 Anon. (1998b). Yhe Spey District Fishery Board Management Report 1998 and Policy

Review.Scotland. Spey District FisheryBoard.

Anon. (1999). Glas-Y-Dorlan 25, Environment Agency Wales, Spring.

Anon. (2001). Report of the working group on north Atlantic salmon, ICES CM

2001/ACFM:15 Part one,Aberdeen, Scotland, 2-11 April 2001.

Anon. (2003a). 77zeFoyle Fisheries Commission Annual Report and Financial Statements,

1997-1999.Belfast: TSO.

Anon. (2003b). Economic evaluation of sport angling versus commercial netting. Dublin:

CentralFisheries Board.

Anon. (2003c). Report of the 2dh Annual Meeting of the Council of NASCO. Edinburgh,

Scotland.

Azurnaya, T. & Ishida, Y. (2004). An evaluation of the potential influence of SST and

currents on the oceanic migration of juvenile and immature chum salmon

(Oncorhynchusketa) by a simulationmodel. FisheriesOceanography 13,10-23.

Beaugrand, G. & Reid, P.C. (2003). Long-term changesin phytoplankton,zooplankton

and salmonrelated to climate.Global ChangeBiology 9,1-17.

106 Beverton,R. J. H., and Holt, S.J. (1957). On the dynamicsof exploited fish populations.

FisheryInvestigations, London Series 2 19,1-533.

Bosch, M., Oro, D., Cantos,F. J. and Zabala,M. (2000). Short-termeffects of culling on

the ecology and population dynamics of the yellow-legged gull. Joumal ofApplied

Ecology37,369-385.

Buck, R.J. and Hay, D.W., (1984). The relation between stock size and progeny of

Altantic salmon, Salmo salar L., in a Scottish stream. Joumal of Fish Biology 24,

1-11.

Byrne, C.J., Poole, R., Rogan, G., Dillane, M. & Whelan, K.F. (2003). Temporal and

environmentalinfluences on the variation in Atlantic salmonsmolt migration in the

B urrishoolesystem 1970-2000. Joumal of Fish Biology 63,1552-1564.

Cass,A. J., andWood, C.C. (1994). Evaluationof the depensatoryfishing hypothesisas an

explanationfor populationcycles in FraserRiver sockeyesalmon (Oncorhynchus

nerka). CanadianJournal of Fisheriesand Aquatic Science51,1839-1854.

Chadwick, E.M. P. (1982). Stock-recruitmentrelationships for Atlantic salmon (Salmo

salar) in Newfoundland rivers. Canadian Journal of Fisheries and Aquatic

Sciences39,1496-150 1.

107 Chadwick,E. M. P. (1985a). The influenceof spawningstock on productionand yield of

Atlantic salmon in Canadianrivers. Aquaculture and Fisheries Management1,

111-119.

Chadwick, E.M. P. (1985b). Fundamentalresearch problems in the managementof

Atlantic salmon (Salmo salar) in Atlantic Canada. Joumal of Fish Biology 27

(Suppl. A), 9-25.

Chapman,D. W., (1986). Salmonand Steelheadabundance in the Columbia River in the

nineteenthcentury. Transactions of theAmerican Fisheries Society 115,662-670.

Chase,S. (2003). Closing the North American mixed-stockcommercial fishery for wild

Atlantic salmon(Salmo salar L. ). In Salmonat the Edge(Mills, D. ed), pp. 84-92,

Oxford, Blackwell Publishing.

Clark, C.W. (1973). The Economicsof Overexploitation.Science, 181,630-634.

Claytor, R. R. (2000). Conflict resolution in fisheries management using decision rules:an

exampleusing a mixed-stockAtlantic Canadianherring fishery. ICES Journal of

Marine Science 57,1110-1127.

Crozier, W.W. and Kennedy, G.J. A. (1994a). Application of semi-quantitative

electrofishingto juvenile salmonidstock surveys.Joumal of Fish Biology 45,159-

164.

108 Crozier, W.W., KennedyG. J. A. (1994b). Marine exploitation of Atlantic salmon(Salmo

salar L. ) from the River Bush,Northern Ireland. FisheriesResearch 19,141-155.

Crozier, W.W., Kennedy,G. J. A., Boylan, P. & Kennedy,R. (2003). Summaryof salmon

fisheriesand statusof stocksin Northern ireland for 2002. ICES Working Group

Paper no.

Dempson, J.B., Schwarz, C.J., Reddin, D.G., O'Connell, M. F., Mullins, C.C. and

Bourgeois, C.E. (2001). Estimation of marine exploitation rates on Atlantic salmon

(Salmosalar L. ) stocksin Newfoundland,Canada. ICES Journal ofMarine Science

58,331-341.

Dickson, R.R. and Turrell, W.R. (2000). The NAO: The Dominant AtmosphericProcess

Affecting OceanicVariability in Home, Middle and Distant Waters of European

Atlantic Salmon, In 77zeOcean Life of Atlantic salmon: Environmental and

biological factors influencing survival. (Mills, D., ed.), pp. 92-115. London:

FishingNews Books.

Drinkwater, K. F. (2000). Changes in Ocean Climate and its General Effect on Fisheries:

Examplesfrom the North-west Atlantic, In The Ocean Life of Atlantic salmon:

Environmentaland biologicalfactors influencingsurvival. (Mills, D., ed.), pp. 116-

136. London:Fishing NewsBooks.

109 Dumas, J. & Prouzet,P. (2003). Variability of demographicparameters and population

dynamicsof Atlantic salmon(Salmo salar L. ) in a southwestFrench river. ICES

Journal ofMarine Science60,356-370.

Egglishaw,H. J., and Shackley,P. E. (1977). Growth, survival and productionof juvenile

salmonin a Scottishstream, 1966-1975. Joumal of Fish Biology 11,647-672.

Egglishaw,H. J., and Shackley,P. E. (1985). Factorsgoverning the productionof juvenile

Atlantic salmon in Scottish streams. Joumal of Fish Biology 27,27-33.

Ekerholm, P., Oksanen,L. & Oksanen,T. (2001). Long-term dynamics of voles and

lemmingsat the timberline and above the willow limit as a test of hypothesison

trophic interactions.Ecography 24,555-568.

Einarsson, S.M., Gudbergsson,G. (2003). The effects of the net fishery closure on angling

catch in the River Hvita, Iceland. Fisheries Management and Ecology 10,73-78.

Einum, S., Fleming, I.A., Cote, I. M. & Reynolds,J. D. (2003). Population stability in

salmon species:effects of population size and female reproductive allocation.

Joumal ofAnimal Ecology 72,811-82 1.

Elliott, J.M. (1985). Population dynamicsof migratory trout, (Salmo truua) in a Lake

District stream,1966-83, and their implicationsfor fisheriesmanagement. Joumal

of Fish Biology 27,3543.

110 Elliott, J.M. (1989). The natural regulation of numbers and growth in contrasting

populationsof brown trout, Salmotrutta, in two Lake District streams.Freshwater

Biology 21,7-19.

Elliott, J.M. (1994).Quantitative Ecology and the Brovm Trout. Oxford Seriesin Ecology

andEvolution, Oxford: Oxford University Press.

Elliott, J.M., Hurley, M. A. & Elliott, J.A. (1997). Variable effects of droughts on the

density of a sea-trout population over 30 years. Joumal of Applied Ecology 34,

129-138.

Elliott, J.M. and Hurley, M. A. (1997). A functional model for maximum growth of

Atlantic salmonparr, Salmosalar L., from two populations in northwestEngland.

Functional Ecology 11,592-603.

Elliott, J.M. & Hurley, M. A. (1998). An individual-basedmodel for predicting the

emergence period of sea-trout fry in a Lake District stream. Joumal of Fish

Biology 53,414-433.

Elliott, J.M. (2001). The relative role of density in the stock-recruitment relationship of

salmonids. In Stock, recruitment and reference points. assess7nentand

managementof Atlantic salmon,( Prevost,E. & Chaput,G. ed.), pp. 25-66, Paris:

III Elliott, J.M., Hurley, M.A. & Maberly, S.C. (2000). The emergenceof seatrout fry in a

Lake District streamcorrelates with the North Atlantic Oscillation.Joumal of Fish

Biology 56,208-210.

Elliott, S.R., Coe, T.A., Helfield, J.M. & Naiman, R.J. (1998). Spatial variation in

environmentalcharacteristics of Atlantic salmon(Salmo salar) rivers. Canadian

Journal of Fisheriesand Aquatic Sciences55 (Suppl. 1), 267-280.

Elson, P.F. & Tuomi, A. L. W. (1975). 77zeFoyle Fisheries, A New Basis for Rational

Management.Londonderry: Foyle FisheriesCommission.

Frederiksen,M., Lebreton,J. -D. and Bregnballe,T. (2001) The interplay betweenculling

and density-dependencein the great cormorant:a modelling approach. Journal of

Applied Ecology 38,617-627.

Friedland, K.D. (1998). Oceanclimate influenceson critical Atlantic salmon (Salmo

salar) life history events. Canadian Journal of Fisheries and Aquatic Sciences

55 (Suppl. 1), 119-130.

Friedland, K. D., HansenL. P., Dunkley, D. A. & MacLean, J.C. (2000). Linkage

between ocean climate, post-smolt growth, and survival of Atlantic salmon

(Salmosalar L. ) in the North Sea area. ICES Journal of Marine Science57,

419-429.

112 Friedland,K. D., Reddin,D. G. & Castonguay,M. (2003a). Oceanthermal conditions in

the post-smoltnursery of North American Atlantic salmon. ICES Joumal of

Marine Science60,343-355.

Friedland, K.D., Reddin, D. G., McMenemy, J.R. & Drinkwater, K.F. (2003b).

Multidecadaltrends in North American Atlantic salmon(Salmo salar) stocksand

climate trends relevant to juvenile survival. CanadianJournal of Fisheries and

Aquatic Sciences 60,563-583.

Forchhammer,M. C., Post,E. & Stenseth,N. C. (2002). North Atlantic Oscillation timing

of long- and short-distance migration. Joumal ofAnimal Ecology71,1002-1014.

Gardiner, R., and Shackley, (1991). Stock and recruitment and inversely density

dependentgrowth of salmon,in a Scottishstream. Joumal of Fish Biology 38,69 1-

696.

Gee,A. S., Milner, N.J. and Hernsworth,R. J. (1978). The effect of densityon mortality in

juvenile Atlantic salmon (Salmo salar). Joumal ofAnimal Ecology 47,497-505.

Gillett, N.P., Graf, H.F. & Osborne,T. J. (2003). Climate changeand the North Atlantic

Oscillation, In 77zeNorth Atlantic Oscillation: Climate Significance and

EnvironmentalImpact. GeophysicalMonograph Series (Hun-ell, J. W., Kushnir, Y.,

Ottersen,G. andVisbeck, M., eds.) 134,193-209.

113 Gillooly, J.F., O'Keefe,T. C., Newman,S. P. and Baylis, J.R. (2000). A long-termview of

density-dependentrecruitment in smallmouthbass from Nebish Lake, Wisconsin.

Joumal of Fish Biology. 56,542-55 1.

Grant, J.W. A., Steingrimsson,S. O., Keeley,E. R. & Cunjak, R.A. (1998). Implicationsof

territory size for the measurementand prediction of salmonid abundancein

streams.Canadian Joumal of FisheriesAquatic Science55 (Suppl. 1), 181-190.

Haldane,J. B. S. (1953). Animal populationsand their regulation.New Biology 15,9-24.

Haldane,J. B. S. (1956). The relation betweendensity regulation and natural selection.

Proceedings of the Royal Society B 145,306-308.

Hansen,L. P. & Quinn, T.P. (1998). The marine phaseof the Atlantic salmon (Salmo

salar) life cycle, with comparisons to Pacific salmon. Canadian Journal of

Fisheries and Aquatic Sciences55 (Suppl. 1), 104-118.

Harris, M.P. (1970). Territory limiting the size of the breeding population of the

Oystercatcher (Haematopus ostralegus) -a removal experiment. Journal ofAnimal

Ecology 39,707-713.

Harwood, A. J., Annstrong, J.D., Metcalfe, N.B. & Griffiths, S.W. (2003). Does

dominancestatus correlate with growth in wild stream-dwellingAtlantic salmon

(Salmosalar)?. BehavioralEcology 14,902-908.

114 Hassell, M.P. & May, R.M. (1990). Population regulation and dynamics.Philosophical

Transactions of the Royal Society of London 330,121-304.

Hay, D.W. (1984). The relationshipbetween redd counts and the numbersof spawning

salmonin the Gimock Bum (Scotland), InternationalCouncil for the Exploration

of the Sea,C. W. 1984/M:22, Anadromousand CatadromousFish Committee.

Hay, D.W. (1987). The relationshipbetween redd counts and the numbersof spawning

salmonin the Gimock Bum, Scotland.Journal du Conseil43,146-148.

Heggberget,T. G. (1988). Timing of spawning in Norwegian salmon (Salmo salar L. ).

CanadianJoumal of Fisheriesand Aquatic Science45,845-849.

Heggenes,J., and Saltveit, S.J. (1990). Seasonaland spatial microhabitat*selection and

segregation in young Atlantic salmon and brown trout in a Norwegian river.

Joumalof Fish Biology 36,707-720.

Hicks, B.J. (2003).Rock type and channelgradient structure salmonid populations in the

Oregon Coast Range. Transactions of the American Fisheries Society 132,468-

482.

Hilborn, R. (1997). Comment:Recruitment paradigms for fish stocks. CanadianJournal

of Fisheriesand Aquatic Sciences54,984-985.

115 Hill, M. O. (1979). TWINSPAN -A FORTRAN programfor arrangingmultivariate data

in orderedtwo-way tablesby classificationo the individualsand attributes.Ecology

and Systematics,New York: Cornell University.

Hill, M. F., Botsford, L. W. & Hastings, A. (2003). The effects of spawning age distribution

on salmon persistenceinfluctuating environments. 7he Journal ofAnimal Ecology

72,736-744.

Hilton J., Welton J.S., ClarkeR. T. & Ladle M. (2001). An assessmentof the potentialfor

the application of two simple models to Atlantic salmon, Salmo salar, stock

managementin chalk rivers. Fisheries Management and Ecology, 8,189-205.

Holm, M., Holst, J.C. & Hansen,L. P. (2000). Spatial and temporal distribution of post-

smolts of Atlantic salmon (Salmosalar L. ) in the Norwegian Sea and adjacent

areas. ICES Journal ofMarine Science 57,955-964.

Horton, R.E. (1945). Erosionaldevelopment of streamsand their drainagebasins: a hydro-

physical approach to quantitative morphology. Bulletin of the Geological Society

ofAmerica 56,275-350.

Hurrell, J.W., Kushnir, Y., Ottersen,G. & Visbeck, M. (2003).An overview of the North

Atlantic Oscillation. 77teNorth Atlantic Oscillation: Climate Significance and

Environmental Impact. Geophysical Monograph Series (eds. Hurrell, J.W.,

Kushnir, Y., Ottersen,G. andVisbeck, M. ) 134,1-35.

116 Hurrell, J.W. (1995). Decadal trends in the North Atlantic Oscillation: regional

temperaturesand precipitation. Science 269,676-679.

Hutchinson P. & Mills D. H. (2000). Executive Summary. In 71e Ocean Life of Atlantic

salmon: Environmental and biological factors influencing survival. (D. Mills ed.)

pp. 7-18, London:Fishing News Books.

Isaak, D. J., Thurow, R.F., Rieman, B. E. & Dunham, J.B. (2003). Temporal variation in

synchronyamong chinook salmon(Oncorhynchus tshawytscha) redd countsfrom a

wildemess area in centml Idaho. Canadian Journal of Fisheries and Aquatic

Science60,840-848.

James,G. W. A. andDonald, K. L. (1990). Territory size as a predictorof the upper limit to

population density of juvenile salmonids in streams. Canadian Journal of Fisheries

and Aquatic Science47,1724-1737.

Jenkins,D., Watson,A. & Miller, G.R. (1963).Population studies on Red GrouseLagopus

lagopus scoticus (lath. ) in north-east Scotland. Joumal ofAnimal Ecology 32,317-

376.

117 Jensen,A. J., Zubchenko,A. V., Heggberget,T. G., Hvidsten, N. A., Johnsen,B. 0.,

Kuzmin, 0., Loenko,A. A., Lund, R. A., Martynov, V. G., Nxsje, T. F., Sharov,A.

F., and Okland, F. (1999). Cessationof the Norwegiandrift net fishery: changes

observedin Norwegianand Russianpopulations of Atlantic salmon.ICES Journal

of Marine Science56,84-95.

Jones,J. W. (1959).7he Salmon, London: Collins, St JamesPlace.

Jones, M., Laurila, A., Peuhkuri, N., Piironen, J. & Seppa, T. (2002). Timing an

ontogeneticniche shift: responsesof emerging salmon alevins to chemical cues

from predatorsand competitors. Oikos 102,155-163.

Johnston,P. (2002). Benefitsof nettingbuy-out. Trout & SalmonJune, 23-25.

Jonsson, N., Jonsson,B. & Hansen, L. P. (2003). The marine survival and growth of wild

and hatchery reared Atlantic salmon (Salmo salar L. ). Joumal ofApplied Ecology

40,900-911.

KennedyG. J. A. (1983). Someobservations on salmonidecology in upland streams,Irish

Fisheries Investigations Series A. 23.

Kennedy, G.J. A. and Strange,C. D. (1982). The distribution of salmonids in upland

streamsin relationto depthand gradient. Joumal of Fish Biology 20,579-591.

118 Kennedy, G.J. A. and Strange,C. D. (1986a). The effects of intra- and inter-specific

competitionon the survival and growth of stockedjuvenile Atlantic salmon,Salmo

salar L., and residenttrout, Salmotrutta L., in an upland stream. Joumal of Fish

Biology 28,479-489.

Kennedy, G.J. A. and Strange,C. D. (1986b). The effects of intra- and inter-specific

competitionon the distributionof stockedjuvenile Atlantic salmon,Salmo salar L.,

in relation to depth and gradient, in an upland trout, SaInw trutta L., stream.

Joumal of Fish Biology 29,199-214.

Kennedy,G. J. A. & Crozier, W.W. (1993). Juvenile Atlantic salmon (Salmosalar L. ) -

productionand prediction. In Productionof JuvenileAtlantic salmon(Salmo salar

L) in natural waters. Canadian Special Publication of Fisheries and Aquatic

Sciences118,179-187. (eds.Gibson, R. J. & Cutting, R.E. ).

Kennedy, G.J. A. & Crozier, W.W. (1997). What is the value of a wild salmon smolt,

(SalmosalarL. )? Fisheries Management and Ecology 4,103-110.

Klemetsen,A., Amundsen,P. A., Dempson,J. B., Jonsson,B., Jonsson,N., O'Connell,

M. F. & Mortensen,E. (2003). Atlantic salmon (Salmo salar L. ), brown trout

(Salmotmtta L. ) and Arctic chaff (Salvelinusalpinus L. ): a review of aspectsof

their life histories. Ecologyof FreshwaterFish 12,1-59.

119 Kocik, J.F. & Ferreri, C.P. (1998). Juvenileproduction variation in salmonids:population

dynamics, habitat, and the role of spatial relationships. Canadian Journal of

Fisheries and Aquatic Sciences55 (Suppl. 1), 191-200.

Krebs, J.R. (1970). Regulation of numbers in the great tit (Aves: Passeriformes). Joumal

of Zoology (U)ndon) 162,317-333.

Lachin, J.M. (2000). Biostatistical Methods - The Assessmentof Relative Risks. New York: Wiley.

Langeland, A. & Pedersen,T. (2000). A 27-year study of brown trout population

dynamicsand exploitation in Lake Songsjoen,central Norway. Journal of Fish

Biology 57,1227-1244.

Laurian C.; Ouellet J-P.; Courtois R.; Breton L.; St-Onge S. (2000). Effects of intensive

harvesting on moose reproduction. Joumal ofApplied Ecology, 37,515-53 1.

Logerwell, E.A., Mantua, N., Lawson, P.W., Francis, R.C. & Agostini, V.N. (2003).

Tracking environmental processesin the coastal zone for understandingand

predicting Oregon coho (Oncorhynchus kisutch) marine survival. Fisheries

Oceanography 12,554-568.

Magee,J. A., Obedzinski,M., McConnick, S.D. & Kocik, JR (2003). Effects of episodic

acidification on Atlantic salmon (Salmo salar) smolts. Canadian Journal of

Fisheriesand Aquatic Sciences60,214-22 1.

120 Marschall,E. A. and Crowder,L. B. (1995). Density dependentsurvival as a function of

size in juvenile salmonids in streams. Canadian Journal of Fisheries and Aquatic

Science52,136-140.

Martin, J.H. A. & Mitchell, K.A. (1985). Influence of seatemperature upon the numbers

of grilse and multi-sea-winterAtlantic salmon(Salmo salar) caught in the vicinity

of the River Dee (Aberdeenshire).Canadian Journal of Fisheries and Aquatic

Science42,1513-152 1.

Milner, N.J., Wyatt, R.J. and Broad, K. (1998). HABSCORE - applicationsand future

developmentsof related habitat models. Aquatic Conservation: Marine and

FreshwaterEcosystems 8,633-644.

Milner, N.J., Elliott, J.M., Armstrong, J.D., Gardiner, R., Welton, J.S. and Ladle, M.

(2003). The natural control of salmonand trout populationsin streams.Fisheries

Research 62,111-125.

McCain, M., Fuller, D., DeckerL., and OvertonK. (1990). StreamHabitat Classification

and Inventory Proceduresfor Northern California. FHR Currents, R-5s Fish

Habitat RelationshipsTechnical Bulletin, No. 1, U. S. Department of Agriculture,

Forest Service, Pacific Southwest Region.

121 McCormick, S.D., Hansen, L.P., Quinn,T. P. & Saunders,R. L. (1998). Movement,

migration and smolting of Atlantic salmon (Salmo salar). Canadian Journal of

Fisheries and Aquatic Sciences55 (Suppl. 1), 77-92.

McFarlane, G.A., King, J.R. & Beamish, RJ. (2000). Have there been recent changes in

climate? Ask the fish. Progress in Oceanography 47,147-169.

McKinnell, S.M. & Karlstrom, 0. (1999). Spatial and temporal covariation in the

recruitmentand abundanceof Atlantic salmonpopulations in the Baltic Sea. ICES

Journal of Marine Science56,433-443.

Mikio Inoue, Shigeru Nakano and Futoshi Nakamura (1997). Juvenile masu salmon

(Oncorhynchusmasou) abundanceand stream habitat relationships in northern

Japan.Canadian Journal of Fisheriesand Aquatic Science54,1331 - 1341.

Milner, N.J., Wyatt, R.J., and Scott, M. D. (1993). The use of habitat models to interpret

salmonid distribution and abundancein streams. In ProceedingsFSBI Annual

Symposium,Conwy.

Mueter, F.J., Peterman, R. M. & Pyper, B. J. (2002). Opposite effects of ocean temperature

on survival ratesof 120 stocksof Pacific salmon(Oncorhynchus spp. ) in northern

and southernareas. CanadianJournal of Fisheriesand Aquatic Sciences59,456-

463.

122 Myers, R.A., Barrowman,LA., Hutchings,J. A., and Rosenberg,A. A. (1995). Population

dynamicsof exploitedfish stocksat low populationlevels. Science 269,1106-1108.

Myers, R.A., Rosenberg,A. A., Mace,P. M., Barrowman,and Restrepo,V. R. (1994).ICES

Journal ofMarine Science51,191-205.

Newton,1. (1998). PopulationLimitation in Birds. London:Academic Press.

O'Connell, M. F., Dempson, J.B. and Reddin, D. G. (1992). Evaluation of the impacts of

major managementchanges in the Atlantic salmon (Salmo salar L. ) fisheries of

Newfoundlandand Labrador,Canada, 1984-1988. ICES Journal of Marine Science

49,69-87

Olsen,P. D. and Olsen,J. (1989). Breedingof the Peregrinefalcon (Falco peregrinus):III.

Weather, nest quality and breeding success.Emu 89,6-14.

Parrish, D. L., Behnke, R. J., Gephard, S.R., McConnick, S.D. & Reeves, G.H. (1998).

Why aren't there more Atlantic salmon (Salmo salar)? Canadian Journal of

Fisheries and Aquatic Sciences55 (Suppl. 1), 281-287.

Pess,G. R., Montgomery,D. R., Steel,E. A., Bilby, R.E., Feist, BE and Greenberg,H. M.

(2002). Landscapecharacteristics, land use, and coho salmon (Onchorhynchus

kisutch) abundance,Snohomish River, Washington,U. S.A.. CanadianJournal of

Fisheriesand Aquatic Science59,613-623.

123 PetermanR. M., Pyper, B.J. & MacGregor,B. W. (2003). Use of the Kalman filter to

reconstruct historical trends in productivity of Bristol Bay sockeye salmon

(Oncorhynchusnerka). CanadianJournal of FisheriesScience 60,809-824.

Pierce,G. J. & Boyle, P.R. (2003). Empirical modelling of interannualtrends in abundance

of squid(Loligoforbesi) in Scottishwaters. FisheriesResearch 59,305-326.

Poff, N.L. & Huryn, A.D. (1998). Multi-scale determinantsof secondaryproduction in

Atlantic salmon (Salmo salar). Canadian Journal of Fisheries and Aquatic

Sciences55 (Suppl. 1), 201-217.

Punt,A. E., and Hilborn, R. (1997). Fisheriesstock assessmentand decisionanalysis: The

Bayesianapproach. Reviews in Fish Biology and Fisheries7,35-63.

QuizilbashM. (2001). SustainableDevelopment: Concepts and Rankings.7he Journal of

Development Studies 37,134-16 1.

Reddin D.G., Helbig, J., Thomas, A., Whitehouse, B.G. & Friedland, K. D. (2000).

Survival of Atlantic Salmon(Salmo salar L. ) related to Marine Climate. In 7he

Ocean Life of Atlantic salmon: Environmental and biological factors influencing

survivaL(D. Mills ed.) pp. 88-91,London: Fishing NewsBooks.

Reddin, D.J., Friedland, K.D., Downton, P., Dempson, B.J. & Mullins, C.C. (2004).

Thermal habitat experiencedby Atlantic salmon (Salmosalar L. ) kelts in coastal

Newfoundlandwaters. FisheriesOceanography 13,24-35.

124 Reid, P.C, & Planque, B. (2000). Long-term planktonic variations and the climate of the

North Atlantic. In 7he Ocean Life of Atlantic salmon: Environmental and

biologicalfactors influencingsurvivaL (D. Mills ed.) pp.153-169, London: Fishing

News Books.

Ricker, W.E. (1954). Stock and Recruitment.Journal FisheriesResearch Board Canada

11,559-623.

Ritter, J.A. (1997). Atlantic Salmon Maritimes Region Overview. DFO ScienceStock

Status Report, D3-14.

Rodrfguez-SAnchez,R., Lluch-Belda, D., Villalobos, H. & Ortega-Garcia,S. (2002).

Dynamic geographyof small pelagic fish populations in the California Current

Systemon the regimetime scale(1931-1997). CanadianJournal of Fisheriesand

Aquatic Science59,1980-1988.

Romakkanierni, A., Perd, I., Karlsson, L., Jutila, E., Carlsson, U. & Pakarinen, T. (2003).

Developmentof wild Atlantic salmonstocks in the rivers of the northernBaltic Sea

in responseto managementmeasures. ICES Journal of Marine Science60,329-

342.

Ross, M. R., and Almeida, F. P. (1986). Density-dependentgrowth of silver hakes.

Transactionsof the AmericanFisheries Society 115,548-554.

125 Ruggerone,G. T. & Rogers,D. E. (2003). Multi-year effects of high densitiesof sockeye

salmonspawners on juvenile salmongrowth and survival: a casestudy from Exxon

Valdezoil spill. FisheriesResearch 63,379-392.

Ruggerone,G. T., Zimmennann,M., Myers, K.W., Nielsen, J.L. & Rogers,D. E. (2003).

Competitionbetween Asian pink salmon (Oncorhynchusgorbuscha) and Alaskan

sockeyesalmon (0. nerka) in the North Pacific Ocean.Fisheries Oceanography 12,

209-219.

Schnute,J. T., and Kronlund, A.R. (1996). A managementorientated approach to stock

recruitmentanalysis. Canadian Joumal of Fisheriesand Aquatic Science53,128 1-

1293.

Shearer,W. (1992). 7he Atlantic Salmon: Natural History, Exploitation and Future

Management. London: Fishing News Books.

Silveria, L., Rodrigues,F. H. G., Ja.A como, A. T.A. and Filho, J.A. F. D. (1999). Impact of

wildfires on the megafaunaof Emas National Park, central Brazil. Oryx 33,108-

114.

Sinclair, A. E.G. (1989). The regulation of animal populations. In Ecological concepts.

British Ecological Society Symposium(Cherrett, M., ed.), pp. 197-241,Oxford:

Blackwell ScientificPublications.

126 Small, 1. (1991).Exploring dataprovided by angling for salmonidsin the British Isles. In

in Freshwater Fisheries Catch Effort Sampling Strategies - their Application

Management(Cowx, I. G. ed.), Oxford: Blackwell Scientific Publishing.

Smith, J.N. M., Taitt, M.J., Rogers,C. M., Arcese,P., Keller, L.F., Cassidy,A. L. E.V. and

Hochachka,W. M. (1996). A metapopulationapproach to the populationbiology of

the SongSparrow Melospiza melodia. This138,120-128.

Smith, A. D.M., and Walters, C.J. (1981). Adaptive Managementof Stock-Recruitment

Systems.Canadian Journal of Fisheriesand Aquatic Science38,690-703.

Solomon,D. J. (1985). Salmonstock and recruitment,and stock enhancement.Joumal of

Fish Biology 27 (Suppl. A), 45-57.

Stenning,M. J., Harvey, P.H. and Campbell,B. (1988). Searchingfor density-dependent

regulation in a population of pied flycatchers Ficeduld hypoleuca Pallas. Journal of

Animal Ecology 57,307-317.

Titus, R.G. and Mosegaard,H. (1992).Fluctuating recruitment and variable life history of

migratory brown trout in a small unstable stream. Journal of Fish Biology 41,239-

255.

Walters, C.J. (1981). Optimum Escapementsin the Face of Alternative Recruitment

Hypotheses.Canadian Joumal of Fisheriesand Aquatic Science38,678-689.

127 Walters, C.J., and Ludwig, D. (1981). Effects of MeasurementErrors on the Assessment

of Stock-RecruitmentRelationships. Canadian Journal of Fisheries and Aquatic

Science38,704-7 10.

Waring, C.P. & Moore, A. (2004). The effect of atrazine on Atlantic salmon (Salmo salar)

smoltsin fresh waterand after seawater transfer.Aquatic Toxicology66,93-104.

Whitehead,P. (1992). Examplesof recent models in environmentalimpact assessment.

Journal of the Institution of Waterand EnvironmentalManagement 6,475-484.

Whitehead, A. (2003) Progress in ending mixed-stock interceptory fisheries: United

Kingdom. In Salmon at the Edge, (Mills, D. ed), pp.78-83, Oxford, Blackwell

Publishing.

Wright, JR, Armitage, P.D., Furse, M. T., and Moss, D. (1989). Prediction of invertebrate

communities using stream measurements. Regulated Rivers. Research &

Management 4,147-155.

Wright, U., Blackburn, J.H., Gunn, R. J.M., Furse, M. T., Armitage, P.D., Winder, J.M.,

and Symes,K. L. (1996). Macroinvertebratefrequency data for the RIVPACSIII

SITES in Great Britain and their use in conservation evaluation. Aquatic

Conservation:Marine and FreshwaterEcosystems 6,141-167.

128