ecology and physiology of chaetognaths (semi-gelatinous ... · les taux de prédation quotidiens...

158
Ecology of chaetognaths (semi-gelatinous zooplankton) in Arctic waters Thèse Jordan Grigor Doctorat interuniversitaire en océanographie Philosophiæ doctor (Ph. D.) Québec, Canada © Jordan Grigor, 2017

Upload: others

Post on 18-Jul-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

Ecology of chaetognaths (semi-gelatinous zooplankton) in Arctic waters

Thèse

Jordan Grigor

Doctorat interuniversitaire en océanographie

Philosophiæ doctor (Ph. D.)

Québec, Canada

© Jordan Grigor, 2017

Page 2: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

Ecology of chaetognaths (semi-gelatinous zooplankton) in Arctic waters

Thèse

Jordan Grigor

Sous la direction de :

Louis Fortier, directeur de recherche

Page 3: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

iii

Résumé

Les chaetognathes sont d’importants membres des communautés mésozooplanctoniques de

l’Arctique en ce qui a trait à l'abondance et à la biomasse. Les chaetognathes de l’Arctique

se répartissent en trois espèces principales qui sont considérées comme étant strictement

carnivores : Eukrohnia hamata, Parasagitta elegans et Pseudosagitta maxima. Cette étude

utilise un ensemble de données de filet planctoniques recueillies sur une période de 5 ans

dans les régions européennes, canadiennes et de l'Alaska de l’Arctique (2007, 2008, 2012,

2013, 2014) et comprend un cycle annuel complet dans l'Arctique canadien (2007-2008), le

but étant d’améliorer notre compréhension sur les distributions, les cycles de vie et les

stratégies d'alimentation du E. hamata et du P. elegans. Dans la présente thèse, les points

suivants seront abordés : (1) la stratégie d'alimentation et la maturité du P. elegans dans

l'Arctique européen durant la nuit polaire en 2012 et 2013, (2) les cycles de croissance et de

reproduction, les stratégies d'alimentation et les distributions verticales du E. hamata et du

P. elegans dans l'Arctique canadien de 2007 à 2008, et (3) les différences spatiales dans les

stratégies d'alimentation du E. hamata et du P. elegans à l'automne 2014. Afin d’étudier leurs

stratégies d'alimentation, des analyses de contenu du tube digestif ainsi que des techniques

biochimiques ont été utilisées. Dans l'Arctique canadien, le E. hamata et le P. elegans vivent

tous deux pendant environ 2 ans. Le P. elegans colonise principalement les eaux

épipélagiques, tandis que le E. hamata colonise principalement les eaux mésopélagiques.

Dans cette région, P. elegans se reproduit en continue de l'été au début de l'hiver, dans la

période de forte biomasse de copépodes, qui constituent ses proies, dans les eaux proches de

la surface, un mode de reproduction basé sur l’apport immédiat d’énergie. Cependant, les

résultats ont révélé que E. hamata a engendré des couvées distinctes dont on peut voir

l’évolution au cours de fenêtres de reproduction séparées, à la fois durant les périodes de

printemps-été et d’automne-hiver, ce qui suggère une reproduction basée sur les réserves.

Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif

sont généralement restés faibles pour les deux espèces de chaetognathes. Toutefois, pour E.

hamata et P. elegans, les taux de prédation inférés en été-automne ont dépassé ceux de

l’hiver-printemps. Des études d’alimentation ont révélé que E. hamata consomme de la

matière organique particulaire (éventuellement des chutes de neige marine) tout au long de

l'année, mais surtout en été, alors que le P. elegans se nourrit différemment. Les deux espèces

Page 4: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

iv

sont caractérisées par une forte croissance estivale. La croissance hivernale du P. elegans

était grandement restreinte, tandis que celle du E hamata l’était moindrement. En somme, les

différences dans la façon dont les lipides et la neige marine sont utilisés par les deux espèces

pourraient expliquer les différences dans leurs cycles de reproduction et leurs patrons de

croissance saisonnière.

Page 5: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

v

Abstract

Chaetognaths are important members of Arctic mesozooplankton communities in terms of

abundance and biomass. Despite this, the bulk of seasonal studies have focused on grazing

copepods. Arctic chaetognaths comprise three major species which are thought to be strict

carnivores: Eukrohnia hamata, Parasagitta elegans and Pseudosagitta maxima. This thesis

uses datasets collected from plankton net sampling during five years in European, Canadian

and Alaskan areas of the Arctic (2007, 2008, 2012, 2013, 2014) and includes a full annual

cycle in the Canadian Arctic (2007-2008), the purpose being to improve our understanding

of the distributions, life history and feeding strategies of E. hamata and P. elegans. The

following topics are addressed: (1) the feeding strategy and maturity of P. elegans in the

European Arctic during the polar night in 2012 and 2013; (2) the growth, breeding cycles,

feeding strategies and vertical distributions of E. hamata and P. elegans, in the Canadian

Arctic from 2007 to 2008; and (3) spatial differences in the feeding strategies of E. hamata

and P. elegans in autumn 2014. To investigate feeding strategies, a combination of gut

contents and biochemical techniques was used. In the Canadian Arctic, both E. hamata and

P. elegans live for around 2 years. P. elegans mainly colonized epi-pelagic waters, whereas

E. hamata mainly colonized meso-pelagic waters. In this region, P. elegans reproduced

continuously from summer to early winter when copepod prey peak in near-surface waters.

This is characteristic of income breeders. However, results for E. hamata revealed that this

species spawned distinct and traceable broods during separate reproductive windows in both

spring-summer and autumn-winter, suggesting capital breeding. Daily predation rates

inferred from gut content analyses appeared to be generally low in the two chaetognath

species, though inferred predation rates in summer-autumn exceeded those in winter-spring.

Feeding studies revealed that E. hamata consumed particulate organic matter (possibly

falling marine snow) throughout the year but especially in the summer, whereas P. elegans

did not feed in this way. High summer growth seems to be a characteristic of both these

species. Growth during winter was highly restricted in P. elegans, to a lesser extent in E.

hamata. In summary, differences in how lipids and marine snow are utilised by the two

species could explain differences in their breeding cycles and seasonal growth patterns.

Page 6: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

vi

Table of contents

Résumé iii

Abstract v

Acknowledgements xviii

Foreword xx

1. Chapter 1 – General Introduction 1

1.1 The Arctic Ocean 1

1.1.1 Physical environment 2

1.1.2 Arctic marine ecosystems 3

1.2 Zooplankton and their polar adaptations 5

1.3 Chaetognaths 8

1.3.1 Arctic species and distributions 9

1.3.2 Lifespans 10

1.3.3 Reproductive strategy 11

1.3.4 Oil vacuoles in Eukrohnia spp. 11

1.3.5 Feeding strategies and trophic importance 12

1.3.6 Fatty acids and stable isotopes 14

1.4 Climate change and other challenges for Arctic marine life 14

1.5 Study areas 17

1.6 Aims and objectives 18

2. Chapter 2 – Polar night ecology of a pelagic predator, the chaetognath Parasagitta elegans

20

2.1 Résumé 20

2.2 Abstract 21

2.3 Introduction 22

2.4 Method 23

2.4.1 Study area 23

2.4.2 Physical and biological environment 24

2.4.3 Zooplankton sampling 24

2.4.4 Sample processing 25

2.4.5 Gut content analyses 26

2.4.6 Food-containing ratio and feeding rate 26

2.4.7 Stable isotope analyses 27

2.4.8 Determination of trophic level 28

2.4.9 Fatty acid analyses 28

2.4.10 Mid-winter maturity status 29

2.5 Results 30

2.5.1 Chaetognath abundance and prey field 30

Page 7: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

vii

2.5.2 Gut contents 32

2.5.3 Body composition 34

2.5.4 Mid-winter maturity status 36

2.6 Discussion 36

2.6.1 Studies during the polar night 36

2.6.2 Feeding activity and rates 36

2.6.3 Energetics 38

2.6.4 Lipid profile 38

2.6.5 Reproduction 40

2.7 Concluding remarks 41

3. Chapter 3 – Growth and reproduction of the chaetognaths Eukrohnia hamata and

Parasagitta elegans in the Canadian Arctic Ocean: capital breeding versus income breeding

42

3.1 Résumé 42

3.2 Abstract 43

3.3 Introduction 44

3.4 Method 46

3.4.1 Study area 46

3.4.2 Sampling 47

3.4.3 Chaetognath body size and sampler efficiency 48

3.4.4 Hatching and cohort development 49

3.4.5 Estimation of maturity and oil vacuole area 50

3.4.6 Vertical distributions 50

3.5 Results 51

3.5.1 Physical environment and primary production in the Amundsen Gulf 51

3.5.2 Physical environment and primary production in autumn (other sampling

locations) 51

3.5.3 Abundances and vertical distributions of chaetognath species 52

3.5.4 Length distributions and sampler efficiency 53

3.5.5 Timing of reproduction 54

3.5.6 Eukrohnia hamata length cohorts and life cycle 55

3.5.7 Parasagitta elegans length cohorts and life cycle 62

3.6 Discussion 67

3.6.1 Chaetognath cohort interpretation and lifespans 67

3.6.2 Resource partitioning in the sympatric Eukrohnia hamata and Parasagitta elegans

67

3.6.3 The potential role of lipid reserves: contrasting growth in the two species 69

3.6.4 Maturation 71

3.6.5 Capital versus income breeding in a warming Arctic 72

3.7 Conclusions 73

4. Chapter 4 – Feeding strategies of arctic chaetognaths: are they really “tigers of the

plankton”? 74

Page 8: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

viii

4.1 Résumé 74

4.2 Abstract 75

4.3 Introduction 76

4.4 Method 77

4.4.1 Study areas 77

4.4.2 Sampling in the Amundsen Gulf 79

4.4.3 Sampling in the Chukchi Sea and Baffin Bay 80

4.4.4 Abundance of zooplankton 80

4.4.5 Gut contents 81

4.4.6 Fatty acids 82

4.4.7 Carbon and nitrogen 82

4.5 Results 84

4.5.1 Amundsen Gulf 84

4.5.1.1 Phenology of algae blooms 84

4.5.1.2 Zooplankton community 84

4.5.1.3 Visible prey items and predation rates 86

4.5.1.4 Lipid droplets and detritus 87

4.5.1.5 Scanning Electron Microscope observations 89

4.5.2 Chukchi Sea and Baffin Bay 90

4.5.2.1 In-vitro feeding observations 90

4.5.2.2 Lipid amounts 90

4.5.2.3 Fatty acid profiles: species differences 91

4.5.2.4 Fatty acid profiles: station and regional differences in Eukrohnia hamata and

Parasagitta elegans 91

4.5.2.5 Carbon and nitrogen isotopes 94

4.6 Discussion 96

4.6.1 Prey items 96

4.6.2 Predation rates 97

4.6.3 Omnivory in arctic chaetognaths 98

4.6.4 Fatty acids and stable isotopes confirm gut contents 99

4.6.4.1 18:1 (n-9)/(n-7) ratios and Calanus copepod FATMs 99

4.6.4.2 Algal FATMs 100

4.6.4.3 δ13C values 100

4.6.4.4 δ15N values and inferred TLs 101

4.6.4.5 C/N ratios 101

4.6.5 Morphological explanations 102

4.7 Concluding remarks and future studies 102

5. Chapter 5 – General Conclusions 103

5.1 Resource use by arctic chaetognaths 103

5.2 Winter ecology in Svalbard 103

5.3 Life histories, habitats and spawning times 104

5.4 Food resources 105

5.5 Limitations of the study 107

5.5.1 Sampling limitations 107

Page 9: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

ix

5.5.2 Stable isotopes and trophic levels 107

5.5.3 Length cohorts 108

5.6 Future research avenues 108

5.7 A note on climate change 109

Bibliography 110

Appendix A. Information on the zooplankton samples included in Chapter 2 126

Appendix B. Vertical profiles of temperature and salinity in Chapter 2 128

Appendix C. Stations sampled in Chapter 3 129

Appendix D. Vertical profiles of temperature, salinity and chlorophyll a in Chapter 3 131

Appendix E-1. Amundsen Gulf stations sampled in Chapter 4 132

Appendix E-2. Chukchi Sea and Baffin Bay stations sampled in Chapter 4 136

Page 10: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

x

List of tables

Table 2.1 Total water-column abundances (ind. m-3) of a polar night zooplankton community

(Rijpfjorden 2012), ordered according to abundance per taxonomic group. Net sampling

(Multi-Plankton Sampler; 0.25-m2 opening, 0.2-mm mesh) was used. As larger

chaetognaths may have avoided the smaller MPS net (see Grigor et al. 2014),

chaetognath abundances presented here are likely to be underestimates. Mean

abundances for species and copepod stages were first calculated over two hauls (one at

midday and the other at midnight at various depth intervals, see Appendix A), and data

were summed across all sampling depths. Copepod stages are CI-CV, AM (adult male)

and AF (adult female). Functional (feeding) groups were extracted from Søreide et al.

(2003). “Small Calanoida” comprised the following taxa: Acartia longiremis, Aetideidae

CI-CIII, Bradyidius similis, Microcalanus spp. and Pseudocalanus spp. Only taxa with

abundances of ≥0.1 ind. m-3 are shown. 31

Table 2.2 Stable carbon and nitrogen isotope values for Parasagitta elegans sampled by the

MIK (3.14-m2 opening, 1.5-mm mesh) at various trawl depths (20, 30, 35, 60, 75 and

225 m) in Isfjorden and Rijpfjorden (2012). The average δ13C and δ15N composition (‰)

in replicate samples (usually three but six from 225 m in Rijpfjorden) containing 25

pooled individuals (10-50 mm), average proportions of animal dry weight (DW)

comprising carbon and nitrogen (%), and C/N weight ratios. All values are accompanied

by standard deviations. Trophic levels (TL) were calculated for P. elegans from each

fjord from mean δ15N (‰) values (see ‘Method’). 34

Table 2.3 Average fatty acid profile for Parasagitta elegans in 2012. Results are given as

average percentages of the various fatty acids identified across all samples from

Isfjorden and Rijpfjorden (see ‘Method’), alongside the standard deviations. Only fatty

acids with mean percentages of =>0.5 ± 0.1 across both fjords are shown. In Isfjorden,

the mean percentages of 15:0 FA and 18.2(n-6) FA differed between sampling depths

(indicated by a † symbol, one-way ANOVA, P ˂ 0.05). In Rijpfjorden, the proportion of

every fatty acid was similar between sampling depths (one-way ANOVA, P ˃ 0.05). 35

Table 3.1 Timing of peak development of maturity features in autumn and spring broods of

Eukrohnia hamata and summer brood of Parasagitta elegans, and of the oil vacuole in

E. hamata. Ages in months at peak development are given in parentheses. Monthly

collections consisted of up to 397 and 434 E. hamata individuals from its autumn and

spring broods respectively, and 177 P. elegans. 59

Table 4.1 Composition of the mesozooplankton (30 most abundant taxa) sampled in the

Amundsen Gulf from November 2007 to July 2008, based on data from 70 Multinet

hauls. *Taxa could not be identified to species level. 85

Table 4.2 Frequency of first- and second-year Eukrohnia hamata individuals with green lipid

droplets or macroaggregates in guts. Separate results are presented for the autumn

(October) and spring (April) broods of E. hamata (see Chapter 3), from approximate

month of hatching (no gut content data were available for October). 88

Page 11: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

xi

Table 4.3 Mean sample lipid, fatty acid amounts and fatty acid proportions (± 1 SD) in (a)

Eukrohnia hamata and (b) Parasagitta elegans at different stations (stns) in autumn

2014. Sampling locations are abbreviated: CS, Chukchi Sea; SIF, Scott Inlet Fjord; GF,

Gibbs Fjord; SBB, southern Baffin Bay. Also: number (n) of samples analysed and

chaetognaths per sample, and chaetognath body lengths in mm. † Significant species

differences (P ≤ 0.05). 92

Table 4.4 Mean sample carbon (C) and nitrogen (N) masses, C/N ratios, δ13C and δ15N values

and derived trophic levels (TLs) (± 1 SD) in Eukrohnia hamata and Parasagitta elegans

at different stations in autumn 2014. Sampling locations are abbreviated: CS, Chukchi

Sea; SIF, Scott Inlet Fjord; GF, Gibbs Fjord; SBB, southern Baffin Bay. Also: number

(n) of samples analysed and chaetognaths per sample, as well as body lengths of

chaetognaths in mm. 95

Page 12: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

xii

List of figures

Figure 1.1 Bathymetric map of the central Arctic Ocean and several marginal seas

(http://recherchespolaires.inist.fr/?L-ocean-Arctique-physiographie). 1

Figure 1.2 Water masses in the Arctic Ocean. Reproduced from Stein & Macdonald (2004).2

Figure 1.3 The planktonic food web showing protozoan and metazoan consumers at arrow

caps. Adapted from Hopcroft et al. (2008). 4

Figure 1.4 Photographs of living holo-zooplankton captured in-situ by a zooplankton imager

(Lightframe On-sight Key Species Investigation System; Schmid et al. 2016), during

deployments in the Canadian Arctic (summer 2014). 6

Figure 1.5 Typical physiology of a chaetognath belonging to the family Sagittidae.

Reproduced from Margulis & Chapman (2010). 8

Figure 1.6 Photographs of the heads and the bodies of the major arctic chaetognaths. (a)

Eukrohnia hamata with characteristic marsupial sac and oil vacuole, (b) and (c)

Parasagitta elegans and Pseudosagitta maxima without marsupial sacs and oil vacuoles.

All images except P. maxima head from:

http://www.arcodiv.org/watercolumn/Chaetognaths.html (courtesy of Russ Hopcroft). 9

Figure 1.7 Photograph of a chaetognath’s head taken by Scanning Electron Microscopy (left)

with parts labelled (right). Reproduced from Bieri & Thuesen (1990). 12

Figure 1.8 Diagram illustrating the wide variety of sampling activities conducted during the

Circumpolar Flaw Lead System Study (2007-2008) from the icebreaker CCGS

Amundsen and at an ice camp, in order to study Arctic systems. Reproduced from Barber

et al. (2010). 18

Figure 2.1 Map showing the locations of the stations sampled for chaetognaths in January

2012 and 2013. 24

Figure 2.2 Proportions (%) of Parasagitta elegans individuals per haul with gut contents

(FCRmax), identifiable prey (FCRmin), and empty guts in each fjord. The horizontal line

inside each boxplot shows the median of the proportions over multiple haul samples in

a fjord. The lower and upper boxes show the lower and upper quartiles, respectively, and

the vertical lines outside the boxes the differences between these quartiles and the lowest

and highest proportions observed. Each dot represents an outlying data point. nhauls =

numbers of hauls for each fjord. Hauls with ˂3 individuals analysed were not included.

As only one haul was analysed for Rijpfjorden in 2013, full boxplots could not be shown.

See Appendix A for numbers of individuals analysed per haul. 32

Page 13: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

xiii

Figure 2.3 Proportions (%) of feeding Parasagitta elegans individuals per haul in each fjord

with different types of gut content. For details on the features of the boxplots and the

data, see Figure 2.2. 33

Figure 2.4 Gut contents in ascending head width size classes: proportions of Parasagitta

elegans (%) per haul with gut contents and of feeders with each gut content type.

Includes all dissected specimens from Isfjorden (50 individuals, 4 hauls) and Rijpfjorden

(152 individuals, 13 hauls) in 2012. For details on the features of the boxplots and the

data, see Figure 2.2. ninds. = total numbers of individuals for each size class. 33

Figure 3.1 Bathymetric maps of the Canadian Arctic Ocean indicating the regions, and

positions of stations (black circles) where chaetognaths were sampled from August 2007

to September 2008. Station IDs provided. 47

Figure 3.2 Illustrations and photographs of Eukrohnia hamata and Parasagitta elegans. (a)

Diagrams of the two species indicating maturity features and the centrally-positioned oil

vacuole in E. hamata. (b) Photographs of live E. hamata (top) and P. elegans (bottom)

taken in-situ by a zooplankton imager (Schmid et al. 2016) in the Canadian Arctic.

Specimens ~20 mm. (c) Photograph of oocytes in a stained P. elegans individual. (d)

Photograph of ovaries in an E. hamata individual (imaged in-situ). (e) Photograph of tail

sperm, seminal vesicles and seminal receptacles in a stained P. elegans individual. 49

Figure 3.3 Relative frequency of Eukrohnia hamata, Parasagitta elegans and Pseudosagitta

maxima in relation to bottom depth at Amundsen Gulf stations. Multinet and square-

conical net collections included. b) Weighted mean depths of E. hamata and P. elegans

normalized relative to the bottom depth at Amundsen Gulf stations (see ‘Method’).

Multinet collections. Standard deviation bars also given. The blue area indicates the

Pacific Halocline (60-200 m) and the red area indicates the Atlantic Layer ˃ 200 m

(Geoffroy et al. 2011). 53

Figure 3.4 Length frequency (mean % ± 1 SD) distributions of Eukrohnia hamata and

Parasagitta elegans in the square-conical (S-C) net (1 m2 aperture, 200 µm mesh) and

Multinet (0.5 m2 aperture, 200 µm mesh) in the Amundsen Gulf. Note the different scales

for the two species. k is the number of collections. 54

Figure 3.5 Abundance (mean numbers m-2 + 1 SD) of newborn Eukrohnia hamata and

Parasagitta elegans (body lengths of 2-4 mm) in monthly Multinet deployments from

October to September. August 2007 was inserted between July and September 2008 to

provide a complete composite of the annual cycle. Number of collections shown above

bars. The solid horizontal bar above month labels indicates sampling in the Amundsen

Gulf. 55

Figure 3.6 Monthly length frequency distributions of Eukrohnia hamata. Frequencies of

newborns highlighted in orange. Visually identified length cohorts shown as normal

distributions (in red) with red dots indicating the mean length. Each of the five cohorts

is labelled with a capital E and a number from oldest (1) to youngest (5). Blue line is the

Page 14: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

xiv

total distribution obtained by summing the modelled distributions. Chi-square values for

the goodness-of-fit of the total distribution to the data are given. Sampling regions are

abbreviated in each panel: PC, Parry Channel; AG, Amundsen Gulf; BB, Baffin Bay. k

is the number of collections and n the number of length measurements included (see

‘Method’). 56

Figure 3.7 a) Monthly mean length (± 1 SD) of the five cohorts of Eukrohnia hamata starting

from a major birth month; b) Composite growth trajectories of the autumn brood (born

in October) and spring brood (born in April) assuming a 2-y lifespan; c) Growth-age

curves of the autumn and spring broods. Circles show mean values of each characteristic

(± 1 SD shown as ribbons). 58

Figure 3.8 Development of sexual features (a, oocyte number; b, ovary length; d, sperm load;

e, seminal vesicle width; f, seminal receptacle diameter) and oil vacuole area (c) with

length for the autumn and spring broods of Eukrohnia hamata. Circles show mean values

of each characteristic (± 1 SD shown as ribbons). Vertical line indicates average length

(15.4 mm) at one year of age. Maturity results for each brood were obtained from the

analyses of up to 283 individuals from each 1 mm length class. 60

Figure 3.9 Mean frequency (± 1 SD) of Eukrohnia hamata with oil in vacuole and/or

digestive tract by months. Number of water column collections shown inside bar (33 –

446 individuals per haul). 61

Figure 3.10 Abundances (ind. m-3) of Eukrohnia hamata age classes in discrete depth layers

of the Amundsen Gulf, characterized by their mid-points (black squares). Top panels:

autumn brood individuals. Bottom panels: spring brood individuals. Seabed shown in

brown, unsampled sections of the water column in gray. See ‘Method’ for further details.

62

Figure 3.11 Monthly length frequency distributions of Parasagitta elegans. Frequencies of

newborns highlighted in orange. Visually identified length cohorts shown as normal

distributions (in red) with red dots indicating the mean length. Each of the three cohorts

is labelled with a capital P and a number from oldest (1) to youngest (3). Blue line is the

total distribution obtained by summing the modelled distributions. Chi-square values for

the goodness-of-fit of the total distribution to the data are given. Sampling regions are

abbreviated in each panel: NF, Nachvak Fjord; PC, Parry Channel; AG, Amundsen Gulf;

BB, Baffin Bay. k is the number of collections and n the number of length measurements

included (see ‘Method’). 63

Figure 3.12 a) Monthly mean length (± 1 SD) of the three cohorts of Parasagitta elegans

starting from a major birth month; b) Composite growth trajectory of the single brood.64

Figure 3.13 Development of sexual features (a, oocyte number; b, ovary length; c, sperm

load; d, seminal vesicle width; e, seminal receptacle diameter) with length in Parasagitta

elegans. Circles show mean values of each characteristic (± 1 SD shown as ribbons).

Page 15: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

xv

Vertical line indicates average length (20.7 mm) at one year of age. Maturity results were

obtained from the analyses of up to 63 individuals from each 1 mm length class. 65

Figure 3.14 Abundances (ind. m-3) of Parasagitta elegans age classes in discrete depth layers

of the Amundsen Gulf, characterized by their mid-points (black squares). Seabed shown

in brown, unsampled sections of the water column in gray. See ‘Method’ for further

details. 66

Figure 4.1 Bathymetric maps of the Arctic Ocean, showing the positions and IDs of sampling

stations (black circles) in (a) Amundsen Gulf, (b) north-eastern Chukchi Sea and (c)

Baffin Bay (SIF = Scott Inlet Fjord; GF = Gibbs Fjord; SBB = southern Baffin Bay).

Details of sampling stations are shown in Appendices E-1 and E-2. 79

Figure 4.2 Vertical distributions of chlorophyll a biomass (mg m-3) in the upper 200 m of the

water column along the ship track from November 2007 to July 2008 in the Amundsen

Gulf. Black dots indicate sampling depths. Details of sampling stations are shown in

Appendix E-1. Chl a data were provided by Michel Gosselin (Université du Québec à

Rimouski). 84

Figure 4.3 Average number of prey per chaetognath gut (npc) in square-conical (S-C) net and

Multinet collections from November 2007 to August 2008 in the Amundsen Gulf.

Numbers of individuals shown above data points. k is the number of collections. Inset

bottom: photograph of a Parasagitta elegans specimen (12 mm) with a relatively large

Pseudocalanus spp. copepod in the gut (from November 2 2007). 86

Figure 4.4 Photographs of substantial amounts of green macroaggregates (˃ 500 µm) in the

guts of Eukrohnia hamata, but not in the guts of Parasagitta elegans. Specimens from

the Amundsen Gulf on May 31 2008 (20-40 m depth), 20-30 mm. 87

Figure 4.5 Scanning Electron Microscope photographs of some items in Eukrohnia hamata

guts (left box) and Parasagitta elegans guts (right box), in different months. Examples:

hooks of a chaetognath, and mandibles of copepods (arrow caps). Specimens 20-30 mm.

89

Figure 4.6 Photographs of live Eukrohnia hamata individuals consuming green

macroaggregates in-vitro in the Chukchi Sea. Specimens 20-30 mm. Bottom right: An

inverted microscope photograph of phytoplankton such as Ceratium spp. in one gut. 90

Figure 4.7 Fatty acid biomarkers in Eukrohnia hamata and Parasagitta elegans at different

stations in autumn 2014. Sampling locations are abbreviated: CS, Chukchi Sea; SIF,

Scott Inlet Fjord; GF, Gibbs Fjord; SBB, southern Baffin Bay. From top panel to bottom

panel: mean ratios of the carnivory biomarker 18:1 (n-9)/(n-7), mean proportions of the

Calanus biomarkers ΣC20:1+C22:1 MUFA, and mean ratios of the algal biomarker

16:1/16:0 (± 1 SD). No bars – no data available. 94

Page 16: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

xvi

Figure 4.8 Mean sample δ13C and δ15N values and corresponding trophic levels (± 1 SD) of

Eukrohnia hamata and Parasagitta elegans. Station IDs shown next to data points.

Results from Chukchi Sea stations inside red oval. Results from Baffin Bay region

stations inside blue oval (locations are abbreviated: SIF, Scott Inlet Fjord; GF, Gibbs

Fjord; SBB, southern Baffin Bay). 96

Page 17: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

xvii

Dedicated to the memory of Helen McMahon, my granny, who made it possible for me to

first pursue my interest in the Arctic regions.

23 August 1929 – 25 March 2015

Angelina Kraft

JP Aubé

Page 18: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

xviii

Acknowledgements

I am now nearing the end of my journey to gain a Ph. D. in Oceanography, and there are

several people who have supported me whom I would like to acknowledge in this thesis.

Firstly, I would like to thank my director of research, Dr. Louis Fortier for allowing me to

realize this project on the enigmatic chaetognaths. I would also like to thank Louis for giving

me opportunities to attend and present my research at several national and international

conferences, to voyage to the Arctic/sub-Arctic on three separate research cruises, and for

the many ways in which he helped me improve my experience in scientific writing. Thanks

to Drs. Jean-Éric Tremblay, Guillaume Massé and Stéphane Plourde for evaluating the ‘dépôt

initial’ of my thesis and thesis defense presentation, and providing useful suggestions for

improvements. Thanks also to Dr. Maurice Levasseur for participating in my Ph. D.

committee, and to Dr. Ladd Johnson for presiding over my defense presentation.

I express my gratitude to my colleagues Drs. Øystein Varpe (UNIS), Stig-Falk Petersen

(Akvaplan-niva), Dominique Robert, Eric Rehm, Guillaume Massé and Marcel Babin

(Université Laval), Roxane Barthélémy and Jean-Paul Casanova (Aix-Marseille Université),

and Mrs. Ariane Beauféray (Université Laval), all of whom helped to shape the direction of

this thesis. I would like to specifically acknowledge Dr. Øystein Varpe who has consistently

shown a keen interest in my research, and been happy to offer constructive ideas and

suggestions. Thanks to Drs. Catherine Lalande (Université Laval), and Tara Connelly

(Memorial University of Newfoundland), as well as anonymous journal referees for Polar

Biology and Journal of Plankton Research, for offering valuable comments on chapters.

I am grateful to the crew and scientific staff on expeditions with R/V Helmer Hanssen and

CCGS Amundsen for research support. Dr. Gérald Darnis’ assiduity onboard CCGS

Amundsen, and in the laboratory, was fundamental to the success of zooplankton sections of

the CFL project. Mr. Carl Ballantine collected samples from Svalbard in 2013. I greatly

appreciate the hard work of my student assistants; Mrs. Ariane Beauféray, Miss. Marianne

Caouette, Miss. Vicky St-Onge, Mr. Joël-Fortin Mongeau and Mr. Pierre-Olivier Sauvageau,

for helping me to analyse the vast number of chaetognaths included in this thesis. Sincere

thanks also to the following individuals and lab groups for analysing fatty acids and stable

Page 19: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

xix

isotopes: Dr. Guillaume Massé, Miss Caroline Guilmette, Mr. Jonathan Gagnon, the Institute

for Energy Technology and UNILAB in Norway. Thanks to Drs. Jean-Éric Tremblay and

Michel Gosselin, and Miss Marjolaine Blais for contributing environmental datasets.

I would like to thank my close friends Moritz Schmid, Kai Shapiro and Sophie Regueiro for

their moral support during the most challenging stages of my Ph. D. Many others, including

Cyril Aubry, Kevin Gonthier, Arnaud Pourchez, Noémie Friscourt, Max Geoffroy and other

members of the Fortier lab, have given me great memories from my time in Québec and

Canada. I am grateful to my parents (Ali and Dave), sister Ara, as well as the rest of my

family, who were ready to Skype at a moment’s notice and were always ready to offer

support, advice and encouragement. Thanks to the School for Science and Math at Vanderbilt

(Nashville, USA) for more recently giving me the opportunity to share my passion for the

polar regions with the next generation of scientists.

Bursaries from Takuvik and Québec-Océan allowed me to keep the focus on all my doctoral

activities and reach these final stages of my Ph. D. course.

Page 20: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

xx

Foreword

This doctoral thesis comprises a General Introduction (Chapter 1), three scientific articles

(Chapters 2, 3 and 4), and a number of General Conclusions relating to the objectives of the

thesis (Chapter 5). Chapter 2 is published in Polar Biology as part of a Special Issue on Polar

Night ecology. Chapters 3 and 4 are currently in preparation for publication:

Chapter 2

Grigor J.J., Marais A., Falk-Petersen S., Varpe Ø. (2015) Polar night ecology of a pelagic

predator, the chaetognath Parasagitta elegans. Polar Biology 38:87-98 (reproduced with

permission of the publisher). Minor formatting changes have been made to the material

published in Polar Biology, to ensure compatibility with the other chapters in this thesis (e.g.

species names are written out in full the first time they are used in every new paragraph, and

thereafter abbreviated to their shorter form. Citations follow the same format as in the other

chapters).

Chapter 3

Grigor J.J., Schmid M.S., Fortier L. Growth and reproduction of the chaetognaths

Eukrohnia hamata and Parasagitta elegans in the Canadian Arctic Ocean: capital breeding

versus income breeding. This has been revised and recently re-submitted to Journal of

Plankton Research.

Chapter 4

Grigor J.J., Barthélémy R.-M., Massé G., Casanova J.-P., Fortier L. Feeding strategies of

arctic chaetognaths: are they really “tigers of the plankton”? This will be submitted to Journal

of Plankton Research.

The aims and objectives of this thesis were designed by myself. All the analyses were

performed by myself (or by students under my supervision). To collect material for the thesis,

I participated in research cruises in winter 2012, autumn 2013 and autumn 2014. Chapters 2-

4 have benefited from the corrections and comments of my co-authors. Results of these three

Page 21: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

xxi

chapters have been presented at the following national and international scientific

conferences:

[9] Grigor J.J., Marais A.E., Schmid M.S., Varpe Ø., Fortier L. (2015) Ecology of arrow

worms in the Arctic: are they really “tigers of the zooplankton”?? APECS Online

Conference – New Perspectives in the Polar Sciences

[8] Grigor J.J., Varpe Ø., Marais A.E., Schmid M.S., Rehm E., Fortier L. (2014) Ecology

of arrow worms in the Arctic: are they really “tigers of the zooplankton”?? Arctic Change

Conference (Ottawa, Canada)

[7] Grigor J.J., Marais A.E., Schmid M.S., Rehm E., Fortier L. (2014) Arrow worm

ecology in the Canadian Arctic. Arctic Change Conference (Ottawa, Canada)

[6] Grigor J.J., Marais A.E., Schmid M.S., Rehm E., Fortier L. (2014) Arrow worm

ecology in the Canadian Arctic. Assemblée générale annuelle de Québec-Océan (Rivière-

du-Loup, Canada)

[5] Grigor J.J., Marais A.E., Schmid M.S., Fortier L. (2013) Seasonal ecologies of

chaetognaths (gelatinous zooplankton) in the Canadian Arctic. ArcticNet Annual Science

Meeting (Halifax, Canada) *Awarded best poster in its category: Marine Natural Science

[4] Grigor J.J., Søreide J.E., Varpe Ø., Fortier L. (2013) Seasonal ecology and life history

strategies of the chaetognath Parasagitta elegans in the Canadian and European Arctic:

The story so far. Assemblée générale annuelle de Québec-Océan (Rivière-du-Loup,

Canada)

[3] Grigor J.J., Søreide J.E., Varpe Ø. (2013) The annual routine of a predatory arctic

chaetognath in a highly seasonal environment. Arctic Frontiers: Geopolitics and marine

production in a changing Arctic (Tromsø, Norway)

Page 22: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

xxii

[2] Grigor J.J., Søreide J.E., Varpe Ø. (2012) The annual routine of a predatory

zooplankter in a highly-seasonal Arctic environment. Annual ArcticNet Science Meeting

(Vancouver, Canada)

[1] Grigor J.J., Søreide J.E., Varpe Ø. (2012) The annual routine of a predatory

zooplankter in a highly-seasonal Arctic environment. Québec Océan 10-yr conference: A

reality check on oceans’ health (Montreal, Canada)

During my Ph. D., I participated as a co-author in the writing of one scientific article that

used an underwater camera system, the Lightframe On-sight Keyspecies Investigation system

(LOKI), to capture excellent images of zooplankton and document their fine-scale vertical

distributions. The study also developed a model to automatically identify the species and

stages of imaged animals:

[1] Schmid M.S., Aubry C., Grigor J.J., Fortier L. (2016) The LOKI underwater imaging

system and an automatic identification model for the detection of zooplankton taxa in the

Arctic Ocean. Methods in Oceanography 15:129-160

Page 23: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

1

1. Chapter 1 – General introduction

1.1 The Arctic Ocean

The Arctic area is characterized by strong seasonal cycles in solar radiation and surface sea-

ice cover. The sun remains more than 12° below the horizon at the height of winter (the ‘polar

night’) and does not set at the height of summer (the ‘midnight sun’). The Arctic Ocean

covers an area of 14 million km2 (Huntington & Weller 2005) and is divided by a submarine

ridge of continental crust (the Lomonosov Ridge) into two major basins, the Eurasian and

Amerasian Basins (Figure 1.1). Vast continental shelves comprise 50 % of the Arctic Ocean

area (Figure 1.1), and these are amongst its most biologically productive areas (Sakshaug

2003, see ‘Marine ecosystems’ section).

Figure 1.1 Bathymetric map of the central Arctic Ocean and several marginal seas

(http://recherchespolaires.inist.fr/?L-ocean-Arctique-physiographie).

Page 24: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

2

1.1.1 Physical environment

Through dynamic and thermodynamic processes (movement, growth and melt), the extent of

surface sea-ice increases to its maximum in March and declines thereafter towards its

minimum in September (Stroeve et al. 2008). Sea-ice acts as a barrier between the underlying

ocean and the atmosphere affecting heat exchange and the transmission of light that algae in

the water require for primary production. Sea-ice is also a habitat for several specialized fish,

seabird and mammal species, and influences global climate. Ice formation in key locations is

important for the formation of cold and highly saline water. This briny water sinks, initiating

deep-water currents that move into other oceans as an integral part of the ‘thermohaline

circulation’ (Loeng et al. 2005). Surface water/ice follows two important surface currents: in

the western Arctic the clockwise-turning Beaufort Gyre; and in the eastern Arctic the

Transpolar Drift, which exports ice into the Atlantic (Stein & Macdonald 2004). Freshwater

contributions from ice melt, precipitation and large non-freezing rivers (e.g. Ob and Yenisei

in Eurasia, the McKenzie and Yukon in N. America), increase the buoyancy of surface waters

(i.e. the Polar Mixed Layer). Great volumes of warm, relatively saline Atlantic water, also

enter the Eurasian Basin through the deep (2600 m) Fram Strait (Figure 1.2) and the Barents

Sea. Lower volumes of cold, relatively fresh Pacific water, are supplied to the Amerasian

Basin through the 45 m-deep Bering Strait (Stein & Macdonald 2004, Figure 1.2).

Figure 1.2 Water masses in the Arctic Ocean. Reproduced from Stein & Macdonald (2004).

Page 25: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

3

In the Amerasian Basin, the Atlantic Water cools and becomes denser, plunging below the

Pacific water to depths ˃200 m, where it resides for ~30 years (Stein & Macdonald 2004,

Figure 1.2). The inflows of Pacific water in the Amerasian Basin and of Atlantic water in the

Eurasian basin create salinity gradients (haloclines) that protect sea-ice from warm waters

below (Winsor & Bjork 2000). Arctic deep water (cold and relatively saline), can occur below

900 m in the deep areas of the basins, where it has a residence time of 75-300 years (Stein &

Macdonald 2004, Figure 1.2).

1.1.2 Arctic marine ecosystems

Arctic marine ecosystems are considered less complex than most temperate and tropical

systems, with lower productivity and biodiversity in general (Loeng et al. 2005). In Arctic

marine ecosystems, extreme cycles in solar illumination and sea-ice cover lead to strong

seasonal cycles in primary production by autotrophic algae, with a considerable reduction in

winter (Ji et al. 2013). However, biological hotspots occur where a good nutrient supply to

surface waters supports high primary productivity. For instance, large blooms of

phytoplankton in some parts of the Chukchi Sea near Alaska support sizable communities of

invertebrates, both in the water column and on the seabed (Hopcroft et al. 2004, Grebmeier

et al. 2006, Nishino et al. 2016). Furthermore, polynyas (permanently open water areas

otherwise surrounded by sea-ice) are viewed as biological ‘oases’, with phytoplankton

blooms occurring in polynyas earlier in the season than in any surrounding areas (Ingram et

al. 2002, Hannah et al. 2009). This provides less-seasonally restricted access to food for

grazers and in turn their predators. Polynyas can be formed by the upwelling of warm

seawater which prevents ice from forming (i.e. sensible heat polynya), and/or by the influence

of wind in driving ice away (i.e. latent heat polynya; Hannah et al. 2009). The large North

Water Polynya in northern Baffin Bay (Canadian Arctic) is mainly a latent heat polynya in

winter and spring, however, sensible heat is also important for its growth in late spring

(Ingram et al. 2002).

In ice-covered seas, a bloom of under-ice algae is initiated during spring which is then

succeeded by a bloom of pelagic phytoplankton when the ice retreats in the spring/summer.

Page 26: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

4

Some locations can have another distinct phytoplankton bloom in autumn (Ardyna et al.

2014). Increases in the occurrence of autumn blooms across the Arctic may be a symptom of

global warming (see ‘Climate change and other challenges for Arctic marine life’ section).

On average, phytoplankton production is higher (12 to 50 g C m-2 yr-1) than ice algae

production (5 to 10 g C m-2 yr-1). However, ice algae production exceeds that of

phytoplankton in locations with thick multi-year ice cover (Legendre et al. 1992, Leu et al.

2011). Bloom phenology varies with latitude and timing of open water (Zenkevich 1963,

Falk-Petersen et al. 2009, Tremblay et al. 2012), and the magnitude of a phytoplankton bloom

during the ice-free period is determined by nitrogen availability (Tremblay & Gagnon 2009),

and also influenced by zooplankton grazing (e.g. Banse 2013).

Arctic marine ecosystems contain at least 5000 species of invertebrate animal taxa from 24

phyla (Hoberg et al. 2013 and references therein), 91 % of which are associated with the

seabed (i.e. ‘benthic’), 8 % with the water column (i.e. ‘pelagic’) and 1 % with the sea-ice

(i.e. ‘sympagic’). Copepods belonging to the genus Calanus (Figure 1.3) are known as ‘key

species’ in the Arctic seas. They convert carbohydrates and proteins taken up from the algae

into high-energy lipids (wax esters) that comprise the main energy source within the food

web (e.g. Falk-Petersen et al. 2009).

Figure 1.3 The planktonic food web showing

protozoan and metazoan consumers at arrow

caps. Adapted from Hopcroft et al. (2008).

Page 27: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

5

Amongst the vertebrates, up to 250 fish species reside in the Arctic area defined by the

Conservation of Arctic Flora and Fauna (an Arctic Council Working Group), for part or all

of the year (Hoberg et al. 2013). The polar cod Boreogadus saida is an endemic species which

can channel up to ~75 % of energy flow from zooplankton to vertebrate predators (Welch et

al. 1992). In addition, 35 species of mammals and 60 species of seabirds are observed in

Arctic waters, with the Baffin Bay little auk population considered to be the largest seabird

population in the world (Hoberg et al. 2013). Heterotrophic microbes and protists (Figure

1.3) remineralize dissolved organic carbon (˂0.45 µm) excreted from these consumers into

inorganic forms of carbon which can then be taken up again by protozoans (Hopcroft et al.

2008).

1.2 Zooplankton and their polar adaptations

Zooplankton are aquatic animals with body sizes ˂ 0.002 mm to ˃ 200 mm which have limited

ability to control their horizontal positions against water currents. As zooplankton have little

capacity to evade warming waters and are typically not harvested by humans, variations in

their distribution or phenology may strongly herald climate change effects (Richardson

2008). Zooplankters include herbivores, omnivores and carnivores, and are food for a wide

diversity of invertebrate and vertebrate predators. Zooplankton fecal pellets can be an

important source of carbon sequestration (e.g. Wassmann et al. 1999). However, fecal pellets

excreted in epi-pelagic waters are typically removed rapidly in surface waters. This is due to

degradation by bacteria, protozooplankton and copepods (Morata & Seuthe 2004, Turner

2015). Copepods, chaetognaths and other zooplankters are classified as ‘holoplankton’ taxa

(that drift in the sea throughout their entire life cycle). In contrast, a variety of fish, decapods

and other species are plankton only when they are larvae (‘meroplankton’), before then

continuing life as nekton or benthos (Stübner et al. 2016). Some zooplankton species are

capable of migrating considerable distances through the water column over day-night,

seasonal or ontogenetic cycles. These behaviors may be associated with finding food,

avoiding predation, or gaining an energy bonus from resting in colder waters, and they alter

the spatial distributions of food, materials and energy (Longhurst et al. 1984, Haney 1988).

Page 28: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

6

Arctic zooplankton biomass is typically dominated by Calanus copepods, followed by non-

copepod taxa such as amphipods and chaetognaths (Figure 1.4). Zooplankton abundance is

dominated by small cyclopoid and calanoid copepods including Oithona similis and

Pseudocalanus spp., amongst others (e.g. Søreide et al. 2003, Hopcroft et al. 2004, Darnis &

Fortier 2014). For stable co-existence, similar species must occupy slightly different

ecological niches, efficiently partition available food and share habitat resources (Gause

1934, Ross 1986).

Figure 1.4 Photographs of living holo-zooplankton captured in-situ by a zooplankton imager

(Lightframe On-sight Key Species Investigation System; Schmid et al. 2016), during

deployments in the Canadian Arctic (summer 2014).

Due to reduced growth rates in colder temperatures, zooplankton and other polar

invertebrates often have longer lives and they reach maturity at older ages than those at lower

latitudes (Hoberg et al. 2013). High-latitude zooplankton display a range of adaptations to a

seasonal food source associated with, for instance, the timing and extent of growth and

Page 29: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

7

reproduction, diapause, lipid content and feeding strategy (e.g. Conover 1988, Falk-Petersen

et al. 2009, Varpe 2012, Grigor et al. 2014).

For grazers, such as Calanus glacialis, timing their growth and reproduction to coincide with

the ice algae bloom can be of the utmost importance, as it would enable offspring to hatch

~3 weeks later during the phytoplankton bloom (Søreide et al. 2010, Leu et al. 2011, Daase

et al. 2013). The high-quality food provided during this latter bloom allows for critical early

growth of copepod offspring, without risk of starvation. Such a ‘match’ between the

development of primary production and copepod offspring benefits copepod populations and

the food web in general. However, in cases where offspring hatch before or after the peak of

the phytoplankton bloom, they miss the phytoplankton food needed for healthy growth

(Søreide et al. 2010, Leu et al. 2011). Such a ‘mismatch’ scenario has been shown to

drastically reduce Calanus biomass (Leu et al. 2011).

Storage lipids are important for Arctic organisms, in which they may be used to fuel one or

more of the following processes; basal metabolism, growth, maturation and reproduction

(Kattner et al. 2007). Wax esters (comprising one fatty acid and one fatty alcohol) are the

main long-term energy deposit in zooplankton, whereas triacylglycerols (comprising three

hydrocarbon chains on a glycerol backbone) provide energy for activities in the short term

(Lee et al. 2006). In winter, Calanus copepods and other grazers survive by entering diapause

in relatively deep waters during which time accumulated wax ester stores are metabolized

for energy. Capital breeders (e.g. C. hyperboreus) use their wax esters to fuel reproduction,

and can produce offspring earlier in the season giving them a longer time to develop their

energy reserves, required for surviving the next winter. Reproductive fitness is also improved

(Varpe et al. 2007). Income breeders need recent food to reproduce, and mixed capital-

income breeders use a combination of the two strategies (e.g. Sainmont et al. 2014a).

Carnivores are assumed to have a low capacity for long-term energy storage (Hagen 1999),

however, this may not be true for some species. Carnivorous and omnivorous zooplankters,

like amphipods (Figure 1.4) are also assumed to feed opportunistically year-round (Hagen

1999).

Page 30: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

8

1.3 Chaetognaths

Chaetognaths (Figure 1.5) are a phylum of semi-gelatinous (dry weight ˃5 % wet weight;

Larson 1986) marine mesozooplankton. They are represented worldwide by at least 127

pelagic and benthic species from 23 genera with body lengths 2-120 mm (e.g. Bone et al.

1991, Foster 2011). The name ‘chaetognath’ (Chaeto = bristle, gnath = jaw) relates to the

presence of hooks on the head (Figure 1.5). Chaetognaths are also known as ‘arrow worms’

due to their lean appearance and rapid motion. These coelomates have existed from the

Middle Cambrian and possibly before then (Walcott 1911). The global biomass of

chaetognaths has been estimated at 10-30 % of that of copepods (Bone et al. 1991). The

majority of species are residents of the upper 200 m of the water column (i.e. epi-pelagic;

Pierrot-Bults & Nair 1991).

Figure 1.5 Typical physiology of a chaetognath belonging to the family Sagittidae.

Reproduced from Margulis & Chapman (2010).

Chaetognaths occur in the diets of zooplankton such as amphipods (Gibbons et al. 1992),

jellyfish (e.g. Zavolokin et al. 2008) and other chaetognaths (e.g. Pearre 1982), and recently

have received attention as the preferred prey for the larvae of tropical fish and commercially

Page 31: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

9

important decapods (e.g. Sampey et al. 2007, Saunders et al. 2012). In the Arctic, fish such

as the polar cod (Rand et al. 2013), seabirds such as the little auk (Mehlum & Gabrielsen

1993) and baleen whales (Pomerleau et al. 2012), feed on chaetognaths. Suthers et al. (2009)

described chaetognaths as “the tigers of the plankton”. Animals in the phylum are generally

thought to be opportunistic, but strict carnivores (Marazzo et al. 1997). For further details,

see ‘Feeding strategies and trophic importance’ section.

1.3.1 Arctic species and distributions

In contrast to the high chaetognath diversity observed in other seas (e.g. De Souza et al.

2014), only three major species (Eukrohnia hamata, Parasagitta elegans and Pseudosagitta

maxima) are frequently reported in Arctic plankton surveys (e.g. Kramp 1939, Buchanan &

Sekerak 1982, Sameoto 1987, Figure 1.6). Another two species have been detected in Arctic

bathy-pelagic surveys; Heterokrohnia involucrum (Dawson 1968) and H. mirabilis (Kapp

1991). Kosobokova & Hopcroft (2009) reported that chaetognaths represented, on average,

about 13 % of the zooplankton biomass in the Canadian Arctic in summer 2005.

Figure 1.6 Photographs of the heads and the bodies of the major arctic chaetognaths. (a)

Eukrohnia hamata with characteristic marsupial sac and oil vacuole, (b) and (c) Parasagitta

elegans and Pseudosagitta maxima without marsupial sacs and oil vacuoles.

All images except P. maxima head from:

http://www.arcodiv.org/watercolumn/Chaetognaths.html (courtesy of Russ Hopcroft).

Page 32: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

10

Parasagitta elegans (previously Sagitta) is restricted to the North Atlantic, North Pacific and

Arctic Oceans, with its lower distribution limit at 41°N and a northern limit short of the

central Arctic basin (e.g. Bogorov 1946, Bieri 1959, Grainger 1959). It is generally an epi-

pelagic, neritic species throughout most of its range (e.g. Terazaki 2004). Terazaki (2004)

described P. elegans as the best-studied chaetognath species. In contrast to P. elegans,

Eukrohnia hamata and Pseudosagitta maxima are cosmopolitan species occurring in all

oceans, with predominantly meso- or bathy-pelagic affinities (e.g. Bigelow 1926, Bieri 1959,

Cheney 1976, Thuesen et al. 1993). All three species occasionally co-occur at epi-pelagic

depths in the Arctic (e.g. Alvarino 1964, Hopcroft et al. 2005, Bieri 1959). The life cycle and

feeding strategy of E. hamata is more thoroughly studied in the Southern Ocean (e.g.

Øresland 1990, Øresland 1995, Froneman & Pakhomov 1998, Froneman et al. 1998, Kruse

2009, Kruse et al. 2010), and few studies exist on the ecology of P. maxima in the Arctic (see

Sameoto 1987).

Some chaetognaths may migrate long distances vertically through the water column. Grigor

et al. (2014) reported seasonal vertical migrations (SVM) in a fjord population of Parasagitta

elegans in the European Arctic. In this Svalbard fjord, the youngest age class resided near

the surface during the summer phytoplankton bloom, whereas all three age classes mostly

remained below 60 m from September to April (seasonal migrations closely mirrored those

of Calanus prey). Daase et al. (2016) also reported diel (day-night cycle) vertical migrations

(DVM) in a population of Eukrohnia hamata in another Svalbard fjord, towards the end of

the midnight sun period. Migrations involved upward movements of E. hamata to surface

waters, apparently to feed at night, and downward movements to deeper waters, apparently

to avoid predators in well-lit waters during the day.

1.3.2 Lifespans

The life history characteristics of chaetognath species (e.g. life span, size at maturity, time

and number of spawning cycles per year) differ throughout their geographical ranges in

response to variations in temperature, food quantity and quality (Pearre 1991, Alvarino 1990,

Terazaki 2004). In temperate seas, both Eukrohnia hamata and Parasagitta elegans may

Page 33: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

11

have lifespans of less than a year (e.g. Russell 1932, Terazaki & Miller 1986, Sameoto 1971,

Zo 1973, King 1979, Terazaki & Miller 1986). In the Arctic, two-year lifespans have been

suggested for E. hamata (Sands 1980, Sameoto 1987, Pearre 1991); one to three-year

lifespans have been suggested for P. elegans (e.g. Dunbar 1940, McLaren 1961, Timofeev

1995, Welch et al. 1996, Grigor et al. 2014); and the larger Pseudosagitta maxima may live

for four to five years (Sameoto 1987).

1.3.3 Reproductive strategy

Chaetognaths are hermaphrodites, able to reproduce by both cross and self-fertilization. Male

gonads in the tail region (testes and seminal vesicles, Figure 1.5) produce and transmit sperm,

whilst female gonads in the trunk region (ovaries and seminal receptacles, Figure 1.5)

produce ova and receive sperm from a neighbor (Alvarino 1992). Although cross-fertilization

is more common due to sperm maturing earlier than oocytes, this strategy makes self-

breeding possible when access to mates is limited (Alvarino 1992). The eggs of chaetognaths

hatch a few days after oocytes are fertilized (e.g. Kotori 1975). Hatching takes place inside a

marsupial sac in Eukrohniidae (Figure 1.6a) and in open water in Sagittidae (Alvarino 1968,

Kotori 1975, Hagen 1985). Kuhl (1938) suggested that chaetognaths die immediately after

spawning. However, several other authors have disputed this finding, instead showing

iteroparity and continued spawning after laying their first batch of eggs (Conway & Williams

1986, Grigor et al. 2014). Whilst Eukrohnia hamata is reported to breed year-round in the

Southern Ocean (Øresland 1995), breeding of Parasagitta elegans in the Arctic mainly

coincides with peak abundance of zooplankton food (e.g. Kramp 1939, Dunbar 1940,

Grainger 1959, McLaren 1961, Dunbar 1962, Grigor et al. 2014).

1.3.4 Oil vacuoles in Eukrohnia spp.

The middle section of the gut in Eukrohnia spp. contains an elliptical oil vacuole (Figure

1.6a), also described in earlier reports as an ‘oil droplet’ (Øresland 1990, Froneman &

Pakhomov 1998, Froneman et al. 1998, Kruse et al. 2010, Giesecke & Gonzalez 2012, Pond

2012). Such an oil vacuole is not present in Sagittidae species (e.g. Figure 1.6b and Figure

1.6c). Oil vacuoles in Eukrohnia species may store wax esters. Wax esters seem to occur in

Page 34: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

12

much higher amounts in Eukrohnia hamata compared to Parasagitta elegans and

Pseudosagitta maxima (e.g. Lee & Hirota 1973, Falk-Petersen et al. 1987, Donnelly et al.

1994, Lee et al. 2006, Connelly et al. 2012, Connelly et al. 2016). Large wax ester supplies

in Eukrohnia chaetognaths could suggest a potential for metabolic reduction/capital breeding,

as occurs in overwintering zooplankton, though the centralized position of the oil vacuole in

the body also hints at a buoyancy control function (Pond 2012).

1.3.5 Feeding strategies and trophic importance

Zooplankton prey includes a wide range of ichthyoplankton and zooplankton, particularly

copepods (e.g. Terazaki 2004). Prey is sensed mechanically; chaetognath hairs detect

vibrations at distances of 1-3 mm (Feigenbaum & Reeve 1977, Goto & Yoshida 1981).

Specific prey types can be differentiated by their unique movements (Tonnesson & Tiselius

2005). The eyes of chaetognaths lack lenses capable of forming images but they do contain

a pigment that detects light direction (Pearre 1973). Therefore, their eyes may help them to

detect prey outside the maximum distance for hair detections (Goto & Yoshida 1981). Dorso-

ventral undulating movements are used to dart at prey, which is then seized by the hooks and

punctured by rows of small teeth (Figure 1.7). The locomotion appendages, as well as the

feeding appendages, differ between chaetognath species, for instance Eukrohnia hamata

lacks anterior teeth (Furnestin 1965), present in P. elegans (Terazaki 1993). Relatively large

items are consumed by expansions of the head (Bieri & Thuesen 1990). Chaetognaths may

paralyze prey by secreting a bacterial-produced neural poison (tetrodotoxin) from vestibular

ridges (Thuesen & Kogure 1989, Bieri & Thuesen 1990, Figure 1.7).

Figure 1.7 Photograph of a chaetognath’s head taken by Scanning Electron Microscopy

(left) with parts labelled (right). Reproduced from Bieri & Thuesen (1990).

Page 35: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

13

In the waters of Nova Scotia (Canada), Sameoto (1971, 1972) suggested that the abundant

Parasagitta elegans could remove considerable fractions of the copepod standing stock per

day (36 % to ˃100 % depending on season). Welch et al. (1992) suggested that Eukrohnia

hamata and Pseudosagitta maxima may ingest up to 51 % of the copepod biomass in

Lancaster Sound annually, but noted that their conclusions for these chaetognaths were “little

better than guesses” (Welch et al. 1992). The above estimates of predation rates or proportion

of prey standing stock removed, were based on observations of gut contents in dead

chaetognaths, and specifically, methodology which treated visible lipid droplets in their

bodies as evidence of recent prey digestion (e.g. Sameoto 1971, Sameoto 1972, Sameoto

1987). This approach was also adopted for E. hamata in the Southern Ocean (Froneman &

Pakhomov 1998). Including lipids in predation rate calculations may be suitable for species

without considerable capacity for lipid storage. For Eukrohnia species it seems there is a risk

of confusing diagnostic lipid stores with oil droplets from copepods, which is likely to have

occurred in at least the Sameoto (1987) study (Øresland 1990). Several other polar studies

on E. hamata, as well as P. elegans and P. maxima, have reported much lower predation rates

when lipid droplets were excluded from the estimates (e.g. Sameoto 1987, Øresland 1995,

Froneman & Pakhomov 1998, Brodeur & Terazaki 1999, Bollen 2011, Giesecke & Gonzalez

2012).

Whilst chaetoganths are generally recognized as predators, several studies have also observed

algae (e.g. Alvarino 1965, Boltovskoy 1981, Alvarez-Cadena 1993, Kruse et al. 2010) or

high amounts of algae biomarkers (e.g. Philp 2007, Connelly et al. 2014) in chaetognath guts,

possibly indicating that these opportunistic feeders also ingest algae in seawater or graze

them from a substrate. Chaetognaths belonging to the genera Archeterokrohnia and

Heterokrohnia reside in the nepheloid layer, where they are reported to feed on organic

matter and bacteria from sediments (Casanova 1986). In an interesting recent study,

Casanova et al. (2012) suggested facultative osmotrophy (ingestion of dissolved organic

matter in ambient seawater) as the main mode of nutrition in chaetognaths.

Page 36: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

14

1.3.6 Fatty acids and stable isotopes

Fatty acid profiles and stable isotope signals can also reveal information about consumer

diets, trophic positions, as well as the main primary producers in the food web. Chaetognaths

are easily stressed and damaged in plankton sampling nets. Potential food loss from damaged

guts or stress-induced regurgitation and defecation, as well as cod-end feeding, can all bias

gut content analyses and predation rate estimates (see Baier & Purcell 1997 and references

therein). In contrast to gut contents, fatty acid and stable isotope signals can persist for weeks

or months (Dalsgaard et al. 2003, Arim & Naya 2003). Fatty acid analysis is based on the

premises that (a) fatty acids are produced by specific organisms, and (b) the fatty acids

undergo limited breakdown and transformation in consumers (Dalsgaard et al. 2003, Arim &

Naya 2003). Nitrogen isotopes are used to assess trophic position, based on the premise that

the 15N/14N ratio undergoes a constant level of change between trophic levels (Minagawa &

Wada 1984, Michener & Schell 1994). Previous studies have used this technique to infer that

Eukrohnia hamata may feed more omnivorously than Parasagitta elegans in the Arctic

(Søreide et al. 2006, Hop et al. 2006). Carbon isotopes can indicate the carbon source in a

food web since the 13C/12C ratio differs little between a consumer and its food (Hobson &

Welch 1992).

1.4 Climate change and other challenges for Arctic marine life

Over millions of years, Arctic marine ecosystems have been shaped by profound

environmental changes that include transitions between glacial periods and interglacial

periods, in addition to large variations in sea level (Darby et al. 2006). However, the 2-3 ºC

rise in mean annual air temperatures since 1950 (Chapman & Walsh 2003) has had an

anthropogenic source (greenhouse gas emissions, and CO2 in particular). Sea surface

temperatures are increasing 2-3 times as fast as in the temperate oceans (known as ‘Arctic

amplification’), as elevated influx of warm Atlantic waters causes major heat advection

(Spielhagen et al. 2011). Overland et al. (2014) predicted further warming by up to 13 °C

prior to the autumn of 2100 if greenhouse gas emissions continue unabated. Consistent with

these changes, conditions in the Arctic Ocean are becoming more similar to those in

temperate oceans (referred to as ‘Atlantification’, see Wassmann et al. 2006).

Page 37: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

15

Arctic seas are acidifying twice as fast as temperate seas due to enhanced CO2 solubility in

colder waters (Bates et al. 2011). This shift could directly affect zooplankters such as the

pteropod mollusc Limacina helicina, a prominent component of Arctic zooplankton

communities (Gannefors et al. 2005). L. helicina precipitates CaCO3 to build a shell required

for its survival. However, experiments suggest that the lower pH values expected to occur by

the end of the century will be associated with a drastic 28 % decrease in shell calcification

rate and the death of these animals (Comeau et al. 2009).

Previous studies have reported decreases in both the annual mean extent of Arctic sea-ice by

3.5-4.1 % per decade (1979-2012; IPCC 2014), and mean sea-ice thickness at the end of the

melt season by 1.6 m (1958-2000; Kwok et al. 2009) as well as a 5-day increase in the length

of the summer melt-season (1979-1996; Smith 1998). Further sea-ice loss would be

disastrous for species that depend on sea-ice to live, hunt and breed, such as some

invertebrates and vertebrates including the polar bear. Ice algae would also lose their

substrate (Loeng et al. 2005). Despite this, some phytoplankton taxa may benefit because a

major barrier to sunlight access would be removed. Whilst oceanic primary production has

recently decreased at lower latitudes, Pabi et al. (2008) reported an overall ~30 % Arctic

increase in oceanic primary production between 1998 and 2006. Ardyna et al. (2014) also

noted that phytoplankton blooms in autumn are becoming more common throughout the

Arctic, as reduced ice cover allows winds to stir the water, breaking down any existing water

column stratification, and making nutrients previously at depth available to phytoplankton

near the surface. Another consequence of the recent sea-ice melt combined with an increased

river discharge (Peterson et al. 2002) has been an increase in water column stratification

(McLaughlin & Carmack 2010). This has created a scenario in which nutrients in shallow

water can become more quickly depleted by its algal residents. Li et al. (2009) have already

reported a corresponding shift from larger cells (i.e. diatoms) to smaller algal cells (i.e. pico-

phytoplankton) which can extract available nutrients more efficiently. Such changes in the

composition or biomass of primary producers could result in unpredictable ‘bottom-up’

effects that ripple through planktonic grazers to higher trophic levels. In the future, possible

loss of ice algae blooms and changes in the timing of phytoplankton blooms could cause

Page 38: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

16

more ‘mismatch’ scenarios in the development of primary producers and grazers, with

consequences for entire food webs (Søreide et al. 2010, Leu et al. 2011).

Changes in zooplankton abundances may also affect the survival of consumers which

selectively take larger lipid-rich taxa. Falk-Petersen et al. (2007) suggested that warming in

the Barents Sea (+2-3 ºC) could result in the replacement of two large lipid-rich species of

copepods (Calanus glacialis and C. hyperboreus) by a smaller, lipid-poor North Atlantic

congener species (C. finmarchicus). This may adversely affect specialist feeding seabirds

whilst benefiting herring and baleen whales, herring predators. Global warming is also

associated with northern movements of previously southern fish and mammals. There are

recent reports of northern movements of NE Atlantic Cod in the Barents Sea (e.g. Kjesbu et

al. 2014) and of killer whales in the Canadian Arctic (e.g. Higdon et al. 2012). As particularly

voracious predators on seals and fish, killer whales could exert top-down effects on lower

trophic levels, with unpredictable effects on zooplankton (Ferguson et al. 2010).

Arctic seas provide a wide range of ecosystem services for sustenance, recreation and the

well-being of people living within and outside the Arctic (e.g. petroleum products, fish,

animal furs, tourist destinations). Seabed beyond the Arctic Circle is thought to contain 30 %

and 13 % of the world’s undiscovered oil and gas respectively, with the majority of

undiscovered oil located in Arctic Alaska (Bird et al. 2008). With ice-free waters expected

to occur as early as September 2045 (Laliberté et al. 2016), these supplies will become

increasingly accessible and may be exploited as global population grows. Potential oil spills

are a concern because Arctic ecosystems are frail and Arctic seas present significant

containment and recovery challenges. The ecological impacts of previous large spills in sub-

Arctic waters have been considerable (e.g. Exxon Valdez 1989 in the Gulf of Alaska), with

spilt oil persisting in marine habitats for decades after spillages (e.g. Peterson 2001).

Shipping has also increased: with a change in the number of recorded vessels travelling

through the Northern Sea Route from 4 in 2010 to 71 in 2013 (Linstad et al. 2016). This also

exposes ecosystems to increased pollution, and so careful regulation placed on activities will

be critical to maintaining Arctic biodiversity. Worldwide increases in jellyfish populations

in recent years have been a concern for human health and marine operations such as fishing,

Page 39: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

17

drilling etc. These observations may be related to increasing temperatures, low populations

of forage fish (competitors for food) and the presence of humans, amongst other factors (e.g.

Brotz et al. 2012, Purcell 2012). Brotz et al. (2012) concluded with high certainty that

jellyfish numbers will increase in the Southern Ocean but it is unclear how they will fare in

the Arctic where less restricted access offers new prospects for human activity. Similarly,

few predictions have been made into how polar chaetognaths might respond to climate

change and other environmental pressures.

1.5 Study areas

This thesis includes depth-integrated and depth-stratified sampling of zooplankton in several

different regions of the Arctic Ocean. The major sampling regions were the Svalbard

archipelago in the European Arctic (Chapter 2), and Amundsen Gulf, located in the south-

eastern Beaufort Sea of the Canadian Arctic (Chapters 3 and 4). Additional collections from

other regions of the Canadian Arctic (the Labrador Sea, Baffin Bay and the Canadian

archipelago) were included in the analyses for Chapters 3 and 4. Finally, collections from the

Alaskan part of the Chukchi Sea were included in Chapter 4.

The Svalbard archipelago: The first of our main sampling regions; the Svalbard archipelago

in Arctic Norway, is a meeting point for Arctic water and Atlantic water (Strömberg 1989).

We studied chaetognaths in several fjords on the west and north coasts of Svalbard.

Parasagitta elegans is the dominant chaetognath species in the Barents Sea (Falkenhaug

1991) and along the west coast of Svalbard (Grigor et al. 2014, Hirche et al. 2015), but

Eukrohnia hamata dominates to the north in regions with pack ice and reduced influence of

Atlantic water (Daase & Eiane 2007, Daase et al. 2016). Several Arctic fjords including those

in Svalbard and Baffin Bay are particularly useful for zooplankton studies due to having

‘sills’, seabed intrusions that reduce advection of water masses, and associated plankton

communities (e.g. Syvitski et al. 1990, Bell & Josenhans 1997, Arnkvaern et al. 2005, Grigor

et al. 2014). Svalbard fjords in recent years have been the focus of novel polar night research

cruises led by Norwegian universities to investigate pelagic biology and winter ecology

(Lønne et al. 2014).

Page 40: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

18

The Amundsen Gulf: The second of our main sampling regions; the Amundsen Gulf in the

southeastern Beaufort Sea, was the site of a unique overwintering sampling campaign in

which an icebreaker was used to sample zooplankton and other ecosystem components year-

round (Figure 1.8). This formed a key part of the Circumpolar Flaw Lead (CFL) System

Study of 2007-2008. Several studies on copepod ecology have already been published from

this time (e.g. Forest et al. 2010, Wold et al. 2011a, Darnis & Fortier 2014).

Figure 1.8 Diagram illustrating the wide variety of sampling activities conducted during the

Circumpolar Flaw Lead System Study (2007-2008) from the icebreaker CCGS Amundsen

and at an ice camp, in order to study Arctic systems. Reproduced from Barber et al. (2010).

1.6 Aims and objectives

The central aim of this doctoral thesis is to improve our understanding of the ecology of two

major arctic chaetognath species; Eukrohnia hamata and Parasagitta elegans (Pseudosagitta

maxima was generally excluded from the thesis because of its rarity or absence at the

sampling locations). Specifically, the thesis aims at improving our understanding of these

chaetognaths’ feeding strategies and life cycles in the European Arctic, Canadian Arctic and

Alaskan Arctic. To that end, the thesis comprises three main chapters (2-4), presented as

scientific articles. The seasonal adaptations of arctic chaetognaths are interesting given the

Page 41: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

19

high seasonality in the abundance of known prey (copepods) in epi-pelagic waters, which

could require these predators to carry out seasonal vertical migrations (SVM), or be able to

quickly switch their food sources. Chapters 2-4 each comprise winter sampling, addressing

the knowledge gap of biological processes during the polar night (Lønne et al. 2014). This

will allow us to gain a more holistic understanding of how chaetognaths in the Arctic could

react to climate change and other environmental pressures. The inclusion of locations with

different environmental conditions at the same time of the year allowed me to investigate

possible differences in life cycles and feeding strategies between populations.

The objective of Chapter 2 was to shed light on the polar night ecology of Parasagitta elegans

in fjords of the Svalbard archipelago, where P. elegans is previously reported to be an

abundant chaetognath species. Feeding ecology and reproductive dynamics were examined

in four different fjords during January 2012 and January 2013. Gut content analyses were

used to assess diets, feeding rate and potential impacts of P. elegans on copepod populations.

To gain further insights on the trophic positions of chaetognaths, stable isotope and fatty acid

trophic marker (FATM) analyses were carried out.

The objective of Chapter 3 was to describe the life histories of Eukrohnia hamata and

Parasagitta elegans in the Canadian Arctic, mainly in the Amundsen Gulf, over a full annual

cycle (August 2007 to August 2008). Growth and reproductive dynamics were inferred from

body length, and the condition of oil vacuoles and maturity features. I tested the hypothesis

that E. hamata is a capital breeder that spawns during most of the year, whilst P. elegans is

an income breeder, reproducing exclusively during periods of high prey availability.

The objective of Chapter 4 was to understand how food resources are utilized year-round by

co-existing Eukrohnia hamata and Parasagitta elegans in the Canadian and Alaskan Arctic.

In the Amundsen Gulf, gut contents were examined to reveal diets from November 2007 to

August 2008. FATM and stable isotope signatures of E. hamata and P. elegans in two distant

regions (the Alaskan portion of the Chukchi versus Sea and Baffin Bay) were compared,

based on collections made in autumn 2014.

Page 42: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

20

2. Chapter 2 – Polar night ecology of a pelagic predator, the

chaetognath Parasagitta elegans

2.1 Résumé

Les routines annuelles et l'écologie saisonnière des espèces de zooplancton herbivores sont

relativement bien connues en raison de leur couplage étroit avec leur source de nourriture à

variation saisonnière, la production primaire. Quant aux niveaux trophiques planctoniques

supérieurs, ces interactions saisonnières sont moins bien comprises. Le présent chapitre

traitera de l’alimentation des chaetognathes dans les écosystèmes des fjords du haut-Arctique

au milieu de l’hiver. Les chaetognathes sont des prédateurs planctonophages très présents

dans les mers de haute latitude. Nous avons étudié l’espèce commune, le Parasagitta elegans,

autour de l'archipel du Svalbard (78-81ºN) au cours des hivers 2012 et 2013. Nos échantillons

étaient constitués de spécimens (longueur de corps : 9-55 mm), échantillonnés dans trois

fjords, dont le contenu du tube digestif (n = 903), la présence d’isotopes stables, la

composition en acides gras et l'état de maturité (n = 352) ont été étudiés. Environ un quart

des spécimens présentaient du contenu dans leur tube digestif, principalement des

gouttelettes de lipide et de débris chitineux, alors que seulement 4 % d’entre eux contenaient

des proies identifiables, principalement des copépodes Calanus spp. et Metridia longa. La

teneur élevée en δ15N du P. elegans et son niveau trophique moyen de 2.9 a confirmé son

profil de carnivore. Dans la même perspective, le profil d'acides gras [en particulier ses

niveaux élevés de 20:1 (n-9) et 22:1(n-11)] a confirmé la consommation de Calanus. Des

observations de gonades non-développées dans la plupart des P. elegans de grande taille,

ainsi que l'absence de petits spécimens ˂10 mm, a suggéré que la reproduction n'a pas

commencé si tôt dans l'année. Le taux d'alimentation moyenne du P. elegans selon les fjords

et les années était de 0,12 proie/ind.-1/jour-1, ce qui est faible par rapport aux mesures du taux

d'alimentation printanier et estival dans des environnements de haute latitude. Nos résultats

suggèrent une activité d'alimentation réduite pendant l'hiver. Par ailleurs, la prédation du P.

elegans semble avoir peu d'effets sur la mortalité des copépodes.

Page 43: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

21

2.2 Abstract

The annual routines and seasonal ecology of herbivorous zooplankton species are relatively

well known due to their tight coupling with their pulsed food source, the primary production.

For higher trophic levels of plankton, these seasonal interactions are less well understood.

Here, we study the mid-winter feeding of chaetognaths in high-Arctic fjord ecosystems.

Chaetognaths are planktivorous predators which comprise high biomass in high-latitude seas.

We investigated the common species Parasagitta elegans around the Svalbard archipelago

(78-81ºN) during the winters of 2012 and 2013. Our samples consisted of individuals (body

lengths 9-55 mm) from three fjords, which were examined for gut contents (n = 903), stable

isotopes, fatty acid composition, and maturity status (n = 352). About a quarter of the

individuals contained gut contents, mainly lipid droplets and chitinous debris, whilst only 4

% contained identifiable prey, chiefly the copepods Calanus spp. and Metridia longa. The

δ15N content of P. elegans, and its average trophic level of 2.9, confirmed its carnivorous

position and its fatty acid profile [in particular its high levels of 20:1 (n-9) and 22:1 (n-11)]

confirmed predation on Calanus. Observations of undeveloped gonads in many of the larger

P. elegans, and the absence of small individuals ˂10 mm, suggested that reproduction had

not started this early in the year. Its average feeding rate across fjords and years was 0.12

prey ind.-1 day-1, which is low compared to estimates of spring and summer feeding in high-

latitude environments. Our findings suggest reduced feeding activity during winter and that

predation by P. elegans had little impact on the mortality of copepods.

Page 44: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

22

2.3 Introduction

The Arctic is a highly seasonal environment, with conditions in the marine pelagic varying

greatly between summer and winter. The extreme cycles in solar illumination and sea-ice

cover create strong seasonality in primary production (Ji et al. 2013), which affects food

availability for organisms at higher trophic levels (Varpe 2012). Food availability for

herbivores such as Calanus copepods is severely reduced during the polar night. Many

grazers survive winter food shortages by resting at depth and sustaining themselves on

storage lipids accumulated when food was last available (e.g. Conover 1988). Although

pelagic omnivores and carnivores may experience a less-pulsed food source than for the

herbivores, they may also display adaptations to a seasonal food source (e.g. Newbury 1971,

Choe et al. 2003, Kraft et al. 2013). However, for many predators we lack knowledge of their

feeding opportunities and activities during winter.

The chaetognath Parasagitta elegans (Verrill 1873) is highly represented in Arctic

mesozooplankton communities, numerically and in biomass terms (e.g. Søreide et al. 2003,

Hopcroft et al. 2005). This typically neritic species (Bieri 1959) can exert high predation

pressure on Arctic copepods and compete with larval fish (e.g. Sameoto 1987). Some authors

have proposed that chaetognaths are not particularly sensitive to seasonality (e.g. Hagen

1999), suggesting that their varied diets (reviewed by Terazaki 2004) and non-visual food

search should allow them to find adequate nutrition throughout the year. Lipid-rich copepods,

often preferred prey, diapause in large densities and in an unalert state during winter, so they

may be easy prey for some carnivores (Darnis et al. 2012). However, some chaetognath

species at high latitudes are found to contain considerable amounts of lipids, including wax

esters, which may be used for storage (e.g. Lee 1974, Kruse et al. 2010). Furthermore, winter

growth rates of P. elegans in the Arctic is lower than those in spring and summer (Grigor et

al. 2014), and these observations could indicate reduced winter feeding and resting strategies

in this chaetognath.

We report here on the winter feeding ecology of Parasagitta elegans in the European high

Arctic. We used three different methods to assess its diet and trophic level: gut content

Page 45: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

23

analysis, which records recent feeding, as well as stable isotope and fatty acid trophic marker

(FATM) analyses, which provide additional information on its feeding history, over time

frames of weeks to months (Graeve et al. 2005). We calculated feeding rates and possible

impact on copepod populations. Although we did not expect mid-winter reproduction (e.g.

Kramp 1939, Grigor et al. 2014), we also examined their level of maturity, allowing us to

infer how close P. elegans is to reproducing, which links to how actively individuals are

expected to feed.

2.4 Method

2.4.1 Study area

Chaetognaths were collected from waters around Svalbard (78-81ºN) during the winters of

2012 and 2013, during the ARCTOS ‘Polar Night Cruises’ (8 to 21 January 2012 and 9 to 18

January 2013).

The Svalbard archipelago is a meeting place for warm Atlantic water sourced from more

southern regions and colder Arctic water from the north. The northward moving West

Spitsbergen Current deposits Atlantic water into fjords on the west coast, with one branch

finally turning north-east at the top of Spitsbergen and affecting conditions in the Arctic

Ocean (Saloranta & Haugen 2001). We sampled three fjords in 2012: Rijpfjorden (station

R3), Isfjorden proper, and Adventfjorden, a tributary fjord of Isfjorden (station ISA, see

Figure 2.1). The latter two locations are hereafter referred to as ‘Isfjorden’. Another fjord,

Kongsfjorden, was sampled in 2013, and Rijpfjorden was re-sampled. See Appendix A for

full details on the sampling activities. Isfjorden (78ºN, 14ºE) and Kongsfjorden (79ºN,

11.5ºE) are influenced by Atlantic water masses and the outer parts are ice free for much of

the year (Svendsen et al. 2002, Nilsen et al. 2008), causing important seasonal changes in

biodiversity and animal populations in these fjords. Rijpfjorden, in contrast (80ºN, 22ºE), is

influenced by Arctic water, typically remaining ice-covered from December/February to July

(Wallace et al. 2010).

Page 46: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

24

Figure 2.1 Map showing the locations of the stations sampled for chaetognaths in January

2012 and 2013.

2.4.2 Physical and biological environment

At all stations, vertical profiles of salinity, temperature, density, and fluorescence were

obtained using a CTD (SBE 9) and processed following standard Sea Bird Electronics (SBE)

procedures. In 2012, a steep thermocline was observed in Rijpfjorden between ~70 and 100

m, over which temperature rose to 2 ºC and then fell again to 0.5 ºC at ~200 m (Appendix

B). At the mouth of Isfjorden, the thermocline was much weaker, with temperature increasing

gradually to 2 ºC at ~200 m. At Isfjorden (station ISA), temperature varied little with depth

(Appendix B). At all stations, salinity varied with depth in similar ways to temperature. A

strong halocline occurred in Rijpfjorden only, at the same depths as the thermocline

(Appendix B). Fluorescence was very low (˂0.6 g l-1) at all stations in January 2012.

2.4.3 Zooplankton sampling

Horizontal trawls of the large Methot Isaac Kidd gear (MIK, 3.14-m2 opening, 1.5-mm mesh)

were used to obtain large numbers of Parasagitta elegans for the various dietary analyses

(Appendix A). MIK sampling depths in Rijpfjorden and Isfjorden (in 2012) captured all water

Page 47: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

25

masses in the fjords (melt water, Arctic water/mixed Fram Strait water and Arctic deep

water). In contrast, most MIK sampling for chaetognaths in 2013 (Kongsfjorden and

Rijpfjorden) was carried out in the upper 20 m (Appendix A). This sampling decision was

taken to reduce haul time, as chaetognaths can easily become damaged and stressed during

net sampling, which poses several problems for gut content analyses (Baier & Purcell 1997).

In Rijpfjorden in 2012, vertical hauls of the smaller Multi-Plankton Sampler gear (MPS,

0.25-m2 opening, 0.2 mm mesh) were also performed to collect zooplankton community data

(see also Daase et al. 2014) and to source smaller Parasagitta elegans for gut content analyses

(sampled strata: 260-200 m, 200-100 m, 100-50 m, 50-20 m, 20-0 m; see Appendix A for

details). The MPS captured higher densities of smaller P. elegans (10-19 mm) in the upper

20 m than the MIK (0.5 ± 0.4 compared to 0.1 ± 0.1 ind. m-3), but fewer individuals ≥20 mm

(1.8 ± 1.7 compared to 5.9 ± 8.2 ind. m-3). Individuals ˂10 mm were not captured by either

gear in Rijpfjorden or Isfjorden in 2012. Grigor et al. (2014) similarly showed that the MPS

captured the smaller size fraction of P. elegans more efficiently than a larger ‘WP3’ net with

a 1-m2 opening. MPS samples collected from Isfjorden ISA on 27 January 2012 were also

analysed for abundance data.

2.4.4 Sample processing

Upon retrieval of each MIK haul, the cod end was immediately transferred to a marked bucket

and diluted up to the 10 l mark. A 0.6 l sub-sample was taken for community analysis and

fixed in 4 % buffered formalin-seawater solution. Webster et al. (2013) show presence-

absence data for zooplankton in Rijpfjorden from the community samples from 2012. From

a second 0.6 l subsample, 100 Parasagitta elegans were randomly picked out for gut content

analyses (Appendix A) and preserved in 4 % buffered formalin-seawater solution. On the

2012 cruise, 75-150 P. elegans were also picked out of this second sub-sample for stable

isotope analyses (n individuals = 675, Appendix A), and 30 individuals were randomly

picked out for fatty acid trophic marker (FATM) analyses (n individuals = 290, Appendix

A). These samples were frozen at -80 ºC until the end of the cruise and then stored at -80 ºC

for further processing.

Page 48: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

26

2.4.5 Gut content analyses

A total of 202 individuals from 2012 (body lengths 10-55 mm) and 701 from 2013 (body

lengths 9-42 mm) were dissected and analysed for gut contents. In 2012, a minimum of 10

individuals were dissected from each MIK haul. In 2013, we examined the guts of all 100

individuals preserved from each MIK haul; 3-10 smaller individuals were sourced from each

MPS haul (Appendix A).

Parasagitta elegans individuals were measured to the nearest millimeter (excluding the

caudal fin) and stained with a solution of lignin-red, used to stain tissues of crustacean prey

in their guts (Falkenhaug 1991), and borax carmine, capable of highlighting the gonads of

prey (Pierce 1941), diluted 50 times in water. Head widths were measured to the nearest 0.01

millimeter under the binocular microscope. All visible gut contents were described. Detected

prey was identified to the lowest possible taxonomic level and photographed using a digital

camera connected to a dissecting microscope (Leica DFC 320).

2.4.6 Food-containing ratio and feeding rate

The food-containing ratio (FCR) is the proportion of individuals in a population with food in

their gut. Some only include chaetognaths containing identifiable prey in the FCR calculation

(FCRmin, e.g. Falkenhaug 1991, Kruse et al. 2010). However, this approach may not account

for loss of prey items from guts during net towing, due to damage or stress-induced

regurgitation (Baier & Purcell 1997, see ‘Discussion’). Other studies also included

individuals with other signs of recent prey digestion (gut lipids and debris, e.g. Sameoto 1987,

Froneman & Pakhomov 1998), which may give an upper estimate of the proportion feeding

(FCRmax). FCRmax measurements may not be appropriate for some chaetognath species, such

as those that contain oil vacuoles in the centre of their bodies. This includes another Arctic

species, Eukrohnia hamata (Möbius 1875), in which oil vacuoles may be for storage or

buoyancy (Øresland 1990, Pond 2012, pers. obs.). However, these oil vacuoles do not occur

in Parasagitta elegans. We calculated both FCRmax and FCRmin for P. elegans from each

haul. We also calculated feeding rates (FR: no. of prey items consumed ind.-1 day-1) according

Page 49: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

27

to Eq. (2.1). Only individuals containing identifiable prey were included in the FR

calculation.

FR =𝑛prey x 24

tdig (2.1)

where nprey = mean no. of identifiable prey per chaetognath, and tdig = digestion time in hours.

Feigenbaum (1982) suggested a tdig of 10.2 h for Parasagitta elegans from a laboratory study

on starved specimens from Vineyard Sound, maintained at 0 ºC. As the water temperature in

all our study fjords was close to 0 ºC (Appendix B), we used Feigenbaum’s tdig estimate.

Proportions of individuals per haul containing different gut types were analysed using

Kruskal–Wallis tests.

2.4.7 Stable isotope analyses

As the ratios of 13C/12C and 15N/14N in organisms at different trophic levels tend to differ

considerably, stable isotope analyses can be used to examine marine food web structure and

the transfer of energy between trophic levels (McConnaughey & McRoy 1979). Stable

isotope analyses were performed on our 2012 samples at the Institute for Energy Technology

in Kjeller, Norway, using similar methods to Søreide et al. (2006). Three replicate samples

each containing 25 Parasagitta elegans were analysed from seven of the MIK hauls, whilst

six replicate samples of 25 inds. were analysed from an eighth MIK haul at 225 m in

Rijpfjorden (Appendix A). All P. elegans had body lengths within the range of 10-50 mm.

The whole body of each chaetognath was used, except lipids and non-dietary carbon (i.e.

carbonates), which were removed before analyses by Soxhlet extraction with CH2Cl2. Lipids

have notably lower levels of 13C than proteins and carbohydrates (van Dongen et al. 2002),

and their removal reduces a main source of measurement variability between individuals

(Hobson & Welch 1992). C/N ratios were expressed as the deviation from standards in ppt

(‰) according to Eq. (2.2) (Søreide et al. 2006).

Page 50: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

28

δX=[(𝑅sample/𝑅standard) − 1]×1000 (2.2)

where X = 13C or 15N and R = the corresponding ratios 13C/12C or 15N/14N. International

standards, Pee Dee Belemnite for δ13C (PDB: USGS 24), and atmospheric air for δ15N

(IAEA-N-1 and 2) were used to determine R. Carbon and nitrogen composition were

expressed as percentages of animal dry weight.

2.4.8 Determination of trophic level

The trophic level (TL) of Parasagitta elegans in each fjord was calculated as the difference

between its average δ15N content and that of the winter food web baseline, assuming a

constant fractionation between trophic levels [Eq. (2.3), Søreide et al. 2006].

TL =∝ +δ15N 𝑃.𝑒𝑙𝑒𝑔𝑎𝑛𝑠− δ15N 𝐶.𝑔𝑙𝑎𝑐𝑖𝑎𝑙𝑖𝑠

∆N (2.3)

where α = trophic level of the food web baseline and DN = the trophic enrichment factor for

δ15N (average amplification) per trophic level. We took the abundant grazer Calanus

glacialis (α = 2) to represent the winter food web baseline. This decision was made because

Parasagitta elegans are thought to be carnivores that typically do not feed on primary

production but feed often on Calanus, and primary production is anyway low in winter (pers.

comm.: J Søreide).

We used a δ15N value for Calanus glacialis of 9.6 ± 0.2 ‰, estimated for individuals from

300 m in Svalbard fjords during December (Søreide et al. 2008). We used Δ = 3.4 ‰, also

determined for the European Arctic by Søreide et al. (2006).

2.4.9 Fatty acid analyses

Groups of primary producers and some herbivores produce fatty acids that are unique to

them, known as fatty acid trophic markers (FATMs, Dalsgaard et al. 2003). Higher

consumers possess a lower capacity to synthesise their own fatty acids, or to modify those

received from their prey. Therefore, when FATMs persist in a predator for some time with

little modification, they can be used to quantify its diet (Dalsgaard et al. 2003).

Page 51: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

29

Determination of the fatty acid signatures of Parasagitta elegans from Isfjorden and

Rijpfjorden (in 2012) were performed on triplicate samples from nine MIK hauls (Appendix

A) at UNILAB, Tromsø, Norway, using similar methods to Wold et al. (2011b). Each sample

contained 10 pooled individuals with dry weights between 0.001 and 0.05 g. Total lipid was

extracted in 15 ml 2:1 chloroform–methanol with butylated hydroxytoluene (BHT) (Folch et

al. 1957). Each sample was supplemented with a known amount of the fatty acid 21:0, as an

internal standard, and trans-methylated in methanol containing 1 % sulphuric acid with

toluene for 16 h at 50 ºC. The relative compositions (%) of FA methyl esters were determined

on an Agilent 6890 N gas chromatograph, equipped with a fused silica, wall-coated capillary

column with an Agilent 7683 injector and flame ionisation detection. The fatty acid methyl

esters were identified and quantified by gas chromatography. Results are given as relative

percentages of the various fatty acids identified in specimens from the two fjords. Differences

in average fatty acids proportions between sampled depths in Isfjorden and Rijpfjorden were

analysed using one-way ANOVA. Differences in the fatty acid profile between fjords was

analysed using the Mann-Whitney U test (sum rank test).

2.4.10 Mid-winter maturity status

Chaetognaths are hermaphrodites possessing male and female gonads (Alvarino 1992). The

male gonads (testes and seminal vesicles) produce and secrete sperm, whilst the female

gonads (ovaries and seminal receptacles) produce ova and receive sperm from a sexual

partner (Alvarino 1992). Mature Parasagitta elegans are characterised by advanced ovaries

containing large oocytes, high sperm loads filling the tail section of the animal, and pairs of

pronounced seminal receptacles and vesicles (Russell 1932, Choe et al. 2003). We described

the maturity of 29 of the largest individuals (40-51 mm, 8×1 mm size classes), and in 323

smaller individuals (13-39 mm, 26×1 mm size classes) in MIK samples from all fjords (both

years, Appendix A). Measured individuals were stained with borax carmine solution to

highlight their gonads (Pierce 1941). Maturity assessments were made under the binocular

microscope following the methods of Grigor et al. (2014). Ovaries were measured to the

nearest 0.1 mm, and individuals showing advanced ovary development were taken as those

with ovaries ≥5.4 mm, whilst shorter ovaries ˂5.4 mm were considered to be poorly

developed. The volume of loose sperm in the tail was estimated to the nearest 25 %. Low

Page 52: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

30

sperm loads could suggest that sperm is filling the tail or has already been released. However,

when sperm has already been released, remaining sperm appears more sparsely distributed.

Therefore, spent individuals can be separated from those which have not yet secreted sperm.

The size of the seminal receptacles was also described (see Grigor et al. 2014 for further

details).

2.5 Results

2.5.1 Chaetognath abundance and prey field

In Rijpfjorden in 2012, the most abundant zooplankton taxa based on the MPS sampling were

the larger calanoid copepod Calanus finmarchicus and the smaller copepods Pseudocalanus

spp. (Calanoida) and Oithona similis (Cyclopoida) (Table 2.1). Parasagitta elegans was the

most abundant chaetognath, but overall chaetognaths were also less abundant than the non-

copepod taxa Oikopleura spp. (Tunicata) and Beroe cucumis (Ctenophora) (Table 2.1).

Average Parasagitta elegans abundances in Rijpfjorden MIK trawls ranged from 2.4 ± 1.2

ind. m-3 (225 m) to 14.7 ± 7.0 ind. m-3 (75 m), comparing relatively well with the MPS results.

For further details on the polar night zooplankton community in Rijpfjorden, see Webster et

al. (2013), Kraft et al. (2013), and Daase et al. (2014).

Page 53: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

31

Table 2.1 Total water-column abundances (ind. m-3) of a polar night zooplankton community

(Rijpfjorden 2012), ordered according to abundance per taxonomic group. Net sampling

(Multi-Plankton Sampler; 0.25-m2 opening, 0.2-mm mesh) was used. As larger chaetognaths

may have avoided the smaller MPS net (see Grigor et al. 2014), chaetognath abundances

presented here are likely to be underestimates. Mean abundances for species and copepod

stages were first calculated over two hauls (one at midday and the other at midnight at

various depth intervals, see Appendix A), and data were summed across all sampling depths.

Copepod stages are CI-CV, AM (adult male) and AF (adult female). Functional (feeding)

groups were extracted from Søreide et al. (2003). “Small Calanoida” comprised the

following taxa: Acartia longiremis, Aetideidae CI-CIII, Bradyidius similis, Microcalanus

spp. and Pseudocalanus spp. Only taxa with abundances of ≥0.1 ind. m-3 are shown.

Taxonomic group Functional (feeding) group Abundance (ind. m-3)

Cyclopoida (Copepoda) Omnivores 1467.6

Small Calanoida (Copepoda) Omnivores 1257.5

Calanus spp. (Copepoda) CI-AF Herbivores 881.5

Metridia longa Omnivore 78.6

Tunicata Variable between species 36.9

Ctenophora Carnivores 23.5

Limacina helicina (Pteropoda) Herbivore 17.3

Parasagitta elegans (Chaetognatha) Carnivore 5.5

Clione limacina (Pteropoda) Carnivore 3.9

Other Mollusca Variable between species 3.9

Isopoda Variable between species 3.4

Eukrohnia hamata (Chaetognatha) Carnivore 1.4

Travisiopsis sp. (Polychaeta) Carnivores 1.2

Harpacticoida (Copepoda) Variable between species 0.9

Echinodermata Variable between species 0.6

Hydrozoan medusae Variable between species 0.6

Euphasiacea Variable between species 0.3

Apherusa glacialis (Amphipoda) Herbivore 0.1

Carnivorous Calanoid copepods Carnivores 0.1

Pseudomma truncatum (Mysidacea) Omnivore 0.1

Ostracoda Variable between species 0.1

Page 54: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

32

2.5.2 Gut contents

Observed gut contents were copepods, chitinous debris, lipid droplets, and other detritus that

could not be identified. The proportions of Parasagitta elegans with gut contents (FCRmax)

varied between fjords and years (Kruskal-Wallis test, P < 0.01), ranging from 9 to 53 %

(Figure 2.2). Some individuals contained two or more gut content types. Lipid droplets were

the most common gut observation overall (Figure 2.3 and Figure 2.4), observed in 106 (11.7

%) of the 903 chaetognaths. A total of 33 individuals (3.7 %) contained identifiable copepod

prey, including Calanus finmarchicus, Metridia longa, and harpacticoid copepods.

Proportions containing prey (FCRmin) also varied between fjords and years (Kruskal-Wallis

test, P ˂ 0.05), ranging from 0 to 10 % (Figure 2.2). FRs ranged from 0.00 to 0.24 prey ind.-

1 day-1 (mean = 0.12 prey ind.-1 day-1). Amongst feeding P. elegans, median per haul

proportions with prey varied with fjord (Kruskal-Wallis test, P ˂ 0.005), but proportions with

other gut contents did not (P ˃ 0.05, Figure 2.3). The relationship between chaetognath head

width and body length was linear (y = 0.031x + 0.37, R2 = 0.79), and prey was identified in

individuals with head widths ≥0.87 mm (Figure 2.4). Per haul proportions of P. elegans with

gut contents increased with head width size class (Kruskal-Wallis test, P ˂ 0.05, Figure 2.4),

but amongst feeders, proportions with each gut content type did not vary with head width (P

˃ 0.05, Figure 2.4).

Figure 2.2 Proportions (%) of Parasagitta elegans individuals per haul with gut contents

(FCRmax), identifiable prey (FCRmin), and empty guts in each fjord. The horizontal line inside

each boxplot shows the median of the proportions over multiple haul samples in a fjord. The

lower and upper boxes show the lower and upper quartiles, respectively, and the vertical

lines outside the boxes the differences between these quartiles and the lowest and highest

Page 55: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

33

proportions observed. Each dot represents an outlying data point. nhauls = numbers of hauls

for each fjord. Hauls with ˂3 individuals analysed were not included. As only one haul was

analysed for Rijpfjorden in 2013, full boxplots could not be shown. See Appendix A for

numbers of individuals analysed per haul.

Figure 2.3 Proportions (%) of feeding Parasagitta elegans individuals per haul in each fjord

with different types of gut content. For details on the features of the boxplots and the data,

see Figure 2.2.

Figure 2.4 Gut contents in ascending head width size classes: proportions of Parasagitta

elegans (%) per haul with gut contents and of feeders with each gut content type. Includes

all dissected specimens from Isfjorden (50 individuals, 4 hauls) and Rijpfjorden (152

individuals, 13 hauls) in 2012. For details on the features of the boxplots and the data, see

Figure 2.2. ninds. = total numbers of individuals for each size class.

Page 56: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

34

2.5.3 Body composition

Values for carbon and nitrogen isotopes varied little between fjords and sampling depths

(Table 2.2). The δ13C content of Parasagitta elegans ranged from -22.0 ± 0.3 ‰ (Rijpfjorden)

to -21.5 ± 0.2 ‰ (Isfjorden), whilst δ15N content ranged from 12.5 ± 0.3 % to 12.9 ± 0.1 ‰

(both Isfjorden). Average trophic level (TL) based on the δ15N content was 2.9, and average

C/N ratio over fjords and depths was 3.1 (Table 2.2).

Table 2.2 Stable carbon and nitrogen isotope values for Parasagitta elegans sampled by the

MIK (3.14-m2 opening, 1.5-mm mesh) at various trawl depths (20, 30, 35, 60, 75 and 225 m)

in Isfjorden and Rijpfjorden (2012). The average δ13C and δ15N composition (‰) in replicate

samples (usually three but six from 225 m in Rijpfjorden) containing 25 pooled individuals

(10-50 mm), average proportions of animal dry weight (DW) comprising carbon and

nitrogen (%), and C/N weight ratios. All values are accompanied by standard deviations.

Trophic levels (TL) were calculated for P. elegans from each fjord from mean δ15N (‰)

values (see ‘Method’).

Parasagitta elegans fatty acid signature did not differ between Isfjorden and Rijpfjorden

(Mann-Whitney U test, P ˃ 0.05). Similarly, percentage composition of almost all fatty acids

did not differ between sampling depths in the two fjords (one-way ANOVA, P ˃ 0.05). The

high levels of 18:1 (n-9) compared with those of 18:1 (n-7) strongly indicate carnivory. The

Calanus fatty acid marker 20:1 (n-9) and 22:1 (n-11) were both abundant. The dinoflagellate

marker docosahexaenoic acid (DHA) 22:6 (n-3) was recorded in interestingly high amounts,

Location Trawl depth (m) δ13C (‰) δ15N (‰) DW%C DW%N C/N TL

Mouth of Isfjorden 250 -21.7±0.0 12.5±0.3 48.4±0.5 15.6±0.4 3.1±0.0 2.9

Isfjorden (ISA) 30 -21.5±0.2 12.9±0.1 48.4±0.4 15.5±0.1 3.1±0.0 3.0

-"- 35 -21.5±0.1 12.8±0.2 47.9±1.1 15.4±0.4 3.1±0.0 2.9

-"- 60 -21.6±0.1 12.9±0.1 48.3±0.8 15.5±0.4 3.1±0.0 3.0

Rijpfjorden 20 -22.0±0.0 12.5±0.2 48.7±0.3 15.8±0.1 3.1±0.0 2.9

-"- 75 -21.8±0.2 12.5±0.2 48.0±0.2 15.4±0.0 3.1±0.0 2.9

-"- 75 -22.0±0.3 12.6±0.3 48.9±0.4 15.7±0.2 3.1±0.0 2.9

-"- 225 -21.8±0.1 12.5±0.1 48.6±0.4 15.8±0.2 3.1±0.0 2.9

Page 57: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

35

whilst eicosapentaenoic acid (EPA) 20:5 (n-3) was recorded in moderate to high amounts

(Table 2.3).

Table 2.3 Average fatty acid profile for Parasagitta elegans in 2012. Results are given as

average percentages of the various fatty acids identified across all samples from Isfjorden

and Rijpfjorden (see ‘Method’), alongside the standard deviations. Only fatty acids with

mean percentages of =>0.5 ± 0.1 across both fjords are shown. In Isfjorden, the mean

percentages of 15:0 FA and 18.2(n-6) FA differed between sampling depths (indicated by a

† symbol, one-way ANOVA, P ˂ 0.05). In Rijpfjorden, the proportion of every fatty acid was

similar between sampling depths (one-way ANOVA, P ˃ 0.05).

Fatty acid Mean % Isfjorden

(n samples = 12)

Mean % Rijpfjorden

(n samples = 17)

14:0 FA 5.02±2.09 4.13±0.62

14:1 (n-5) FA 0.54±0.06 0.52±0.08

15:0 FA 0.72±0.07† 0.73±0.06

16:0 FA 11.91±1.16 12.65±1.47

16:1 (n-5) FA 2.97±0.41 3.30±0.39

16:1 (n-7) FA 7.13±0.27 6.05±0.33

18:0 FA 1.09±0.13 1.18±0.20

18:1 (n-7) FA 1.95±0.28 1.74±0.18

18:1 (n-9) FA 5.35±0.65 5.82±0.54

18:2 (n-6) FA 1.28±0.11† 1.35±0.09

18:3 (n-3) FA 1.11±0.11 1.37±0.07

18:4 (n-3) FA 1.76±0.63 2.35±0.35

20:1 (n-9) FA 14.80±1.02 11.46±1.71

20:4 (n-3) FA 0.82±0.10 0.99±0.07

20:5 (n-3) FA (EPA) 11.41±0.68 12.73±0.75

22:1 (n-11) FA 6.77±1.38 5.69±0.71

22:1 (n-7) FA 0.47±0.02 0.78±0.30

22:5 (n-3) FA 0.60±0.14 0.69±0.15

22:6 (n-3) FA (DHA) 18.01±2.33 18.81±1.99

24:1 (n-9) FA 2.88±0.21 3.81±0.62

Page 58: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

36

2.5.4 Mid-winter maturity status

Amongst the largest individuals in the population (body lengths 40-51 mm), 55 % had well-

developed seminal receptacles, whereas 93 % had relatively high sperm volumes in their tails

(50-100 %) and all had advanced ovaries (≥5.4 mm). In smaller individuals (lengths 13-39

mm), 45 % had well-developed seminal receptacles and 42 % had advanced ovaries; 80 %

had synthesized relatively high sperm volumes (filling 50-100 % of the tail area).

2.6 Discussion

2.6.1 Studies during the polar night

Studies of plankton ecology in the high Arctic are typically restricted to spring and summer,

due to the logistical difficulties of sampling during winter. The polar night cruise, of which

our study formed a part, offered unique possibilities for a better understanding of polar night

marine ecology (see Berge et al. 2012, Webster et al. 2013, Daase et al. 2014).

We show that the common chaetognath Parasagitta elegans remains an active carnivore

during the polar night. This arrow worm is not in a dormant and non-feeding state although

its feeding rates may be considerably lower than in spring and summer, and reproduction is

absent (or occurring at very low rates) at this time of year. These observations add to the

increasing awareness of activity levels and ecological interactions of pelagic organisms

during the polar night.

2.6.2 Feeding activity and rates

Copepods are abundant in Svalbard fjords during winter (Table 2.1) and as a non-visual

predator Parasagitta elegans should, as opposed to fish and other visual predators, be able

to encounter and catch them also during the polar night. The behaviour of some copepods

during the polar night (i.e. diapausing in an unalert state) could make them particularly

susceptible to predators such as chaetognaths (Darnis et al. 2012). About a quarter of the P.

elegans individuals showed signs of recent feeding (mainly lipid droplets). The body of P.

elegans does not contain a centralised oil vacuole, suggesting that all lipid droplets in P.

elegans guts remained from recently digested prey, yet only 4 % contained identifiable prey.

Page 59: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

37

Based on Feigenbaum’s (1982) digestion time estimate of 10.2 h at 0 ºC (likely longer in

colder waters), our findings would suggest that most individuals had not fed for several hours

before their capture. In contrast, Falkenhaug (1991) identified prey in 36 % of the P. elegans

population. collected from the Barents Sea in summer. Our average estimate of per capita

feeding in P. elegans was 0.12 prey ind.-1 day-1, corresponding to 0.66 prey ind.-1 m-3

consumed per day by the P. elegans population (based on the abundance data in Table 2.1).

Their predation impact on Calanus is therefore low given the high Calanus abundances

(Table 2.1, see also Daase et al. 2014). Calanus are assumed to be in a dormant, non-feeding

state during winter (Falk-Petersen et al. 2009). Daase et al. (2014) reported on Calanus from

the same cruise in 2012 and found that the Calanus finmarchicus population in Rijpfjorden

comprised mainly copepodite stage CV and some CIVs, which could indicate overwintering.

However, these authors also noted that “the bulk of the C. finmarchicus and C. glacialis

population was found close to the surface and not at greater depth where they presumably

should overwinter”, suggesting they were not in an unalert dormancy phase, at least not in

January. If not, these copepods may be more alert to the presence of predators than assumed,

allowing them to avoid or escape chaetognaths.

Our FR estimate is similar to that obtained for immature specimens in northern Sweden in

autumn-winter (0.2 prey ind.-1 day-1) and much lower than reported for spring-summer (0.7-

0.9 prey ind.-1 day-1, Øresland 1987). These findings suggest that feeding rates for

Parasagitta elegans at high-latitudes drop during winter, corresponding well with the

reduced growth rates observed during winter (Grigor et al. 2014), and the lack of reproduction

at this time of year (this study). In less seasonal environments, energetic requirements may

vary less between the seasons, accounting for relatively higher feeding rates at lower

latitudes, also during winter and early spring (e.g. up to 1.33 prey ind.-1 day-1 in Vineyard

Sound, Feigenbaum 1982).

However, care should be taken when comparing estimates of feeding rates between studies,

because each study detects prey with various levels of precision. For example, we did not

search for mandible remains in guts to detect further signs of recent copepod digestion, as in

Page 60: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

38

other studies (e.g. Falkenhaug 1991, Giesecke and Gonzalez 2004). Feeding estimates in

chaetognaths are also highly sensitive to sampling methodology. As well as the risk of

damage to their fragile bodies, stressed chaetognaths may also regurgitate gut contents,

leading to underestimates of true feeding rates. Baier & Purcell (1997) estimated prey loss

from guts of ~50 % when tows were longer than 2 min. Such short hauls can be unfeasible,

especially when using large trawl nets such as the MIK, and when sampling populations in

deeper waters. Further studies should therefore utilize new zooplankton imaging devices for

observing chaetognaths in the water column without necessarily capturing them in nets.

Optical methods are promising avenues (e.g. Schulz et al. 2010, Sainmont et al. 2014b).

2.6.3 Energetics

Stable isotope and FATM analyses confirmed the position of Parasagitta elegans as a

predator during the polar night. From its average trophic level (TL) of 2.9, it can be classified

as a carnivore according to a trophic model devised for the European Arctic (Søreide et al.

2006, in which carnivores had TLs between 2.9 and 3.3). δ15N (an indicator of protein

content) and δ13C (an indicator of organic matter content) values varied little with station or

depth, and agree well with values from the Barents Sea in March (Søreide et al. 2006). δ15N

values were, however, lower than reported from the Bering Sea (14.7 ± 0.7 ‰, Lovvorn et

al. 2005), suggesting that P. elegans in the sub-Arctic Pacific typically feed higher up the

food chain (also see ‘Lipid profile’ section of ‘Discussion’). Many factors affect the

elemental composition of zooplankton species, including age, size maturity status,

reproductive strategies, as well as variations in the composition and growth rates of primary

producers (Ikeda 1974), yet for P. elegans in the Barents Sea, δ15N values seem to vary little

throughout the year. Søreide et al. (2006) reported δ15N values ranging from 11.9 ± 0.2 ‰ in

spring to 12.2 ± 0.1 ‰ in winter. δ13C values varied slightly more between spring (-19.3 ±

0.8 ‰) and winter (-20.8 ± 0.3 ‰).

2.6.4 Lipid profile

The high levels of the FATMs 20:1 (n-9) and 22:1 (n-11) confirm that Parasagitta elegans

is part of the Arctic Calanus-based food web (Falk-Petersen et al. 2007, Wold et al. 2011b).

Page 61: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

39

Furthermore, the high level of 22:6 (n-3) indicates that the base of the food chain is dominated

by dinoflagellates, which can be very abundant in waters north of Svalbard in spring-summer

(Hegseth & Sundfjord 2008). This lipid signature agrees well with previous results from the

European sub-Arctic (Falk-Petersen et al. 1987). Both the presence of green detritus in P.

elegans guts, and the surprisingly high levels of the dinoflagellate fatty acid marker

docosahexaenoic acid, may suggest that some omnivory and/or detritivory occurs. Non-

carnivorous feeding in chaetognaths has been suggested by Casanova et al. (2012). In the

Canadian Arctic, Eukrohnia hamata and Pseudosagitta maxima have been observed

ingesting green detritus under the microscope (pers. obs.). Eukrohnia species in Antarctica

were found to contain higher amounts of 16:0, reflecting feeding on the Antarctic copepods

including Rhincalanus gigas (Kruse et al. 2010). We suggest that winter feeding is captured

in these results, following the suggestions of Graeve et al. (2005) that a new signal (source

of food) will be visible relatively quickly (~4-8 days after feeding). If P. elegans starves

during autumn and early winter, and if 22:6 (n-3) undergoes little transformation or

metabolism (Dalsgaard et al. 2003), it remains possible that this signal is retained from

feeding on Calanus during a previous summer or autumn bloom (Falk-Petersen et al. 1990,

Wold et al. 2011b). However, this study and others (e.g. Feigenbaum 1982, Øresland 1987,

Falkenhaug 1991) suggest a low likelihood of long-term fasting in P. elegans.

An important difference between the two main chaetognath species in Arctic waters,

Eukrohnia hamata and Parasagitta elegans, is that the former typically possess oil vacuoles

in the centre of their bodies, whilst the latter do not (Øresland 1990, Pond 2012, pers. obs.).

Previous studies that considered the presence of oil vacuoles in this species to reflect recent

feeding may therefore have overestimated feeding rates (e.g. Sameoto 1987, Froneman and

Pakhomov 1998). The role of these vacuoles in E. hamata (e.g. buoyancy, storage) is not yet

clear (Pond 2012), but unlike P. elegans, E. hamata has a cosmopolitan range. In the Arctic,

both species commonly occur at epi-pelagic depths (Bieri 1959), but in the North Pacific, E.

hamata populations are commonly found residing deeper, and throughout a wider depth

range than P. elegans (e.g. Bieri 1959, Alvarino 1964). Maintaining such wide vertical

distributions would certainly require a strong control of buoyancy, which could be offered

by having centrally positioned oil vacuoles (Pond 2012). Lipid reserves form a central

Page 62: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

40

component of the life history of polar zooplankton species (Falk-Petersen et al. 2009, Varpe

et al. 2009). In Antarctica, 34 % of examined E. bathyantarctica individuals were found to

contain oil droplets in summer compared to 57 % of individuals in winter. This species also

contained relatively high amounts of the fatty acid 18:1 (n-9), which is found in storage lipids

(Kruse et al. 2010). If the primary role of oil vacuoles is storage, this could suggest that

survival and possibly reproduction in E. hamata is less dependent on concurrent food intake

than in P. elegans. Studies of the extent of capital breeding (Varpe et al. 2009) in E. hamata

are therefore needed.

2.6.5 Reproduction

Small Parasagitta elegans <10 mm were absent from MPS samples from January, and whilst

many individuals above 20 mm had synthesised high sperm loads, many of the largest P.

elegans specimens in the nets (40-51 mm) still lacked well developed seminal receptacles,

suggesting that they were not fully mature (Russell 1932, Choe et al. 2003). Grigor et al.

(2014) frequently observed pronounced receptacles in individuals ≥20 mm from February to

May, after which they disappeared. In the Arctic, breeding of P. elegans generally takes place

in spring and summer (Kramp 1939, Ussing 1939, Bogorov 1940, Grigor et al. 2014); this is

in contrast to Eukrohnia hamata, which may also reproduce in winter (pers. obs.). Saito &

Kiørboe (2001) showed that P. elegans <5 mm in the North Sea fed almost exclusively on

prey <350 μm in length. In Svalbard, the prey in this size range is represented by small

cyclopoids such as Oithona similis, as well as Calanus nauplii and the young Calanus stages.

In winter, individuals may be waiting for increased food input before they reproduce or for

suitable food for the young and newborns to feed on (in terms of size and availability). By

investing energy into maturation during winter, egg hatching can be scheduled to coincide

with the reproduction of a wide range of copepod prey in spring and summer, and the

buoyancy of chaetognath eggs may also allow them to hatch in shallow waters (Hagen 1999

and references therein) amongst their grazing prey. In another fjord on Svalbard, large

numbers of P. elegans eggs became visible in the water column in March 2003, suggesting

that conditions for reproduction begin to improve at this time of year (Hirche & Kosobokova

2011).

Page 63: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

41

2.7 Concluding remarks

Knowledge of an organism’s activity level and foraging ecology outside of the windows of

the main primary production period (spring–summer) is key to establishing an understanding

of the full annual routine (Varpe 2012). Here, we have focused on the polar night feeding

ecology of a predatory zooplankter, adding to recent work on the life history and vertical

distribution (Grigor et al. 2014). We found the chaetognath Parasagitta elegans to feed

during the polar night, but with feeding rates lower in winter than have been reported from

other seasons. As growth rates (e.g. Dunbar 1962, Grigor et al. 2014) may also decrease at

this time and reproduction does not seem to occur, individuals should have lower-energy

requirements and require less food. Although several copepod taxa may rest at depth in an

unalert state during the Arctic winter, our study shows that mortality on copepods caused by

P. elegans in the water column is low at this time of year. This study did, however, not sample

the zone immediately above the seafloor, the hyperbenthic zone, where high densities of

chaetognaths are known to aggregate during winter (Choe & Deibel 2000), and where

abundances of resting copepods may also peak. The activities of P. elegans in this zone, as

well as those of the other Arctic species Eukrohnia hamata and Pseudosagitta maxima,

require more attention.

Page 64: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

42

3. Chapter 3 – Growth and reproduction of the chaetognaths

Eukrohnia hamata and Parasagitta elegans in the Canadian Arctic

Ocean: capital breeding versus income breeding

3.1 Résumé

Dans les mers arctiques, la production primaire et l’accessibilité de la nourriture du

zooplancton varient fortement durant la courte période estivale. Nous avons testé l'hypothèse

selon laquelle Eukrohnia hamata et Parasagitta elegans, deux chaetognathes arctiques

similaires et sympatriques, partitionnent les ressources selon différentes stratégies de

reproduction. Les deux espèces avaient des longévités naturelles similaires d'environ 2 ans.

Espèce mésopélagique, E. hamata engendre deux couvées distinctes à l'automne et au

printemps. La production de jeunes coïncide avec une baisse de la fréquence d’E. hamata

présentant des réserves lipidiques visibles, ce qui est caractéristique d’une reproduction basée

sur les réserves. La croissance s’est avérée positive d'avril à janvier et négative en février-

mars. La croissance et la maturation étaient similaires pour les deux cohortes. Les réserves

lipidiques contenues dans une vacuole permettent à E. hamata de se reproduire et de croitre

en dehors de la courte saison de productivité. Espèce plus néritique, P. elegans produit une

couvée en été-automne, durant la production biologique maximale dans les eaux près de la

surface, un mode de reproduction basé sur l’apport immédiat d’énergie. Cependant, avec le

réchauffement climatique réchauffe, un bloom phytoplanctonique automnal pourrait

favoriser la couvée d'été-automne chez P. elegans.

Page 65: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

43

3.2 Abstract

In Arctic seas, primary production and the availability of food to the zooplankton are strongly

pulsed over the short productive summer. We tested the hypothesis that Eukrohnia hamata

and Parasagitta elegans, two similar and sympatric arctic chaetognaths, partition resources

through different reproductive strategies. The two species had similar natural longevities of

around 2 years. The meso-pelagic E. hamata spawned two distinct broods in autumn and

spring. Offspring production coincided with drops in the frequency of E. hamata with visible

lipid reserves, characteristic of capital breeders. Growth was positive from April to January

and negative in February and March. Growth and maturation were similar for the two broods.

Storage reserves contained in an oil vacuole may allow E. hamata to reproduce and grow

outside the short production season. The neritic P. elegans produced one brood in summer-

autumn during peak production in near-surface waters, characteristic of income breeders.

However, as the Arctic warms, the development of an autumn phytoplankton bloom could

favour the summer-autumn brood of P. elegans.

Page 66: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

44

3.3 Introduction

Arrow worms or chaetognaths form a phylum of gelatinous zooplankton represented by ~200

species worldwide, with biomass estimated at 10-30 % that of copepods (Bone et al. 1991).

They can represent a significant fraction of the diet of many animals such as amphipods (e.g.

Gibbons et al. 1992), other chaetognaths (e.g. Pearre 1982), seabirds (e.g. Mehlum and

Gabrielsen 1993), and are the preferred prey of the larvae of some tropical fish and

commercially important decapods (e.g. Sampey et al. 2007, Saunders et al. 2012).

Chaetognaths are hermaphrodites. Spermatogonia bud off from testes parallel to the tail walls

to produce spermatocytes, which are transmitted to a conspecific to fertilise its oocytes

(Alvarino 1992, Bergey et al. 1994). Each chaetognath possesses one pair of seminal vesicles

on the tail segment and one pair of seminal receptacles at the posterior ends of the oviducts.

Two individuals first come into contact with each other, before sperm is passed from the

seminal vesicles of one individual to the seminal receptacles of the other (Goto & Yoshida

1985). Hatching strategies vary amongst species. For instance, Parasagitta elegans releases

buoyant eggs that hatch near the surface where small prey is abundant (Kotori 1975, Hagen

1985). Eukrohnia hamata broods its eggs in an external sac from which the young later swim

out, often at depth (Alvarino 1968 and references therein).

In contrast to the high diversity of chaetognaths in other seas (e.g. De Souza et al. 2014), only

three species are frequently reported in Arctic plankton surveys. Parasagitta elegans is a

neritic species, often peaking in abundance in epi-pelagic waters; Eukrohnia hamata is

abundant in meso-pelagic and deep waters; and Pseudosagitta maxima, growing much larger

(up to 90 mm) than the other species (40-45 mm), is typically bathy-pelagic but may also

occur near the surface in the Arctic (Bieri 1959, Alvarino 1964, Terazaki & Miller 1986,

Sameoto 1987). A fourth species, Heterokrohnia involucrum has also been recorded at bathy-

pelagic depths, so far only in the Arctic Ocean (Dawson 1968). Whilst their contribution to

higher trophic levels is poorly documented, arctic chaetognaths may consume an important

fraction of copepod biomass. Although admittedly little better than a guess, Welch et al.

Page 67: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

45

(1992) estimate that chaetognaths in Lancaster Sound ingest 51 % of the copepod biomass

annually (164 of 319 kJ m-2 yr-1).

At high arctic latitudes, resource availability is strongly pulsed seasonally, with maximum

phytoplankton and zooplankton biomass occurring during the short ice-free season in late

summer and autumn (e.g. Falk-Petersen et al. 2007, Falk-Petersen et al. 2009, Søreide et al.

2010). Here, the co-existence of similar species may depend on the seasonal partitioning of

resources, which can be achieved through differences in vertical distribution, diet and

reproductive strategies. For instance, the reproductive strategies of the two dominant arctic

herbivorous copepods differ markedly in ice-covered waters. The large capital-breeder

Calanus hyperboreus reproduces in winter, fuelling egg production with lipids accumulated

in summer (e.g. Conover 1967, Hirche & Niehoff 1996, Pasternak et al. 2001). Calanus

glacialis may combine lipid stores and high-quality food to fuel reproduction (e.g. Conover

1988, Sainmont et al. 2014a). Income breeders do not synthesise reserves as fuel for

reproduction at a later time. Peak reproduction occurs in spring-summer in the epi-pelagic

Parasagitta elegans (e.g. Kramp 1939, Sameoto 1971, Grigor et al. 2014), which suggests

income breeding with the summer maximum in prey abundance fuelling reproduction. By

contrast, year-round breeding in Gerlache Strait, Antarctica (Øresland 1995), and the

accumulation of lipids and wax esters in an oil vacuole (Pond 2012) suggest capital breeding

fuelled by lipid reserves in E. hamata.

We describe the life histories of Eukrohnia hamata and Parasagitta elegans over an annual

cycle based on weekly sampling of zooplankton in the Amundsen Gulf (Beaufort Sea) from

August 2007 to July 2008, complemented by collections from other areas of the Canadian

Arctic in autumn of the two years. In particular, growth of size cohorts, reproduction, and oil

reserves are tracked to explore the hypothesis that E. hamata is a capital breeder that spawns

during most of the year, whilst P. elegans is an income breeder, reproducing during periods

of high prey availability.

Page 68: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

46

3.4 Method

3.4.1 Study area

Sampling was conducted on-board the CCGS Amundsen, primarily from October 2007 to

August 2008 in the Beaufort Sea (Figure 3.1). To complete the annual cycle of observations,

additional collections from the Labrador Sea, the Canadian archipelago and Baffin Bay were

included in the analyses (Figure 3.1). In addition to the 27 stations sampled at weekly or

higher resolution in the Amundsen Gulf region (69-72°N, 121-131°W) from November 2007

to August 2008, two stations were sampled in Nachvak Fjord in August 2007, two stations

in Parry Channel in October 2007, and two stations in northern Baffin Bay in September

2008 (Figure 3.1 and Appendix C). Amundsen Gulf, our main sampling region, is a 400 km

long, 170 km wide channel with a maximum bottom depth of ~630 m that connects the south-

eastern part of the Beaufort Sea to the Canadian archipelago. Amundsen Gulf is typically

covered with sea-ice from October to early June (Barber & Hanesiak 2004). Three water

masses are detected in the Amundsen Gulf (Geoffroy et al. 2011 and references therein); the

Pacific Mixed layer (PML; 0-60 m), the Pacific Halocline (PH; 60-200 m) and the Atlantic

layer (AL; ˃ 200 m). Nachvak Fjord is a pristine 45-km long sill-fjord in Nunatsiavut,

Northern Labrador, with a maximum depth of 210 m (Bell & Josenhans 1997, Simo-Matchim

et al. 2016). The deep semi-enclosed basin of Baffin Bay lies further north, and is connected

via Parry Channel to the Beaufort Sea in the west (Figure 3.1).

Page 69: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

47

Figure 3.1 Bathymetric maps of the Canadian Arctic Ocean indicating the regions, and

positions of stations (black circles) where chaetognaths were sampled from August 2007 to

September 2008. Station IDs provided.

3.4.2 Sampling

In the Amundsen Gulf, estimates of weekly profiles of temperature and salinity between

November 2007 and July 2008 are from Geoffroy et al. (2011), and those of chlorophyll a

(chl a) biomass are from Darnis & Fortier (2014). Salinity, and temperature profiles of the

water column at single stations in Nachvak Fjord, Parry Channel and Baffin Bay, were

obtained using a conductivity-temperature-depth (CTD) system. We also collected seawater

samples from 9-10 depths in the upper 100 m, using the CTD rosette equipped with twenty-

four 12 1 Niskin-type bottles (OceanTest Equipment). Subsamples (500 ml) for the

determination of total chlorophyll a (chl a) were filtered onto 25 mm Whatman GF/F filters

Page 70: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

48

(nominal porosity of 0.7 µm). Chl a was measured using a Turner Designs 10-AU

fluorometer, after 24 h extraction in 90 % acetone at 4 °C in the dark without grinding

(acidification method of Parsons et al. 1984).

Two similar samplers were used to collect zooplankton: a square-conical net with a 1 m2

opening area, 200 µm mesh and a rigid cod-end, and a Hydrobios® Multinet with 9 individual

opening and closing nets with 0.5 m2 openings, 200 µm mesh and rigid cod ends. At 12

stations, the 1 m2 net was deployed cod-end first (non-filtering) to 10 m above the seabed

and hauled back to the surface at a speed of 0.5 m s-1. At 22 stations, the Multinet was

deployed to 10 m above the seabed and retrieved at 0.5 m s-1, sampling sequentially over pre-

set depth intervals, the overall thickness of which varied from 100 to 520 m. Typically, 3×20-

m depth strata were sampled from 10 m above the seafloor upward, and 3×20-m depth strata

were sampled from 60 m depth to the surface. The remaining interval (from 70 m above the

bottom to 60 m below the surface) was then divided into three equal sampling layers

(Appendix C). Samples were fixed in 4 % formaldehyde-seawater solution buffered with

sodium borate.

3.4.3 Chaetognath body size and sampler efficiency

Chaetognaths were sorted from the samples, identified to species and when possible,

measured (nearest mm) from the top of the head to the tip of the tail, excluding the caudal

fin. Heavily damaged or broken individuals were excluded, but those missing heads only (~1

mm) were included in the analyses, as positive identification remained possible (Figure 3.2)

and length could be estimated. Out of 8179 chaetognaths sorted, 7245 chaetognaths (5918

Eukrohnia hamata and 1298 Parasagitta elegans were included. Twenty-nine (29)

Pseudosagitta maxima present in the samples were not analysed further. The relative

efficiency of the Multinet and square-conical net at capturing chaetognaths was assessed by

comparing the size frequency distribution of E. hamata or P. elegans in the two nets.

Page 71: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

49

Figure 3.2 Illustrations and photographs of Eukrohnia hamata and Parasagitta elegans. (a)

Diagrams of the two species indicating maturity features and the centrally-positioned oil

vacuole in E. hamata. (b) Photographs of live E. hamata (top) and P. elegans (bottom) taken

in-situ by a zooplankton imager (Schmid et al. 2016) in the Canadian Arctic. Specimens ~20

mm. (c) Photograph of oocytes in a stained P. elegans individual. (d) Photograph of ovaries

in an E. hamata individual (imaged in-situ). (e) Photograph of tail sperm, seminal vesicles

and seminal receptacles in a stained P. elegans individual.

3.4.4 Hatching and cohort development

The occurrence of newborns (2-4 mm) in Multinet collections was used as an index of recent

hatching. The subsequent growth of the different cohorts was inferred visually from

successive monthly length frequency distributions in collections from both samplers. Lengths

from the Multinet collections were included twice in the histograms to give equal weight to

this sampler (0.5 m2 aperture) and the square-conical net (1 m2 aperture). Each cohort was

detected by visual means, and the MIXdist library in RTM (Du 2002) was used to estimate its

mean length (arithmetic mean), standard deviation and length range in each month. Monthly

Page 72: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

50

growth rates of each cohort were calculated from changes in mean length. In most cases, a

cohort was clearly defined in the month length frequency distributions. In some cases, the

interpretation of cohorts was aided by two assumptions. First, based on published estimates

of growth (Welch et al. 1996, Choe & Deibel 2000, Grigor et al. 2014), we assumed that

monthly growth of a cohort could not exceed 7 mm mo-1. Second, the possibility that a cohort

was absent from the population in one month if clearly present during the preceding or

following month in the time series was excluded. Solutions aimed to minimise cohort length

overlaps. Goodness-of-fit was tested using the Chi-square statistic. Parasagitta elegans data

for August are from Nachvak Fjord in 2007 due to the scarcity of this species in the

Amundsen Gulf collections in that month in 2008.

3.4.5 Estimation of maturity and oil vacuole area

Eukrohnia hamata and Parasagitta elegans were stained with a Borax Carmine solution to

highlight gonads (Pierce 1941, Grigor et al. 2014). Oocytes (Figure 3.2) in single ovaries

were counted and diameters measured under the stereomicroscope. Ovary length, diameter

of the externally protruding part of the seminal receptacles and width of the seminal vesicles

(Figure 3.2) were measured to the nearest 0.01 mm. The distribution of

spermatogonia/spermatocytes in tails was characterised as absent, spermatogonia present

near walls only, or spermatocytes dispersed throughout the tail. The width of seminal vesicles

was measured to the nearest 0.01 mm. The presence or absence of an oil vacuole (Figure 3.2)

was noted in 2089 individuals of E. hamata. The width and height of the vacuole were

measured to the nearest 0.01 mm under the stereomicroscope. Area was calculated as

π×0.5width×0.5height. The presence of oil in the vacuole and/or oil droplets in the gut was

noted.

3.4.6 Vertical distributions

19 Multinet collections (typically 9 samples per haul), 1-2 per month from August 2007 to

September 2008, were analysed to estimate the abundances of Eukrohnia hamata and

Parasagitta elegans populations and age classes in discrete layers of the water column (ind.

m-3). Contour plots were produced in the R computing environment (R Development Core

Page 73: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

51

Team 2008), using Akima bilinear interpolation performed vertically for every meter of the

water column. Assuming a lognormal distribution, species abundances A were log-

transformed as log10(A+1) prior to interpolation, also serving to preserve abundance A = 0

(Legendre & Legendre 2012). Vertical distributions of populations were also characterized

by their weighted mean depths in the water column [Zm, m, Eq. (3.1a)] with standard

deviations [Zs, m, Eq. (3.1b)] as described by Dupont & Aksnes (2012). Zm was normalized

between 0 and 1 to identify the positions of animals relative to a station’s bottom depth.

𝑍𝑚 = ∑ 𝑤𝑖𝑛𝑖=1 𝑑𝑖𝑧𝑖/ ∑ 𝑑𝑖

𝑛𝑖=1 𝑧 (3.1a)

𝑍𝑠 = √(∑ 𝑤𝑑1𝑛𝑖=1 𝑧1

2/ ∑ 𝑑1𝑧1𝑛𝑖=1 ) − 𝑊𝑀𝐷2 (3.1b)

Where wi is the abundance of animals in depth interval i, di is the thickness of interval i, and

zi is its mid-point.

3.5 Results

3.5.1 Physical environment and primary production in the Amundsen Gulf

From November 2007, onwards, the PML in the Amundsen Gulf had temperatures of -1.6 to

-0.6 ºC, but warmed to ˃ 9 ºC at some stations in June and July 2008 on top of a strong

thermocline. Throughout the study, temperatures in the PH were -1.6 to 0 ºC, and

temperatures in the AL were ~0 ºC. Salinities were 30-32 psu in the PML, 32-34.6 psu in the

PH, and ˃ 34.6 psu in the AL (Geoffroy et al. 2011). Ice algae was first detected in late March

in the PML and peaked in late April and early May (~10 mg m-3), when it was succeeded by

a surface phytoplankton bloom. In July, a subsurface chl a maximum (SCM) occurred at ~40

m (16 mg m-3), and the phytoplankton penetrated the PH (Darnis & Fortier 2014).

3.5.2 Physical environment and primary production in autumn (other sampling

locations)

In August 2007, a steep thermocline was observed in Nachvak Fjord between the surface and

60 m, with temperature decreasing from 3.4 °C to -1.6 °C (Appendix D). Chl a concentrations

peaked at 20 m (2.8 µg L-1). In October 2007, temperatures at our Parry Channel station rose

Page 74: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

52

from ˂-1 °C at the surface to 0.4 °C at ~45 m, fell to ˂ 1 °C at 60 m, and then rose again from

-1.5 °C at 200 m to 0.7 °C at 500 m. Chl a concentrations remained below 0.5 µg L-1 at

sampled depths. In September 2008, temperatures at our Baffin Bay station rose from ˂-1 °C

at the surface to ˃ 0.3 °C at ~85 m, fell to ˂ 1.1 °C at 100 m, with smaller fluctuations in

temperature in deeper waters. Chl a concentrations peaked at 45 m (1.1 µg L-1). Strong

haloclines were observed in the upper 100 m at all three stations (Appendix D).

3.5.3 Abundances and vertical distributions of chaetognath species

Eukrohnia hamata was present at all stations in the Amundsen Gulf with abundances ranging

from 56 ind. m-2 in late July to 894 ind. m-2 in mid-April, based on Multinet collections.

Parasagitta elegans was less abundant than E. hamata at all stations in Amundsen Gulf (16

ind. m-2 in November to 244 ind. m-2 in December), and in Parry Channel (320 ind. m-2) and

Baffin Bay (64 ind. m-2) in October and September respectively. Pseudosagitta maxima

occurred in only 4 Multinet collections from the Amundsen Gulf, typically in the Atlantic

Layer, with maximum abundances of 4 ind. m-2. Relative to P. elegans, the abundance of E.

hamata increased with station depth (Figure 3.3a). E. hamata occurred primarily in the deep

AL of the Amundsen Gulf (˃ 200 m, Figure 3.3b), based on weighted mean depth, except on

two dates in late May and late July when the E. hamata population was centred in the PH

(60-200 m). In Parry Channel (October) and Baffin Bay (September), the E. hamata

population remained ˃ 200 m. In the Amundsen Gulf, seasonal vertical migration (SVM) was

most pronounced in P. elegans, which, based on weighed mean depth, resided primarily in

the PH from December to March, rose into the PML (0-60 m) in March/April during the

development of ice algae, and returned to the PH in late June (Figure 3.3b). P. elegans was

39 times more abundant than E. hamata in Nachvak Fjord in August (462 ind. m-2). At

Nachvak Fjord, Parry Channel and Baffin Bay, the P. elegans population remained ˂ 200 m

(Figure 3.3b).

Page 75: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

53

Figure 3.3 a) Relative frequency of Eukrohnia hamata, Parasagitta elegans and Pseudosagitta

maxima in relation to bottom depth at Amundsen Gulf stations. Multinet and square-conical

net collections included. b) Weighted mean depths of E. hamata and P. elegans normalized

relative to the bottom depth at Amundsen Gulf stations (see ‘Method’). Multinet collections.

Standard deviation bars also given. The blue area indicates the Pacific Halocline (60-200

m) and the red area indicates the Atlantic Layer ˃ 200 m (Geoffroy et al. 2011).

3.5.4 Length distributions and sampler efficiency

Eukrohnia hamata in the Amundsen Gulf ranged in length from 2 to 40 mm (≤34 mm in

other sampling regions) and Parasagitta elegans from 2 to 42 mm (up to 45 mm in other

sampling regions). The length frequency distributions of each species were not significantly

different (Kolmogorov-Smirnov test, P ˃ 0.1) between the two samplers (Figure 3.4). The

few E. hamata and P. elegans ≥ 40 mm were caught solely by the larger-aperture square-

conical net.

Page 76: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

54

Figure 3.4 Length frequency (mean % ± 1 SD) distributions of Eukrohnia hamata and

Parasagitta elegans in the square-conical (S-C) net (1 m2 aperture, 200 µm mesh) and

Multinet (0.5 m2 aperture, 200 µm mesh) in the Amundsen Gulf. Note the different scales for

the two species. k is the number of collections.

3.5.5 Timing of reproduction

The hatching seasons of the two chaetognath species were approximately complementary

over the annual cycle (Figure 3.5). Apart from the occurrence of newborns in October,

Eukrohnia hamata hatched from December to July with the production of newborns ramping

up from low values in December-January to a maximum in April, and stopping after July

(Figure 3.5). Except for their absence in our collections in November and January,

Parasagitta elegans newborns were observed continuously from July to February, with

hatching peaking in December (Figure 3.5).

Page 77: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

55

Figure 3.5 Abundance (mean numbers m-2 + 1 SD) of newborn Eukrohnia hamata and

Parasagitta elegans (body lengths of 2-4 mm) in monthly Multinet deployments from October

to September. August 2007 was inserted between July and September 2008 to provide a

complete composite of the annual cycle. Number of collections shown above bars. The solid

horizontal bar above month labels indicates sampling in the Amundsen Gulf.

3.5.6 Eukrohnia hamata length cohorts and life cycle

Four cohorts were distinguished in all months except August and September 2008, when

there were three (Figure 3.6). The cohort of largest/oldest chaetognaths was labelled E1

(Eukrohnia 1), the intermediate-sized cohorts E2 and E3, and the cohort of autumn (October)

newborns E4. The E1 cohort of large chaetognaths increased little in length from October to

March and disappeared from the collections in April. The length-frequency distributions of

intermediate cohorts E2 and E3 shifted to larger length quasi-monotonically throughout the

annual cycle from October to September. Cohort E2 disappeared from the collections in

August. Initially, the length frequency distribution of cohort E4 progressed slowly to longer

lengths or even regressed as some limited numbers of newborns were added to the cohort

Page 78: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

56

starting in December. A new cohort E5 hatched in April (Figure 3.6). Cohort E5 hardly

increased in average length from April to July, as newly hatched chaetognaths were

constantly added to it, but the width of the adjusted Gaussian distribution increased (declining

kurtosis).

Figure 3.6 Monthly length frequency distributions of Eukrohnia hamata. Frequencies of

newborns highlighted in orange. Visually identified length cohorts shown as normal

distributions (in red) with red dots indicating the mean length. Each of the five cohorts is

Page 79: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

57

labelled with a capital E and a number from oldest (1) to youngest (5). Blue line is the total

distribution obtained by summing the modelled distributions. Chi-square values for the

goodness-of-fit of the total distribution to the data are given. Sampling regions are

abbreviated in each panel: PC, Parry Channel; AG, Amundsen Gulf; BB, Baffin Bay. k is the

number of collections and n the number of length measurements included (see ‘Method’).

A similar pattern of annual growth was observed for cohorts E1 to E5 of Eukrohnia hamata

(Figure 3.7). After a period of latency of 3-4 months, the mean length of cohorts E4 and E5

of newly hatched E. hamata started to increase (Figure 3.7a). Apart from this initial latency

in younger animals, cohorts E2-E4 increased more or less regularly in mean length from

March to June. The mean length of all cohorts declined from January to February and this

decline persisted until March in cohorts E3 and E4 (Figure 3.7a).

The average length of cohort E4 and E5 in September corresponded to the average length of

cohort E2 and E3 in October respectively (Figure 3.7a), strongly suggesting the co-existence

of an autumn brood (October) and a spring brood (April) of Eukrohnia hamata. Connecting

the corresponding cohorts created a composite of the growth trajectory of the assumed

autumn and spring broods (Figure 3.7b). In all months, the standard deviations around

average length of the presumed autumn and spring broods were distinct. The E1 cohort of

large chaetognaths likely represented the final months of life of the spring brood (Figure

3.7b). The growth trajectory of the autumn brood terminated after 22 months with the

disappearance of cohort E2 from the collections. The growth trajectory of the spring brood

indicated a slightly longer life cycle of two years (24 months). Mean length at a given age

did not differ significantly (t-test, t = 43 df = -0.36, P = 0.72) between the autumn and spring

broods (Figure 3.7c). For the autumn brood, growth was significantly faster in the second

year of life than in the first (ANCOVA, P ˂ 0.01). In the spring brood, growth rates were not

significantly different between years one and two (ANCOVA, P ˃ 0.05).

Page 80: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

58

Figure 3.7 a) Monthly mean length (± 1 SD) of the five cohorts of Eukrohnia hamata starting

from a major birth month; b) Composite growth trajectories of the autumn brood (born in

October) and spring brood (born in April) assuming a 2-y lifespan; c) Growth-age curves of

the autumn and spring broods. Circles show mean values of each characteristic (± 1 SD

shown as ribbons).

The autumn and spring broods of Eukrohnia hamata matured in different seasons (Table 3.1)

but at similar body lengths (Figure 3.8). In the autumn brood, all variables peaked in the

second year of life (16 to 23 months of age) from the winter (January-March) to July (Table

Page 81: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

59

3.1). Maturity was somewhat less protracted in the spring brood, with ovary length, width of

seminal vesicles and diameter of the protruding seminal receptacles first peaking in August

of the second year of life at 17 months of age (Table 3.1). Maturity indices and oil vacuole

area peaked at different times from November (age 20 months) to March (age 24 months). In

both broods, oocyte number, ovary length and oil vacuole area started to increase by the end

of the first year of life (Figure 3.8a, Figure 3.8b and Figure 3.8c). Some chaetognaths

presented dispersed sperm in the tail during their first year, but sperm frequency started to

increase systematically around the 21th month of life (Figure 3.8d). Seminal vesicle width,

and sperm receptacle diameter began to develop by the 21th month of life as well (Figure 3.8e

and Figure 3.8f).

Table 3.1 Timing of peak development of maturity features in autumn and spring broods of

Eukrohnia hamata and summer brood of Parasagitta elegans, and of the oil vacuole in E.

hamata. Ages in months at peak development are given in parentheses. Monthly collections

consisted of up to 397 and 434 E. hamata individuals from its autumn and spring broods

respectively, and 177 P. elegans.

Characteristic Month(s) of peak maturity (age in months)

E. hamata

autumn brood

E. hamata

spring brood

P. elegans

summer brood

Number of oocytes Jan-Jul (16-23) Mar (24) Feb (20)

Length of ovaries Mar-Jul (18-23) Aug (17), Jan (22) Jan-Aug (19-26)

Area of oil vacuole Mar-Jul (18-23) Dec-Jan (21-22) -

Fraction of inds. with dispersed sperm Mar-Jul (18-23) Nov (20) Feb-Mar (20-21), Aug (26)

Width of seminal vesicles Feb-Jul (17-23) Aug (17), Feb (23) Aug (14), Aug (26)

Diameter of protruding seminal receptacles Mar-Jul (18-23) Aug (17), Feb (23) Jan (19)

Page 82: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

60

Figure 3.8 Development of sexual features (a, oocyte number; b, ovary length; d, sperm load;

e, seminal vesicle width; f, seminal receptacle diameter) and oil vacuole area (c) with length

for the autumn and spring broods of Eukrohnia hamata. Circles show mean values of each

characteristic (± 1 SD shown as ribbons). Vertical line indicates average length (15.4 mm)

at one year of age. Maturity results for each brood were obtained from the analyses of up to

283 individuals from each 1 mm length class.

Page 83: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

61

A vacuole was observed in 99 % of Eukrohnia hamata (n = 2089), the exception being in

newborns 2-3 mm long from the spring brood. Overall, oil was observed in the vacuole or

the digestive tract of 72 % of the E. hamata examined. The fraction of the population with

oil in the vacuole or the body cavity peaked in February (85 %), declined slowly from

February to April, then faster to a minimum in June (54 %) before peaking again in

September (Figure 3.9). Low percentages were observed in October (72 %) and December

(51 %).

Figure 3.9 Mean frequency (± 1 SD) of Eukrohnia hamata with oil in vacuole and/or digestive

tract by months. Number of water column collections shown inside bar (33 – 446 individuals

per haul).

In the Amundsen Gulf, individuals that rose into the PML during summer were mostly

second-year members of the autumn brood and first-year members of the spring-brood

(Figure 3.10).

Page 84: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

62

Figure 3.10 Abundances (ind. m-3) of Eukrohnia hamata age classes in discrete depth layers

of the Amundsen Gulf, characterized by their mid-points (black squares). Top panels: autumn

brood individuals. Bottom panels: spring brood individuals. Seabed shown in brown,

unsampled sections of the water column in gray. See ‘Method’ for further details.

3.5.7 Parasagitta elegans length cohorts and life cycle

Two cohorts were distinguished in all months except August 2007 and July 2008, when there

were three (Figure 3.11). The cohort of largest/oldest chaetognaths was labelled P1

(Parasagitta 1), the intermediate-sized cohort P2, and the cohort of summer (July) newborns

P3. Cohort P1 (monthly mean lengths >32 mm) disappeared from the collections by

September. The length-distribution of P2 (monthly mean lengths = 21-33 mm) shifted to

larger sizes from July to September and then remained constant until April. That of P3

Page 85: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

63

(monthly mean lengths = 5-19 mm) increased from July to November, remained constant or

shifted to smaller sizes from December to April, and then increased again starting in May

(Figure 3.11). A decrease in the length of cohort P3 in December may not be realistic, because

P. elegans reproduced in December (Figure 3.5). However, any winter cohort produced was

not detected in the size-frequency distributions from subsequent months (Figure 3.11).

Figure 3.11 Monthly length frequency distributions of Parasagitta elegans. Frequencies of

newborns highlighted in orange. Visually identified length cohorts shown as normal

distributions (in red) with red dots indicating the mean length. Each of the three cohorts is

labelled with a capital P and a number from oldest (1) to youngest (3). Blue line is the total

distribution obtained by summing the modelled distributions. Chi-square values for the

Page 86: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

64

goodness-of-fit of the total distribution to the data are given. Sampling regions are

abbreviated in each panel: NF, Nachvak Fjord; PC, Parry Channel; AG, Amundsen Gulf;

BB, Baffin Bay. k is the number of collections and n the number of length measurements

included (see ‘Method’).

The growth trajectories of the three cohorts over the annual cycle indicated relatively fast

growth from April to September/October followed by stagnation over the winter months and

even a slight regression in length from January to April (Figure 3.12a). Assembling the

growth trajectories of the individual cohorts of Parasagitta elegans (Figure 3.12a) revealed

a single summer brood, which terminated at 2.2 years with the disappearance of cohort P1

from the collections (Figure 3.12b).

Figure 3.12 a) Monthly mean length (± 1 SD) of the three cohorts of Parasagitta elegans

starting from a major birth month; b) Composite growth trajectory of the single brood.

Page 87: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

65

Figure 3.13 Development of sexual features (a, oocyte number; b, ovary length; c, sperm

load; d, seminal vesicle width; e, seminal receptacle diameter) with length in Parasagitta

elegans. Circles show mean values of each characteristic (± 1 SD shown as ribbons). Vertical

line indicates average length (20.7 mm) at one year of age. Maturity results were obtained

from the analyses of up to 63 individuals from each 1 mm length class.

In Parasagitta elegans, oocyte number, ovary length, the occurrence of dispersed sperm, and

seminal receptacle diameter first peaked in January or February at ages 19 or 20 months

(Table 3.1). Occurrence of sperm peaked again in August (26 months). The width of seminal

Page 88: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

66

vesicles reached a maximum in August early in the second year of life (14 months) and again

in August at 26 months of age (Table 3.1). As P. elegans increased in length, the presence of

oocytes and perceptible ovaries were the first sign of maturation, both indices starting to

increase at 7 mm (Figure 3.13a and Figure 3.13b). Dispersed sperm was first detected around

8 mm and was present in nearly 100 % of individuals >35 mm (Figure 3.13c). Seminal

vesicles and receptacles started to develop around 17 mm in length (Figure 3.13d and Figure

3.13e).

In the Amundsen Gulf, similar SVM behaviour was detected in first- and second-year

individuals. Based on weighted mean depths, the typically resided in the PH from December

to March, ascended in February, were distributed between the PML and the PH from March

to November, and descended in December (Figure 3.14).

Figure 3.14 Abundances (ind. m-3) of Parasagitta elegans age classes in discrete depth layers

of the Amundsen Gulf, characterized by their mid-points (black squares). Seabed shown in

brown, unsampled sections of the water column in gray. See ‘Method’ for further details.

Page 89: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

67

3.6 Discussion

3.6.1 Chaetognath cohort interpretation and lifespans

Terazaki & Miller (1986) reported three separate broods of Eukrohnia hamata (spring,

summer and autumn) in the north Pacific with lifespans of 8-10 months. Two-year lifespans

were suggested from the existence of two distinct size cohorts in summer in northern Norway

(Sands 1980), and Baffin Bay (Sameoto 1987). In the present study, E. hamata autumn and

spring broods (staggered by 6 months) also had lifespans of 2 years.

Parasagitta elegans lifespans in Arctic waters ranged from one year in a land-locked fjord

(McLaren 1961), to three years in the Barents Sea (Falkenhaug 1993) and Svalbard (Grigor

et al. 2014). Two-year lifespans were suggested in Disko Bay (Dunbar 1940) and Lancaster

Sound (Welch et al. 1996). In this study, a single annual summer-autumn brood had two-year

life cycles in the Amundsen Gulf, Parry Channel and Baffin Bay. In the Amundsen Gulf, a

possible fourth cohort that contained offspring in late December may have quickly perished

in the lower part of the Pacific Halocline. McLaren (1969) found that first-feeding P. elegans

preyed upon Pseudocalanus nauplii. Darnis & Fortier (2014) reported that Pseudocalanus

spp. CI resided relatively deeper in the PH/AL in December compared to January.

Admittedly, cohort interpretation for P. elegans could also have been hampered by relatively

low sample sizes. In Nachvak Fjord (August 2007), the capture of several individuals up to

45 mm long could indicate a slightly longer lifespan there, which could complicate using

these data to fill in missing data from the Amundsen Gulf time series. Differences could be

due to lower growth rates, genetic differences or different feeding strategies between

populations (Pearre 1991, Øresland 1995). The relatively shallow waters of Nachvak Fjord

also contained low numbers of Eukrohnia hamata, which could reduce competition for

resources.

3.6.2 Resource partitioning in the sympatric Eukrohnia hamata and Parasagitta elegans

In arctic seas where many planktonic species depend on a single brief annual bout of primary

production in summer, morphologically similar species can co-occur thanks to subtle

differences in the ecological niches occupied. For instance, in the Beaufort Sea the

morphologically identical larvae and juveniles of polar cod (Boreogadus saida) and ice cod

Page 90: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

68

(Arctogadus glacialis) share the same spatio-temporal distribution, hatching season and

growth rate, but a slightly larger size at hatch results in A. glacialis being longer and feeding

on different prey at a given date than B. saida (Bouchard et al. 2016). As well, the herbivorous

calanoid copepod Calanus glacialis is morphologically almost identical to its sympatric

congener C. hyperboreus except for a smaller size starting at the copepodite IV

developmental stage. Within their sympatric distribution, C. glacialis is preferentially

distributed over the shelf while C. hyperboreus concentrates over the slope and deeper basin

(Darnis et al. 2008 and references therein). The two species also differ in their vertical

distribution and ontogenetic migration (Darnis & Fortier 2014).

Our results support the notion that the morphologically similar and sympatric chaetognaths

Eukrohnia hamata and Parasagitta elegans also partition resources in arctic seas. First,

although the two species generally overlapped in time and space, observations confirm the

more pelagic distribution of E. hamata relative to the neritic P. elegans (Figure 3.3) as

reported in previous studies (Hopcroft et al. 2005, Kosobokova & Hopcroft 2009). Darnis et

al. (2008) identified three zooplankton assemblages in southeastern Beaufort Sea: the Shelf

(43-182 m depth range), Polynya (250-537 m) and Slope (435-1080 m) assemblages. The

pelagic E. hamata thus associated with the Polynya assemblage dominated by Calanus

hyperboreus and the Slope assemblage where omnivores and carnivores are abundant (Darnis

et al. 2008). The neritic P. elegans preferentially associated with the Shelf assemblage, which

is strongly dominated by the neritic copepods Pseudocalanus spp. (Darnis et al. 2008).

Interestingly, McLaren (1969) found that first-feeding P. elegans preyed upon

Pseudocalanus nauplii.

Second, the two species produced offspring in mostly opposite seasons, Eukrohnia hamata

mainly from February to July with a second brood in the autumn (October, December); and

Parasagitta elegans from July to February (Figure 3.5). Consistent with our observations, E.

hamata breeds mostly in spring and summer in the waters off west Greenland and northern

Norway, with reports of continuous low reproduction rates during autumn and winter in west

Greenland and Gerlache Strait, Antarctica (Kramp 1939, Sands 1980, Øresland 1995).

Eukrohnia young are retained in the folded lateral fins of the adults and released at meso-

Page 91: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

69

bathy-pelagic depths (Alvarino 1968, Kruse 2009, Terazaki et al. 2013). Emergence at meso-

pelagic levels in the Amundsen Gulf suggests that they may feed on sinking particles,

protozoans, and copepod nauplii present at depth (Makabe et al. 2016), and gut content

observations in 2007 and 2008 in the same region suggest that E. hamata ingested substantial

amounts of large, amorphous macroaggregates (marine snow) throughout the year, while P.

elegans did not feed in this way (unpublished results).

Summer-autumn reproduction in Parasagitta elegans also agrees with previous observations

in Arctic waters (e.g. Kramp 1939, Dunbar 1940, Grainger 1959, McLaren 1961, Sameoto

1971, Grigor et al. 2014). A similar breeding season from July to February in the Canadian

sub-Arctic led Dunbar (1962) to propose that the breeding season of P. elegans is determined

by temperature rather than food availability. However, the epi-pelagic hatching of the

buoyant eggs of P. elegans (Hagen 1985) matches the emergence of the nauplii and CI of the

copepods Pseudocalanus spp. and Oithona similis in the surface layer from July to January

(Darnis & Fortier 2014), potentially ensuring suitably-sized prey (Saito & Kiørboe 2001) to

newborns for the length of the hatching period. At body lengths ˂ 15 mm, first-year P.

elegans participated in the autumn/winter migrations from the PML to the PH (Figure 3.14),

possibly in pursuit of several species of seasonally migrating copepods (Darnis & Fortier

2014).

We suggest that the generally asynchronous reproduction and depth distributions of

Eukrohnia hamata and Parasagitta elegans may reduce potential competition for food during

the early life of arctic chaetognaths.

3.6.3 The potential role of lipid reserves: contrasting growth in the two species

In Eukrohnia hamata, a large fraction (72 %) of total lipids are in the form of the long-term

energy stores steryl esters and wax (Connelly et al. 2012). The lipids contained in the oil

vacuole could have roles in growth, reproduction and/or buoyancy (e.g. Kruse et al. 2010,

Pond 2012, Grigor et al. 2015), although Øresland (1990) doubted a storage role from their

small size. By comparison, lipids in Parasagitta elegans contain low amounts of wax and

Page 92: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

70

steryl esters (˂1-8 % of total lipids), and short-term lipid stores such as triacylglycerol are

present only in low to moderate amounts (Falk-Petersen et al. 1987, Lee et al. 2006, Connelly

et al. 2012, Connelly et al. 2016). Lipids gained from copepod prey by P. elegans during

spring-summer feeding may be rapidly utilized for reproduction (Choe et al. 2003).

Some studies have reported reduced growth in winter in arctic and subarctic chaetognaths

(Dunbar 1962, Grigor et al. 2014), and others have not (Welch et al. 1996). Slow winter

growth has been attributed to poor feeding as indicated by gut content analyses (e.g. Øresland

1987, Falkenhaug 1991, Grigor et al. 2015). In the present study, both Eukrohnia hamata and

Parasagitta elegans exhibited slower growth in winter, but the annual growth pattern differed

between the two species. In E. hamata growth was variable but positive over ten months from

March to January (1.55 and 1.82 mm mo-1 on average for the autumn and spring broods

respectively), followed by a 1 to 2-month period of negative growth in late winter (January

to February-March, Figure 3.8b). In P. elegans, fast growth from April to September or

November (2.1 mm mo-1 in year 1 and 3.7 mm mo-1 in year 2 of life), was followed by a long

period of winter stagnation from September-November to April (Figure 3.12b).

The migration of some copepod prey to meso-pelagic depths in the autumn could decrease

prey availability for epi-pelagic chaetognaths in winter (Hagen 1999). The seasonal vertical

migrations in Parasagitta elegans suggest that conscious efforts were made to reduce

starvation (Figure 3.3 and Figure 3.14), but if winter food access was a problem, arrested

growth from late autumn to April in this species may reflect the absence of lipid stores and

oil vacuole. On the other hand, the return of migrating copepods to epi-pelagic depths as

early as February (Darnis & Fortier 2014) could decrease prey availability for Eukrohnia

hamata in spring, and this species carried out relatively limited seasonal migrations (Figure

3.3 and Figure 3.10). Sullivan (1980) suggested that E. hamata requires less prey than P.

elegans. The morphology, growth and vertical distribution of E. hamata suggests that long-

term energy stores enable this species to limit starvation at depths ˃ 200 m to a brief interlude

in late winter and spring. However, contradicting this interpretation, the frequency of E.

hamata presenting oil reserves peaked from January to March during the months of negative

growth in length, and started to decline in April-May (Figure 3.9) coinciding with peak

Page 93: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

71

reproduction in April (Figure 3.5). The rebuilding of lipid reserve frequency from June to

September was followed by drops in October and then December (Figure 3.9), which both

coincided with short bouts of reproduction (Figure 3.5). The observed negative correlation

between lipid frequency in the population and offspring production is consistent with the

suggestion by Båmstedt (1978) that part of the lipid content of large specimens of E. hamata

is drawn upon during reproduction, thus lowering the average individual lipid proportion in

the population. We conclude that E. hamata invests its lipid reserves primarily into

reproduction rather than growth, which may contribute to a loss of length in February and

March when prey is likely scarce in deep waters.

3.6.4 Maturation

Dispersed sperm in the tail early in the first year of life was the first sign of maturation in

Eukrohnia hamata, confirming the protandric nature of the species (Terazaki & Miller 1982).

All other maturation traits and the oil vacuole developed in the second year of life. In the

North Pacific, only 3 days separated egg release from the appearance of juveniles (Kotori

1975). Accordingly, the successive maturation episodes in spring and autumn (Table 3.1)

coincided with the occurrence of juvenile E. hamata in net collections. These two maturation

cycles at 6-mo intervals could be interpreted as a sign of iteroparity but actually reflect the

asynchronous maturation of the two broods, which supports the suggestion that E. hamata is

semelparous and dies after a single reproduction effort in the second year of life (Kuhl 1938).

Dunbar (1940) and Grigor et al. (2014) suggested an age-at-maturity of one year for

Parasagitta elegans in the Canadian and European Arctic respectively. Consistent with this,

first-year P. elegans (3-20 mm) could not reproduce as they lacked seminal receptacles and

vesicles. However, oocyte numbers and the frequency of dispersed sperm started to increase

at 9-12 mm corresponding to an age of ca. 6 months. Contrary to Eukrohnia hamata, sexual

maturation occurred from January to August in P. elegans, several months prior to the

emergence of the annual brood from July to December.

Page 94: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

72

3.6.5 Capital versus income breeding in a warming Arctic

Capital breeding based on wax esters accumulated in summer to fuel reproduction in winter

so as to match the emergence of offspring with the spring-summer maximum in biological

production has been documented in several arctic zooplankton taxa (e.g. Kattner et al. 2007

for a review). Arctic cod (Boreogadus saida), a pivotal forage fish in the arctic pelagic food

web, also spawn in winter with the buoyant eggs hatching in the surface layer before the

onset of the spring-summer primary production (e.g. Bouchard & Fortier 2011). In the

present study, Eukrohnia hamata relied on wax ester reserves to fuel reproduction. But

contrary to other arctic capital breeders such as Calanus hyperboreus and B. saida that

produce a single spring brood, E. hamata produced distinct broods in the autumn and the

spring. Much of the reproduction occurred in deep waters (Figure 3.10). The comparison of

growth trajectories and maturity at length shows a similar developmental rate in the two

broods, which suggests that E. hamata is impervious to differences in temperature regime or

food availability between the two seasons. The distribution of E. hamata in the meso-pelagic

(Atlantic) layer where temperature is constant and where copepod prey resides for most of

the year (Darnis & Fortier 2014) may explain the parallel existence of two sub-populations

that hatch, develop, and reproduce separated by a 6-month time-lag.

The alternative strategy to capital breeding is income breeding in which reproduction is

fueled by immediate food availability in spring-summer with the production of offspring

coinciding with peak production in summer. By contrast to Eukrohnia hamata, the epi-

pelagic Parasagitta elegans, which accumulates few lipid reserves, produced a single brood

in summer, typical of an income breeder. This strategy, which increases in frequency towards

lower latitudes (e.g. Kattner et al. 2007), is also used for instance by the Pacific sand lance

(Ammodytes hexapterus) a recent subarctic invader of the Beaufort Sea (Falardeau et al.

2014). The end result of the remarkably different reproduction strategies of the two species

is the staggering of three distinct broods of morphologically and ecologically similar

chaetognaths: the E. hamata spring cohort in April, the P. elegans summer cohort in July,

and the E. hamata autumn cohort in October. The sequestration of lipids thanks to the early

development of an oil vacuole may enable young E. hamata at meso-pelagic depths to

overcome potential food shortages in spring and autumn, while the only survival window for

Page 95: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

73

the lipid-poor P. elegans would be during peak availability of prey in the surface layer in

summer.

3.7 Conclusions

Based on our observations from August 2007 to September 2008 in the Canadian Arctic, the

processes of growth and reproduction are more seasonally-restricted in Parasagitta elegans

compared to Eukrohnia hamata. Length and maturity data for the two species from other

years would help to confirm if this is a general trend. As the Arctic warms and the ice-free

season lengthens, wind mixing increasingly triggers a second phytoplankton bloom in the

autumn, with potentially major impacts on arctic marine ecosystems (Ardyna et al. 2014). If

the availability of suitable prey at the first-feeding stage is critical to the survival of young

chaetognaths, we hypothesize that this autumn bloom would more directly and quickly

benefit the income-breeding P. elegans as copepod prey in epi-pelagic waters would increase.

Page 96: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

74

4. Chapter 4 – Feeding strategies of arctic chaetognaths: are they

really “tigers of the plankton”?

4.1 Résumé

Les chaetognathes, aussi connus sous le nom de « tigres du plancton » jouent un rôle

important dans les communautés mésozooplanctoniques, en termes d'abondance et de

biomasse. Bien que traditionnellement considérés comme étant strictement carnivores, des

études récentes suggèrent l’usage de stratégies non -carnivores au sein du phylum. Les

stratégies d'alimentation des chaetognathes de l’Arctique sont particulièrement intéressants

comptes tenus de la forte saisonnalité dans les abondances et les distributions de proies

connues (copépodes). Le présent chapitre traiter des stratégies d'alimentation saisonnières de

deux principales espèces de l’Arctique : Eukrohnia hamata de la zone mésopélagique, ainsi

que Parasagitta elegans de la zone épipélagique. Le contenu du tube digestif de spécimens

récoltés au printemps, en été et en hiver au sud-est de la mer de Beaufort suggère de faibles

taux de prédation chez toutes les espèces (0-0,27 proie ind. d-1). Toutefois, les taux de

prédation de E. hamata et de P. elegans étaient plus élevés au printemps-été par rapport à

l'automne-hiver. E. hamata mange des macroagrégats ne contenant pas de proies (˃500 µm),

probablement de la neige marine, tout au long de l'année. Ce mode d'alimentation

précédemment non répertorié a été confirmé pour E. hamata par des observations

d'alimentation in-vitro dans la mer de Chukchi à l'automne, mais n’a pas été observé chez P.

elegans. Des quantités étonnamment élevées d’acides gras marqueurs de diatomées en

automne dans les spécimens de E. hamata suggèrent fortement que les diatomées ont été

consommées directement, tandis que des ratios inférieurs à 18:1 (n-9)/(n-7), et valeurs

inférieures de δ15N et δ13C comparativement au P. elegans ont tous suggéré son régime

alimentaire plus omnivore. Nous suggérons que par la consommation de neige marine, les

pelotes fécales du E. hamata pourraient contribuer davantage à la séquestration du carbone

qu'on ne le pensait auparavant.

Page 97: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

75

4.2 Abstract

Chaetognaths, also known as the “tigers of the plankton” are important components of

mesozooplankton communities, in terms of abundance and biomass. Although traditionally

considered to be strict carnivores, recent studies suggest non-carnivorous strategies occur

within the phylum. The feeding strategies of arctic chaetognaths are particularly interesting

given the strong seasonality in the abundance of known food items, chiefly copepods. This

chapter addresses the seasonal feeding strategies of two major Arctic species; the meso-

pelagic Eukrohnia hamata and the epi-pelagic Parasagitta elegans. Gut contents of

specimens collected in spring, summer and winter in the south-eastern Beaufort Sea

suggested low predation rates in the two species (0 – 0.27 prey ind. d-1), but predation rates

in E. hamata and P. elegans were higher in spring-summer compared to autumn-winter. E.

hamata ate non-prey macroaggregates (˃ 500 µm, probably marine snow) throughout the

year. This previously unreported feeding mode was confirmed for E. hamata by in-vitro

feeding observations in the Chukchi Sea in the autumn. High levels of diatom fatty acid

markers in E. hamata sampled in autumn, strongly suggested that diatoms were consumed

directly, whilst lower 18:1 (n-9)/(n-7) ratios, δ15N and δ13C, compared to P. elegans

confirmed its more omnivorous diet. We suggest that by containing algal-rich contents, the

fecal pellets of E. hamata could contribute more to carbon sequestration than those of strict

carnivores.

Page 98: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

76

4.3 Introduction

Chaetognaths, the so-called “tigers of the plankton” (Suthers et al. 2009), are semi-gelatinous

zooplankters considered to be primary carnivores in the marine food web (Reeve 1970).

Reported prey includes fish larvae, copepods and other chaetognaths (e.g. Alvarino 1965,

Sullivan 1980, Brodeur & Terazaki 1999). In some locations, chaetognath populations could

control copepod standing stocks (e.g. Sameoto 1972, Williams & Collins 1985). The fecal

pellets of chaetognaths are said to be large and carbon-rich, and, considering the high

abundances of chaetognaths in the ocean, these pellets could be important food for

zooplankton in the water column or in the flux of carbon to the benthos (Dilling & Alldredge

1993, Giesecke et al. 2010).

The presence of algal cells (Alvarino 1965, Alvarez-Cadena 1993, Marazzo et al. 1997, Kruse

et al. 2010), green detritus and phaeophytin (Philp 2007, Grigor et al. 2015) in chaetognath

guts hints at non-carnivorous feeding. However, such vegetal remains are generally

interpreted as originating in the guts of digested animal prey. Based on various aspects of

chaetognath nutrition, notably in-vitro ingestion of seawater and active digestive processes

in gut cells in the absence of visible gut content, Casanova et al. (2012) suggested that

dissolved and fine particulate organic matter were major food sources for chaetognaths.

Several studies on the polar chaetognaths Eukrohnia hamata, Parasagitta elegans and

Pseudosagitta maxima have reported on ˃90 % of individuals lacking visible gut prey (e.g.

Sameoto 1987, Øresland 1995, Froneman & Pakhomov 1998, Brodeur & Terazaki 1999,

Bollen 2011, Giesecke & Gonzalez 2012, Grigor et al. 2015). For many polar zooplankters,

the ability to switch between carnivorous and non-carnivorous feeding (e.g. detritivory or

filtering) is an important adaptation to seasonal cycles in food availability (e.g. Auel et al.

2002, Hirche et al. 2003, Norkko et al. 2007, Søreide et al. 2008). Could the polar

chaetognaths utilize similar seasonal adaptations, even though they do not obviously exhibit

adaptations to filter phytoplankton?

Gut content analysis provides information on recent feeding but the resulting interpretation

of trophic relationships can be flawed if sampling or handling protocols damage the gut,

Page 99: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

77

induce regurgitation or defecation, or if feeding is biased by the concentration of predator

and prey in the sampling gear (Baier & Purcell 1997 and references therein). Additional and

likely less-biased longer-term information on trophic dynamics can be gained from the

analysis of fatty acids and stable isotopes. Some fatty acids specific to a given prey typically

persist in predators for weeks to months with little or no breakdown or transformation, and

can be useful trophic markers (Dalsgaard et al. 2003, Arim & Naya 2003, Grigor et al. 2015).

Nitrogen isotopes are routinely used to infer trophic levels due to a step-wise increase in

15N/14N ratio (δ15N) with every ascending trophic level (Minagawa & Wada 1984, Michener

& Schell 1994). Carbon isotopes can be used to infer the source of primary productivity,

given close similarities in the 13C/12C ratios (δ13C) of an animal and its food (Hobson &

Welch 1992).

This study aims at understanding how food resources are utilized year-round by co-existing

Eukrohnia hamata and Parasagitta elegans in the Canadian and Alaskan Arctic. In the south-

eastern Beaufort Sea, gut contents are examined to reveal diets from November 2007 to

August 2008. In addition, we compare the fatty acid and stable isotope signatures of E.

hamata and P. elegans in the north-eastern Chukchi Sea (Alaskan Arctic) and Baffin Bay

(Canadian Arctic), based on collections made in autumn 2014.

4.4 Method

4.4.1 Study areas

Sampling surveys were carried out on-board CCGS Amundsen between from 2007 to autumn

2008 in the Beaufort Sea, and in autumn 2014 in the Chukchi Sea and Baffin Bay (Figure

4.1). 127 stations were sampled at weekly (or higher) resolution, in the Amundsen Gulf (69-

72°N, 120-131°W) between November 2007 and August 2008 (Figure 4.1a and Appendix E-

1). 6 stations were sampled in the north-eastern Chukchi Sea (71-76°N, 144-168°W) in

September 2014 (Figure 4.1b and Appendix E-2). 3 stations were sampled in the Baffin Bay

region (67-72°N, 61-73°W) in early October 2014; single stations in Scott Inlet Fjord, in

Gibbs Fjord and southern Baffin Bay (Figure 4.1c and Appendix E-2).

Page 100: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

78

The Amundsen Gulf is a bridge between the south-eastern part of the Beaufort Sea to the

Canadian archipelago. This 400-km long gulf has a width of 170 km and a maximum depth

~630 m, and sea-ice cover between October and early June (Barber & Hanesiak 2004,

Geoffroy et al. 2011). Three water masses prevail in the region; the nutrient-poor Pacific

Mixed Layer (PML, 0-50 m depth; salinity ˂ 31.6 psu), the Pacific Halocline (PH, 50–200 m

depth; 32.4-33.1 psu), and lastly the Atlantic Layer (AL, ˃200 m depth; ˃34 psu) (Carmack

& Macdonald 2002). The Chukchi Sea (Alaska’s northernmost shelf sea) is a relatively

shallow environment characterised by hotspots of high primary productivity stimulated by

the northward invasion of nutrient-rich Pacific water (Hopcroft et al. 2004, Weingartner et

al. 2005). Sinking phytoplankton may only be partly exploited by zooplankton, supporting

benthic animals (Hopcroft et al. 2004 and references therein). On the eastern side of the

Canadian archipelago lies the semi-enclosed basin of Baffin Bay, with numerous fjords along

the east coast of Baffin Island. A unique hydrocarbon seep and chemolithic community

occurs in Scott Inlet Fjord (DFO 2015), but the plankton communities in these fjords are

poorly studied.

Page 101: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

79

Figure 4.1 Bathymetric maps of the Arctic Ocean, showing the positions of sampling stations

(black circles) in (a) Amundsen Gulf, (b) north-eastern Chukchi Sea and (c) Baffin Bay (SIF

= Scott Inlet Fjord; GF = Gibbs Fjord; SBB = southern Baffin Bay). Details of sampling

stations are shown in Appendices E-1 and E-2.

4.4.2 Sampling in the Amundsen Gulf

The biomass of chlorophyll a (˃0.7 µm) was used as an indicator of algae blooming. At 54

stations (Appendix E-1), seawater samples were collected from 7-12 depths at using the CTD

rosette equipped with twenty-four 12 1 Niskin-type bottles (OceanTest Equipment). Methods

for the determination of chlorophyll a (chl a) from seawater subsamples are outlined in

Chapter 3 of this thesis.

Page 102: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

80

At 10 stations, zooplankton were sampled using a large square-conical net with a 1 m2

opening area, 200 µm mesh and a 2-L rigid cod-end. At 72 stations, zooplankton were

sampled using a Hydrobios® Multinet, comprising 9 nets with 0.25 m2 apertures, 200 µm

meshes and 2-L rigid cod-ends. Individual nets open and close sequentially to sample

zooplankton from pre-selected depth intervals (Appendix E-1). Samplers were deployed to a

depth of 10 m above the seabed, cod-end(s) first (non-filtering), and then hauled back to the

surface at a constant velocity of 0.5 m s-1. For Multinet sampling, the upper 60 m and lower

60 m of the sampled water column were divided into three 20-m depth layers, and the

remainder was divided into three equal layers. Samples were fixed in 4 % buffered

formaldehyde-seawater solution.

4.4.3 Sampling in the Chukchi Sea and Baffin Bay

Square-conical nets (1 m2 opening areas, 2-L rigid cod-ends) were deployed to collect

chaetognaths for fatty acid and stable isotope analyses (Appendix E-2). At 7 stations, epi-

pelagic chaetognaths were sampled by oblique trawls of 500/750 µm-mesh nets at 90 m. The

ship speed was approximately 1 m s-1, and the cable angle was ~60º. At 6 stations, meso-

pelagic chaetognaths were sampled in vertical tows of 200/500 µm-mesh nets (see Amundsen

Gulf sampling protocol). A maximum of 30 Eukrohnia hamata or Parasagitta elegans

individuals were randomly removed from each sample, and kept frozen at -80 ºC for further

processing.

4.4.4 Abundance of zooplankton

A total of 70 Multinet collections in the Amundsen Gulf (typically 9 samples per haul), 3 to

11 per month from November 2007 to July 2008, were analysed to estimate the abundance

of main zooplankton species. Formalin-preserved samples were sieved into two size

fractions; 0.2-1 mm and >1 mm. From each fraction, known subsamples were taken until 300

copepods were removed, and all animals in the subsamples were counted and identified to

the lowest possible taxonomic level, typically species or stage. Abundances of species

present in the water column were calculated (where possible), summing abundances of

appropriate taxa in the two size fractions.

Page 103: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

81

4.4.5 Gut contents

A total of 21 net collections in the Amundsen Gulf were used to analyze chaetognath gut

contents. 4742 chaetognaths were included in gut content analyses (4078 Eukrohnia hamata

with body lengths between 2 and 39 mm and 654 Parasagitta elegans with body lengths

between 2 and 42 mm). 10 Pseudosagitta maxima also present in the samples (15-65 mm)

were not analysed further. Each individual chaetognath was stained with a solution of Borax

Carmine to highlight potential prey tissues (Sameoto 1987). Gut contents were observed

under the stereomicroscope (maximum magnification = 11.5×). Prey items were identified to

the lowest taxonomic level possible, although accurate identifications of prey species were

often difficult due to their augmented state of digestion. Prey items in the mouths of

chaetognaths were ignored, as they are a likely reflection of the concentration of predator and

prey in the sampling gear, and therefore biased. The average number of prey per chaetognath

gut (npc) was estimated for each collection for the two species. Daily predation rates (DPR;

number of prey items consumed ind.-1 d-1) were calculated after Bajkov (1935) [Eq. (4.1)].

DPR =npc x 24

tdig (4.1)

Where tdig = digestion time in hours at a suitable temperature (here ~0 ºC). For Eukrohnia

hamata, we used 11 h, determined for this species in the Southern Ocean (Giesecke &

Gonzalez 2012). For Parasagitta elegans, we used 10.2 h, determined for this species in

Massachusetts (Feigenbaum 1982).

To detect the presence of different food items, such as algae, the color of lipid droplets and

detritus in the guts of 2118 Eukrohnia hamata and 201 Parasagitta elegans was also

described. A limited subset of individuals (3 E. hamata and 3 P. elegans from single

collections in November, February, May and July) were re-analysed using Scanning Electron

Microscopy (SEM), to detect other evidence of recent feeding only visible at these high

magnifications.

Page 104: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

82

4.4.6 Fatty acids

To determine the fatty acid profiles of Eukrohnia hamata in autumn (2014), we analysed

samples of 5 pooled individuals from 1 station in the Chukchi Sea and 3 in the Baffin Bay

region. We also analysed samples of 5 pooled Parasagitta elegans individuals from 3 stations

in the Chukchi Sea and 2 in the Baffin Bay region (Appendix E-2). For E. hamata, three

samples were analysed per station. For P. elegans, 1-3 samples were analysed per station.

Samples were freeze-dried and an internal standard (5β cholanic acid, 1 µg) was added. Total

lipids were extracted using a mixture of dichloromethane and methanol (2/1; 5 ml; 3×15

minute ultrasonications). The total lipid fraction was dried with N2 and MeOH-H2O KOH

(80/20; 5 %; 3 ml) and was then added to the extract before heating at 90 ºC for 2 hours. The

fraction containing the non-saponifiable lipids was obtained by liquid-liquid extraction, dried

over Na2SO4, and derivatised using BSTFA (70 ºC, 30 minutes) prior to gas

chromatography–mass spectrometry (GC-MS) analysis. The saponification mixture was

acidified with HCl (10N; 2 ml) and the fatty acids extracted with hexane. Fatty acids were

methylated (BF3MeOH; 20/80; 80 ºC; 30 minutes) prior to identification and quantification

of 20 fatty acid methyl esters previously reported in Grigor et al. (2015), plus 20:1 (n-7) by

GC-MS. Results are given as relative percentages of the various fatty acids identified in

specimens at each station. A higher 18:1 (n-9)/(n-7) ratio was used to indicate a greater

tendency towards carnivory (Falk-Petersen et al. 1990, Wang et al. 2015). Higher proportions

of the monounsaturated FAs Σ20:1+22:1 MUFA were taken to indicate a greater contribution

of Calanus copepods in the diet (Falk-Petersen et al. 1987, Wang et al. 2015). A higher ratio

of 16:1/16:0 was used to indicate a greater contribution of diatoms compared to flagellates

(Nelson et al. 2001, Wang et al. 2015).

4.4.7 Carbon and nitrogen

To determine the stable isotope signatures of Eukrohnia hamata in autumn (2014), we

analysed individuals from 3 stations each in the Chukchi Sea and Baffin Bay region. We also

analysed Parasagitta elegans samples from 4 stations in the Chukchi Sea and 2 in the Baffin

Bay region (Appendix E-2). These samples each contained up to 5 individuals. For E.

hamata, 3-6 samples were analysed per station. For P. elegans, 1-5 samples were analysed

per station. Tin-wrapped samples were combusted at a localized temperature up to 1800 °C

Page 105: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

83

using an ECS 4010 Elemental Analyser/ZeroBlank Autosampler (Costech Analytical

Technologies). Carbon and nitrogen masses were determined to the nearest microgram.

Gases produced (N2 and CO2) after oxidation/reduction and water removal were separated

using an internal GC column. Isotope ratios were measured by on-line continuous-flow

isotope ratio mass spectrometry (IRMS) with a Thermo Electron Delta Advantage

spectrometer operating the continuous-flow mode (Thermo Electron ConFlo III). Five of

each of the standards (USGS40 and USGS41, Qi et al. 2003) were analysed at the beginning

and end of each run. One standard was run for every 12 samples to check for combustion and

correct any instrumental drift; isotope ratio errors were ±0.006 or better.

δ13C values were calculated for samples as changes in sample ratios of 13C/12C from those in

the international standard Vienna Pee Dee Belemnite (13C/12C). δ15N values were calculated

as changes in sample ratios of 15N/14N from those in the standard AIR (15N/14N), as in [Eq.

(4.2)].

δ13C or δ15N = [(𝑅sample/𝑅standard) − 1]×1000 (4.2)

Where R = ratio of 13C/12C or 15N/14N

Trophic levels (TLs) of samples were calculated as changes in δ15N values from that of a

typical food-web baseline (TL = 1 in Eq. 4.3). We used particulate organic matter (POM) as

the baseline, in line with other arctic studies (e.g. Iken et al. 2005, Bergmann et al. 2009, Roy

et al. 2015).

TL = 1 +(δ15Nsample−δ15Nbaseline)

δ15Nenrichment per TL (4.3)

Where 1 = TL of POM. We used a δ15Nbaseline value of 6.8 ‰, previously reported for shallow

water POM (˂50 m) in the North Water Polynya in late spring/summer of 1998 (Hobson et

al. 2002). It was assumed that δ15N values increased by 3.8 ‰ with every ascending trophic

level (Hobson & Welch 1992, Hobson et al. 2002, Connelly et al. 2014).

Page 106: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

84

4.5 Results

4.5.1 Amundsen Gulf

4.5.1.1 Phenology of algae blooms

Based on chlorophyll a biomass, a bloom of ice algae was first detected in late March in the

PML (0-60 m), peaked at these depths in late April and early May (chl a biomass ˃ 5 mg m-

3), and succeeded by a surface bloom of phytoplankton (10 mg chl a m-3). In July,

phytoplankton penetrated the PH (60-200 m), and phytoplankton biomass peaked in mid-July

in the PML (˃ 16 mg chl a m-3; Figure 4.2).

Figure 4.2 Vertical distributions of chlorophyll a biomass (mg m-3) in the upper 200 m of the

water column along the ship track from November 2007 to July 2008 in the Amundsen Gulf.

Black dots indicate sampling depths. Details of sampling stations are shown in Appendix E-

1. Chl a data were provided by Michel Gosselin (Université du Québec à Rimouski).

4.5.1.2 Zooplankton community

A total of 119 zoo- and icthyoplankton species and 57 other taxa not identifiable to the species

level were detected in the Amundsen Gulf collections from November 2007 to July 2008.

Numerically, the mesozooplankton was dominated by the usual guild of copepods (Oithona

similis, Triconia borealis, Metridia longa, Calanus glacialis, Microcalanus pygmaeus,

Microcalanus pusillus, Pseudocalanus elongatus, Pseudocalanus minutus and Calanus

hyperboreus), which were present at all stations (Table 4.1). Eukrohnia hamata was also

Page 107: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

85

present at all stations in the Amundsen Gulf with abundances ranging from 4 m-2 in late

March to 3318 m-2 in late July. Parasagitta elegans was less abundant than E. hamata at all

stations in Amundsen Gulf (0 m-2 in December to 1927 m-2 in June).

Table 4.1 Composition of the

mesozooplankton (30 most

abundant taxa) sampled in the

Amundsen Gulf from November

2007 to July 2008, based on data

from 70 Multinet hauls. *Taxa

could not be identified to species

level.

Species / taxon n hauls

present

Mean number m-2

(± 1 SD)

Date of peak

abundance

Oithona similis 70 30475±14290 27/01/08

Triconia borealis 70 17583±12191 17/04/08

Metridia longa 70 15402±10758 05/03/08

Calanus glacialis 70 10255±6702 27/01/08

Microcalanus pygmaeus 70 4818±3750 27/01/08

Microcalanus pusillus 67 4603±3260 19/04/08

Pseudocalanus elongatus 70 4085±3241 27/01/08

Pseudocalanus minutus 70 3431±2874 05/12/07

Cyclopina sp.* 51 1968±2308 10/06/08

Calanus hyperboreus 70 1962±1565 19/04/08

Radiolarians* 66 1053±2262 27/05/08

Scolecithricella minor 70 798±752 18/04/08

Triconia parila/notopus 18 787±1745 01/05/08

Frittilaria sp.* 18 648±3756 06/05/08

Limacina helicina 66 608±964 01/07/08

Clione limacina 68 606±563 03/03/08

Spinocalanus longicornis 66 540±758 27/01/08

Pseudocalanus acuspes 59 467±1305 27/01/08

Eukrohnia hamata 70 456±565 23/07/08

Microcalanus sp.* 18 381±2957 25/07/08

Pseudocalanus sp.* 27 307±830 01/07/08

Boroecia maxima 69 295±274 19/04/08

Bivalves* 27 282±938 23/11/07

Aetideopsis rostrata 55 277±443 19/04/08

Oikopleura sp. 43 273±660 01/07/08

Dimophyes arctica 50 273±386 05/03/08

Paraeuchaeta glacialis 70 253±519 28/06/08

Aglantha digitale 68 252±308 27/07/08

Gaetanus tenuispinus 57 212±260 18/04/08

Parasagitta elegans 68 173±339 10/06/08

Page 108: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

86

4.5.1.3 Visible prey items and predation rates

Prey organisms were detected in 0.9 % of Eukrohnia hamata and 0.8 % of Parasagitta

elegans in the Amundsen Gulf. Copepods comprised 97 % of total prey items in E. hamata

and all detected prey items in P. elegans. Identifiable prey taxa were Oithona similis, Calanus

spp., a female Pseudocalanus spp. (Figure 4.3), and a chaetognath in one E. hamata

individual. Multiple copepods (2 or 3) were detected in five chaetognaths. Daily predation

rates (DPRs) were in the range 0-0.27 prey ind. d-1 for E. hamata, and 0-0.09 prey ind. d-1 for

P. elegans. In general, predation rates were higher on dates in November and December

compared to dates in spring and summer. Average npc values were generally higher in

chaetognaths from Multinet collections compared to square-conical net collections, but note

that collections from the latter gear were often considerably larger (Figure 4.3).

Figure 4.3 Average number of prey per chaetognath gut (npc) in square-conical (S-C) net

and Multinet collections from November 2007 to August 2008 in the Amundsen Gulf.

Numbers of individuals shown above data points. k is the number of collections. Inset bottom:

photograph of a Parasagitta elegans specimen (12 mm) with a relatively large Pseudocalanus

spp. copepod in the gut (from November 2 2007).

Page 109: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

87

4.5.1.4 Lipid droplets and detritus

Lipid droplets occurred in the body cavities of 72 % of 3408 analysed Eukrohnia hamata.

Amounts were substantial though not quantified. The fraction of the E. hamata population

with oil in the body cavity peaked between mid-January and early-May (74-87 %), thereafter

declining quickly to below 50 % in late-May and early-June. Low percentages were observed

in December (51 %). Droplets were typically yellow in color between late November and

early March (72-90 % of droplets), but from then until late July were typically green in color

(58-92 % of droplets). In addition to occurring in the guts of only 4 % of 497 Parasagitta

elegans, lipid droplets were smaller and less abundant in this species. Whilst detritus (mainly

crustaceous debris) was observed in the guts of 12 % of Parasagitta elegans, 38 % of

Eukrohnia hamata contained detritus in their guts, with 79 % of these containing thick green

macroaggregates ˃ 500 µm (see Figure 4.4). The fraction of the E. hamata population

containing detritus in the gut peaked in late May (71 %). Detritus in E. hamata were typically

green in color on all 11 sampling dates (56-100 %). The fractions of the population containing

detritus and green macroaggregates were lower in early May (25 % and 14 % respectively)

and in early June (28 % and 25 %).

Figure 4.4 Photographs of substantial

amounts of green macroaggregates (˃

500 µm) in the guts of Eukrohnia

hamata, but not in the guts of

Parasagitta elegans. Specimens from

the Amundsen Gulf on May 31 2008

(20-40 m depth), 20-30 mm.

Page 110: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

88

Green-colored lipid droplets and macroaggregates were observed in both first-year and

second-year Eukrohnia hamata individuals. The fraction of the second-year individuals

containing such material was ≥ 50 % on all 11 sampling dates from November to July. The

occurrence of such gut contents in the first-year animals increased steadily throughout winter

as the animals grew. (Table 4.2).

Table 4.2 Frequency of first- and second-year Eukrohnia hamata individuals with green lipid

droplets or macroaggregates in guts. Separate results are presented for the autumn

(October) and spring (April) broods of E. hamata (see Chapter 3), from approximate month

of hatching (no gut content data were available for October).

Month Autumn brood Month Spring brood

n % with

green guts

n % with

green guts

Age 0-1

Apr 15 73 Nov 82 17

Early May 14 100 Dec 49 35

Late May 12 58 Jan 1 100

Jun 33 61 Feb 33 21

Jul 6 67 Mar 106 13

Nov 135 67 Apr 12 8

Dec 97 65 Early May 46 48

Jan 83 88 Late May 14 50

Feb 20 90 Jun 75 45

Mar 100 79 Jul 78 72

Age 1-2

Apr 3 67 Nov 159 49

Early May 14 71 Dec 86 57

Late May 7 71 Jan 51 80

Jun 29 52 Feb 67 73

Jul 16 94 Mar 74 82

Nov 50 56 Apr 4 0

Dec 33 79 Early May 35 31

Jan 4 75 Late May 27 22

Feb 19 100 Jun 104 31

Mar 26 100 Jul 47 40

Page 111: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

89

4.5.1.5 Scanning Electron Microscope observations

Spherical structures resembling bacteria were observed in Eukrohnia hamata guts in samples

from November and May (Figure 4.5). The apparent hooks of a chaetognath, and copepod

mandibles and filaments, were observed in Parasagitta elegans guts in samples from

November and May. Individuals of both species also contained unidentified amalgamated

structures in November, February, May and July.

Figure 4.5 Scanning Electron Microscope photographs of some items in Eukrohnia hamata

guts (left box) and Parasagitta elegans guts (right box), in different months. Examples: hooks

of a chaetognath, and mandibles of copepods (arrow caps). Specimens 20-30 mm.

Page 112: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

90

4.5.2 Chukchi Sea and Baffin Bay

4.5.2.1 In-vitro feeding observations

At station CS1 in the Chukchi Sea (Figure 4.1b), in-vitro ingestion of green macroaggregates

was observed for two living Eukrohnia hamata individuals enclosed in a petri dish (Figure

4.6). One Pseudosagitta maxima individual also observed to consumed marine snow in a

petri dish (not shown). Guts of E. hamata in Chukchi Sea samples contained algal cells

including those of the Ceratium spp. (dinoflagellate) and diatoms (inverted microscopy,

Figure 4.6).

Figure 4.6 Photographs of

live Eukrohnia hamata

individuals consuming

green macroaggregates

in-vitro in the Chukchi

Sea. Specimens 20-30 mm.

Bottom right: An inverted

microscope photograph of

phytoplankton such as

Ceratium spp. in one gut.

4.5.2.2 Lipid amounts

Lipid amounts (expressed as a fraction of sample dry weights) did not vary between

Eukrohnia hamata and Parasagitta elegans (Kruskal-Wallis ANOVA, P = 1.00). In each

species, lipid amounts were also similar between stations (E. hamata: One-way ANOVA, P

= 0.12, P. elegans: Kruskal-Wallis ANOVA, P = 0.35).

Page 113: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

91

4.5.2.3 Fatty acid profiles: species differences

Of the 21 fatty acid methyl esters quantified, proportions of 13 varied between Eukrohnia

hamata and Parasagitta elegans (Kruskal-Wallis ANOVA, P ˂ 0.05, Table 4.3). Not

accounting for station differences, amounts of Calanus markers (ΣC20:1+C22:1) were

statistically higher in E. hamata samples (≤30.29 %) than P. elegans samples (≤13.40 %)

(Kruskal-Wallis ANOVA, P = 0.00, Table 4.3 and Figure 4.7). Diatom markers 16:1 (n-7)

and 20:5 (n-3) were dominant fatty acids in both species (Table 4.3). Amounts of 20:5 (n-3)

were similar in samples of the two species (11.12 % - 25.82 % in E. hamata, 8.40 - 22.76 %

in P. elegans), but 16:1 (n-7) amounts were higher in E. hamata (23.42 - 48.93 %) compared

to P. elegans (12.88 - 20.44 %). Some P. elegans samples contained high amounts of the

flagellate marker 16:0 (≤25.69 %), and in general, amounts of 16:0 were lower in E. hamata

samples (3.38 - 15.06 %). Corresponding 16:1/16:0 ratios were higher in E. hamata (˃3

versus ˂2 in P. elegans) (Kruskal-Wallis ANOVA, P = 0.00, Figure 4.7). The bacterial

marker 18:1 (n-7) occurred in moderate amounts in both species, but the animal marker 18:1

(n-9) occurred in lower amounts in E. hamata (2.48-5.33 %) compared to P. elegans (7.15 -

11.55 %. Corresponding 18:1 (n-9)/(n-7) ratios were higher in P. elegans (˃1 versus ˂1 in E.

hamata) (Kruskal-Wallis ANOVA, P = 0.00, Figure 4.7).

4.5.2.4 Fatty acid profiles: station and regional differences in Eukrohnia hamata and

Parasagitta elegans

For Eukrohnia hamata, proportions of only four fatty acids; 14:0, 18:2(n-6), 18:4(n-3) and

20:1 (n-9) differed between stations (One-way ANOVA, P ≤ 0.05). 18:1 (n-9)/(n-7) ratios

did not differ between regions (P = 0.43) or stations (P = 0.38). Similarly, 16:1/16:0 ratios

did not differ between regions (P = 0.65) or stations (P = 0.29). For Parasagitta elegans,

proportions of all fatty acids were similar between sampling stations (Kruskal-Wallis

ANOVA, P ˃ 0.05). Kruskal-Wallis ANOVAs also showed that P. elegans in the two regions

had different 18:1 (n-9)/(n-7) ratios (P = 0.02), and different 16:1/16:0 ratios (P = 0.07).

These differences, however, were not detected at the station level (P = 0.15 and P = 0.33),

which may be an artefact of low sample sizes (n = 1) at two stations (Table 4.3b).

Page 114: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

92

Table 4.3 Mean sample lipid, fatty acid amounts and fatty acid proportions (± 1 SD) in (a)

Eukrohnia hamata and (b) Parasagitta elegans at different stations (stns) in autumn 2014.

Sampling locations are abbreviated: CS, Chukchi Sea; SIF, Scott Inlet Fjord; GF, Gibbs

Fjord; SBB, southern Baffin Bay. Also: number (n) of samples analysed and chaetognaths

per sample, and chaetognath body lengths in mm. † Significant species differences (P ≤ 0.05).

Fatty acid CS4 SIF GF SBB Overall average

(a) E. hamata

n samples 3 3 3 3 12

n individuals per sample 4-5 5 5 4-5

Body lengths (mm) 10-32 9-32 20-32 10-29

Lipids % DW 41.99±6.57 40.31±6.52 48.68±1.41 37.28±4.08 42.06±6.16

FA mg g-1 ind.-1 24.34±10.39 37.73±18.49 47.60±10.92 20.47±7.30 32.54±15.51

14:0† 1.32±0.63 0.86±0.42 1.19±0.34 2.81±0.66 1.55±0.90

14:1 (n-5)† 0.20±0.06 0.14±0.05 0.22±0.05 0.13±0.06 0.17±0.06

15:0† 0.10±0.02 0.07±0.04 0.09±0.03 0.16±0.06 0.11±0.05

16:0† 5.78±0.95 6.62±4.80 4.96±1.19 11.25±3.30 7.15±3.62

16:1 (n-5)† 0.48±0.07 0.42±0.28 0.70±0.17 0.89±0.37 0.62±0.29

16:1 (n-7)† 26.03±3.40 32.05±6.88 29.53±5.42 36.12±12.66 30.93±7.74

18:0† 0.76±0.22 0.82±0.51 0.50±0.14 1.04±0.22 0.78±0.33

18:1 (n-7) 10.21±5.46 7.22±2.58 10.43±2.93 6.22±0.18 8.52±3.44

18:1 (n-9)† 3.03±0.50 3.62±1.48 3.20±0.57 3.44±0.52 3.32±0.78

18:2 (n-6)† 0.84±0.12 0.54±0.10 0.33±0.28 0.86±0.26 0.64±0.29

18:3 (n-3) 0.33±0.06 0.28±0.16 0.08±0.07 0.22±0.09 0.23±0.13

18:4 (n-3) 1.88±0.33 0.61±0.35 1.01±0.67 1.58±0.39 1.27±0.64

20:1 (n-7) 0.47±0.10 0.72±0.29 0.52±0.17 0.54±0.11 0.56±0.19

20:1 (n-9)† 10.41±2.01 13.51±3.44 14.05±0.59 7.23±0.69 11.30±3.34

20:4 (n-3) 0.10±0.02 0.16±0.12 0.14±0.08 0.10±0.08 0.12±0.08

20:5 (n-3) 18.13±6.80 15.04±3.67 16.41±0.36 15.39±6.79 16.24±4.56

22:1 (n-7) 0.22±0.12 0.21±0.22 0.17±0.13 0.34±0.08 0.24±0.14

22:1 (n-11)† 6.66±2.91 10.35±4.46 9.58±0.60 4.70±1.66 7.82±3.36

22:5 (n-3) 0.53±0.34 0.60±0.54 0.53±0.20 0.35±0.15 0.50±0.31

22:6 (n-3)† 12.20±3.70 5.99±0.64 6.22±2.74 6.32±4.29 7.68±3.83

24:1 (n-9)† 0.31±0.07 0.17±0.06 0.15±0.12 0.30±0.13 0.23±0.11

Page 115: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

93

Table 4.3 (cont.)

Fatty acid CS1 (CS) CS3 (CS) CS4 (CS) SIF SBB Overall average

(b) P. elegans

n samples 2 3 1 1 2 9

n individuals per sample 5 3-5 5 4 5

Body lengths (mm) 20-28 18-34 20-27 16-24 14-27

Lipids (% DW) 29.22±20.91 40.49±2.94 41.33 38.00 58.29±11.39 41.66±13.54

FA mg g-1 ind.-1 25.95±10.80 41.85±2.67 9.78 32.35 50.71±5.49 20.88±10.03

14:0† 2.80±0.03 2.49±0.29 2.32 3.66 3.65±0.40 2.93±0.60

14:1 (n-5)† 0.02±0.00 0.01±0.00 0.01 0.01 0.01±0.00 0.01±0.00

15:0† 0.40±0.01 0.43±0.08 0.54 0.26 0.24±0.03 0.38±0.11

16:0† 13.88±0.53 17.57±7.04 20.89 23.03 16.62±2.67 17.51±4.69

16:1 (n-5)† 3.18±0.39 2.90±0.51 3.55 2.83 2.41±0.42 2.92±0.49

16:1 (n-7)† 14.69±0.49 16.63±2.90 20.44 14.67 14.19±1.84 15.86±2.57

18:0† 1.01±0.06 1.19±0.44 1.66 1.23 1.17±0.02 1.20±0.29

18:1 (n-7) 5.83±0.82 6.45±2.50 9.22 13.67 9.46±1.48 8.09±2.95

18:1 (n-9)† 8.08±0.06 8.67±2.50 11.29 9.42 8.53±0.45 8.88±1.60

18:2 (n-6)† 1.03±0.18 1.55±0.06 1.88 0.78 0.86±0.02 1.23±0.41

18:3 (n-3) 0.33±0.11 0.31±0.14 0.29 0.15 0.14±0.17 0.26±0.13

18:4 (n-3) 1.01±0.16 1.25±0.49 0.78 0.51 0.79±0.23 0.96±0.37

20:1 (n-7) 0.56±0.04 0.59±0.08 0.77 0.45 0.61±0.07 0.59±0.10

20:1 (n-9)† 6.63±0.51 7.19±1.04 6.75 7.73 6.50±1.86 6.92±0.95

20:4 (n-3) 0.10±0.04 0.10±0.05 0.06 0.03 0.10±0.03 0.09±0.04

20:5 (n-3) 17.72±0.16 13.98±5.72 8.40 12.33 19.37±4.80 15.20±4.93

22:1 (n-7) 0.35±0.04 0.21±0.08 0.26 0.13 0.27±0.02 0.25±0.08

22:1 (n-11)† 3.11±0.59 2.70±0.69 1.67 3.22 3.84±1.31 2.99±0.91

22:5 (n-3) 0.51±0.23 0.32±0.20 0.15 0.25 0.64±0.24 0.41±0.23

22:6 (n-3)† 18.18±0.63 15.00±7.97 8.52 5.35 10.16±4.85 12.84±6.24

24:1 (n-9)† 0.59±0.17 0.47±0.13 0.55 0.30 0.46±0.05 0.49±0.13

Page 116: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

94

Figure 4.7 Fatty acid biomarkers in Eukrohnia hamata and Parasagitta elegans at different

stations in autumn 2014. Sampling locations are abbreviated: CS, Chukchi Sea; SIF, Scott

Inlet Fjord; GF, Gibbs Fjord; SBB, southern Baffin Bay. From top panel to bottom panel:

mean ratios of the carnivory biomarker 18:1 (n-9)/(n-7), mean proportions of the Calanus

biomarkers ΣC20:1+C22:1 MUFA, and mean ratios of the algal biomarker 16:1/16:0 (± 1

SD). No bars – no data available.

4.5.2.5 Carbon and nitrogen isotopes

In the Baffin Bay region, δ13C values were similar in the two species (Kruskal-Wallis

ANOVA, P = 0.34). In the Chukchi Sea region, δ13C values in Parasagitta elegans remained

similar as in Baffin Bay, but δ13C values in Eukrohnia hamata dropped significantly below

those of P. elegans (Kruskal-Wallis ANOVA, P = 0.00). Interestingly, the lowest mean δ13C

values in the study (-23.1 ± 0.3 ‰), and the highest mean values in the study (-19.2 ± 0.3

‰), respectively occurred in E. hamata and P. elegans sampled at the same Chukchi Sea

station (Table 4.4 and Figure 4.8). In the Baffin Bay region, mean TLs were 2.5-2.7 for E.

Page 117: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

95

hamata and 2.9-3.1 for P. elegans. In the Chukchi Sea, mean TLs were 3.0-3.3 for E. hamata

and 3.5-3.7 for P. elegans (Table 4.4 and Figure 4.8).

Table 4.4 Mean sample carbon (C) and nitrogen (N) masses, C/N ratios, δ13C and δ15N

values and derived trophic levels (TLs) (± 1 SD) in Eukrohnia hamata and Parasagitta elegans

at different stations in autumn 2014. Sampling locations are abbreviated: CS, Chukchi Sea;

SIF, Scott Inlet Fjord; GF, Gibbs Fjord; SBB, southern Baffin Bay. Also: number (n) of

samples analysed and chaetognaths per sample, as well as body lengths of chaetognaths in

mm.

Station

n samples

(n inds. per

sample)

Body lengths

(mm)

C

(% DW)

N

(% DW) C/N

δ13C

(‰)

δ15N

(‰) TL

Eukrohnia hamata

CS2 6 (4-5) 5-32 46.0±5.8 14.3±1.8 3.2±0.0 -23.1±0.3 14.4±0.5 3.0±0.1

CS4 4 (4-5) 4-32 40.3±5.3 13.0±1.5 3.1±0.1 -22.7±0.4 15.5±0.2 3.3±0.1

CS5 4 (5) 10-32 47.8±0.6 14.9±0.2 3.2±0.0 -22.6±0.3 14.4±0.2 3.0±0.1

SIF 3 (2-5) 9-32 42.0±4.9 13.4±1.3 3.1±0.1 -19.8±0.3 13.4±0.2 2.7±0.1

GF 4 (4-5) 5-31 46.9±1.2 14.8±0.4 3.2±0.0 -20.1±0.3 12.9±0.6 2.6±0.1

SBB 5 (3-5) 4-29 39.6±5.4 12.5±1.6 3.2±0.0 -20.0±0.2 12.6±0.4 2.5±0.1

Parasagitta elegans

CS1 5 (5) 19-35 42.8±1.9 13.7±0.6 3.1±0.1 -19.2±0.3 16.8±0.9 3.6±0.2

CS2 5 (5) 17-32 36.2±10.8 11.5±3.4 3.2±0.0 -19.3±0.2 16.3±0.3 3.5±0.1

CS3 5 (1-5) 6-34 43.9±2.1 14.1±0.6 3.1±0.1 -19.9±0.7 17.1±0.4 3.7±0.1

CS4 4 (1-5) 6-27 45.2±2.1 14.3±0.5 3.2±0.1 -20.0±0.4 16.7±0.7 3.6±0.2

SIF 1 (4) 16-24 30.4 9.7 3.1 -19.8 14.9 3.1

SBB 4 (1-5) 7-27 44.2±3.4 14.1±0.8 3.1±0.1 -19.3±0.8 14.2±0.3 2.9±0.1

Page 118: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

96

Figure 4.8 Mean sample δ13C and δ15N values and corresponding trophic levels (± 1 SD) of

Eukrohnia hamata and Parasagitta elegans. Station IDs shown next to data points. Results

from Chukchi Sea stations inside red oval. Results from Baffin Bay region stations inside

blue oval (locations are abbreviated: SIF, Scott Inlet Fjord; GF, Gibbs Fjord; SBB, southern

Baffin Bay).

4.6 Discussion

4.6.1 Prey items

Kruse et al. (2010) reported a high degree of diet flexibility in Eukrohnia hamata from the

Southern Ocean, detecting copepods, jellyfish, radiolarians, tintinnids, and diatoms in their

guts. However, copepods comprised almost all prey for Parasagitta elegans in the Barents

Sea (Falkenhaug 1991). In this study, except for the meagre evidence of cannibalism in two

chaetognaths (see Figure 4.5), copepods were the only prey for P. elegans and E. hamata in

the Amundsen Gulf. Copepods taken were the same as in earlier studies; Pseudocalanus spp.,

Oithona spp. and Calanus spp. (e.g. McLaren 1969, Øresland 1990, Falkenhaug 1991). Small

Page 119: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

97

chaetognaths could ingest relatively large copepods compared to their head width (Figure

4.3).

4.6.2 Predation rates

Our detections of visible prey in only 0.9 % of 4078 Eukrohnia hamata and 0.8 % of 654

Parasagitta elegans (and the corresponding daily predation rates ˂ 0.3 prey ind.-1 d-1), provide

little evidence that either chaetognath species is an important predator on copepods in the

Amundsen Gulf. For P. elegans daily predation rates in Arctic Ocean populations may be

lower than those at lower latitudes (e.g. 5.50 ind.-1 d-1 in the North Sea; Saito & Kiørboe

2001); this is likely an effect of temperature on digestion rates/gut transit times. However,

little information exists on gut transit times in arctic chaetognaths. Lipids inside or outside

the oil vacuole of 63 % of E. hamata may not necessarily reflect recent prey consumption

(Øresland 1990, Kruse et al. 2010), so excluding these from predation rate estimates is wise.

Sameoto (1987) recorded visible prey in just 0.3 % of 2000 E. hamata during autumn 1983

in Baffin Bay, but also reported that lipid “globules” in 63 % of E. hamata resembled copepod

oil sacs. Interpreting these as evidence of recent prey consumption led Sameoto (1987) to

suggest predation rates of 1.8 prey ind. d-1. Based on this earlier estimate, Welch et al. (1992)

suggested that chaetognaths annually ingest 51 % of the copepod biomass in the Lancaster

Sound Region of the Canadian Arctic (164 of 319 kJ m-2 yr-1). However, we suspect that oil

vacuoles in E. hamata individuals were wrongly identified as copepod oil sacs in the Sameoto

(1987) study, and suggest that the estimates of Sameoto (1987) and Welch et al. (1992) are

erroneous.

Øresland (1995) reported little seasonal change in the low predation rates of Eukrohnia

hamata (0.3-0.7 prey ind.-1 d.-1) between austral summer and winter in the Gerlache Strait,

Antarctica. Differences in the seasonal behaviour of Arctic and Antarctic copepods may

affect the seasonal feeding opportunities for meso-pelagic chaetognaths such as E. hamata

(Chapter 3). Our observation of an apparent decline in winter predation in Parasagitta

elegans is consistent with earlier studies in the European Arctic (Øresland 1987, Falkenhaug

1991, Bollen 2011, Grigor et al. 2015). Several studies on this species also reported reduced

Page 120: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

98

winter growth in arctic and sub-arctic seas (Dunbar 1962, Grigor et al. 2015, Chapter 3 of

this thesis). These observations could suggest a winter decline in predation opportunities,

though individuals >12.5 mm can consume a wide variety of prey size classes > 250 μm

(Saito & Kiørboe 2001), and by migrating from the PML to the PH in winter, it seems that

P. elegans pursued several copepods which showed similar SVM behaviour (see Darnis &

Fortier 2014). Therefore, rather than prey access, changes in energy requirements may be the

main reason for the winter reductions in predation and growth rates. Some studies have

detected diurnal differences in predation rates (e.g. Sullivan 1980, Feigenbaum 1982,

Sameoto 1987, Brodeur & Terazaki 1999), a factor not accounted for in the present study.

For the fragile chaetognaths, an apparent scarcity of prey in guts could also be an artefact of

sampling and observation protocols. Baier & Purcell (1997) suggested prey was lost in 50%

of chaetognaths owing to regurgitation and defecation during net sampling, so a subsequent

study on gut contents of Mediterranean chaetognaths doubled npc estimates to account for

this potential loss (Duro & Saiz 2000). Inference of recent copepod prey from their

undigested mandible remains could also be useful. In a study from the Barents Sea

(Falkenhaug 1991), detections of such mandibles at magnifications of up to 50× suggested

that 35.7 % of Parasagitta elegans individuals had recently consumed small copepod prey.

In this study, Scanning Electron Microscope (SEM) detections of chaetognath hooks, and

copepod mandibles and filaments in the chaetognath guts, confirmed the value of high power

microscopy for chaetognath feeding studies (Figure 4.5).

4.6.3 Omnivory in arctic chaetognaths

Although vegetal remains in guts are generally interpreted as originating in the guts of

digested animal prey, gut contents suggested Eukrohnia hamata is more omnivorous than

Parasagitta elegans. Specifically, the presence of vast amounts of green macroaggregates

and green lipid droplets in E. hamata guts (Figure 4.4), and also the ingestion of green

macroaggregates by living E. hamata individuals (Figure 4.6), are two observations offering

strong evidence for omnivory. The absence of these observations in P. elegans from the same

sampling locations, times and depths, suggest reduced overlap in their feeding strategies.

Page 121: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

99

Particulate organic matter sourced in the photic zone during spring and summer sinks in the

water column as marine snow, collecting mucus, detritus, protists, bacteria, algae, other

organisms and inorganic material (Alldredge & Silver 1988). Marine snow is identified as a

food source for meso-pelagic copepods (e.g. Dagg 1993, Dilling et al. 1998, Sano et al. 2013),

euphausiids (Dilling & Brzezinski 2004), polychaete larvae (Bochdansky & Herndl 1992),

and fish (Larson & Shanks 1996). Small meso-pelagic copepods and protists which colonize

marine snow may be an important food source for feeding zooplankton.

For polar zooplankton, omnivorous and detritivorous feeding strategies may be useful to

buffer against seasonal prey shortages (Norkko et al. 2007). For instance, the arctic krill

species Thysanoessa inermis may similarly be omni-carnivorous to support growth and

reproduction (Søreide et al. 2006). Eukrohnia hamata mostly resided at meso-pelagic depths

in the Amundsen Gulf, except for some individuals which entered epi-pelagic depths, the

Pacific Halocline, in summer (Chapter 3 of this thesis). As well as adults, juveniles ate marine

snow (Table 4.2), suggesting a role of omnivory in stimulating early growth. ROV dives have

documented aggregations of arctic chaetognaths near the seabed (L Fortier, unpublished

results), where dissolved and fine particulate organic matter occurs in high concentrations

(Iken et al. 2005). In this area, E. hamata may sit-and-wait for marine snow to sink down,

occasionally also capturing the passing copepod. Other chaetognaths in the genera

Archeterokrohnia and Heterokrohnia, known to feed on organic matter and bacteria from

sediments (Casanova 1986), are also closely associated with the nepheloid layer.

4.6.4 Fatty acids and stable isotopes confirm gut contents

4.6.4.1 18:1 (n-9)/(n-7) ratios and Calanus copepod FATMs

18:1 (n-9)/(n-7) ratios ˃ 1 indicate that Parasagitta elegans are typically carnivores.

However, the lower 18:1 (n-9)/(n-7) ratios seen in P. elegans in Baffin Bay (Figure 4.7)

compared the Chukchi Sea are a sign that detritus does contribute more to its diet in the

former region. Moreover, other studies on P. elegans (e.g. Øresland 1987, Falkenhaug 1991)

do not plead in favor of evident carnivory. Unlike in P. elegans, 18:1 (n-9)/(n-7) ratios in E.

hamata populations from different regions were consistently low (~0.2), characteristic of

detritivores such as mud-ingesting seastars (e.g. Howell et al. 2003). Admittedly, care should

Page 122: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

100

be taken when interpreting 18:1 (n-9)/(n-7) ratios in animals, since 18:1 (n-9) FA could be

synthesized by some primary producers, not only zooplankton. In addition, 18:1 (n-9)/(n-7)

ratios may fluctuate with lipid content (Dalsgaard et al. 2003). Higher amounts of Σ20:1+22:1

MUFA in E. hamata than P. elegans in Scott Inlet Fjord, could indicate that in certain

environments, like fjords, E. hamata takes Calanus copepods more frequently than P.

elegans, which may focus on smaller taxa (e.g. Falkenhaug 1991). Partitioning of food

resources, including zooplankton prey, could allow for the co-existence of the two

chaetognath species in shallow environments.

4.6.4.2 Algal FATMs

Differences in the amounts of 16:0 and Σ16:1 in Eukrohnia hamata and Parasagitta elegans

represent interesting observations of the study. An earlier study conducted in the

hyperbenthic zone of the Beaufort Sea during autumn 2003 and summer 2004 (Connelly et

al. 2014) reported that E. hamata contained the diatom 16:1 (n-7) in similar proportions (21.4

± 4.9 %) to an omnivorous mysid (33.0 ± 5.3 %); (30.93 ± 7.74 in our study), whilst lower

proportions of 16:1 (n-7) occurred in hyperbenthic P. elegans (15.9 ± 2.6 %); (14.5 ± 2.6 in

our study). 16:1/16:0 ratios in hyperbenthic E. hamata were amongst the highest of all the

zooplankters analysed by Connelly et al. (2014), whilst proportions of 16:1 (n-7) in E. hamata

exceeded those in all analysed copepods. Arguably, the high amounts of the diatom markers

16:1 (n-7) and 20:5 (n-3) in E. hamata in both our study (Table 4.3a), and that of Connelly

et al. (2014), reflect direct consumption of diatoms in seawater, consistent with our in-vitro

ingestion observations (Figure 4.6).

4.6.4.3 δ13C values

Several studies have reported a general enrichment in the δ13C values of zooplankton taxa,

including copepods and chaetognaths, as longitude decreased in the Canadian Arctic (e.g.

Dunton et al. 1989, Saupe et al. 1989, Schell et al. 1989, Hobson & Welch 1992, Hobson et

al. 2002). We detected such δ13C enrichment for Eukrohnia hamata but not for Parasagitta

elegans (Figure 4.8). The reason for this difference is unclear, but according to Σ20:1+22:1

MUFA results, the contribution of Calanus copepods to the diet of E. hamata was greater in

Page 123: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

101

the west than in the east (not the case for P. elegans). This, or other regional differences in

diets may contribute to the observed change in δ13C values in E. hamata.

4.6.4.4 δ15N values and inferred TLs

Consistent with other arctic studies presenting stable isotope data for chaetognaths, our δ15N

observations (Figure 4.8) suggest that Eukrohnia hamata generally feeds at a lower trophic

level (TL) than Parasagitta elegans (Søreide et al. 2006, Pomerleau et al. 2011, Connelly et

al. 2014). Mean trophic level estimates were lowest for E. hamata in the Baffin Bay region

(2.5-2.7), and a full trophic level higher for P. elegans in the Chukchi Sea (3.5-3.7). Pelagic

animals with TLs of 2.4-2.8 are omnivores, and those with TLs of 2.9-3.3 carnivores,

according to one food web model (Søreide et al. 2006). However, in our method to estimate

TLs of chaetognaths, a potential weakness was assigning a constant δ15N value (6.8) to the

POM baseline, at all sampling stations. δ15N estimates for POM (or any baseline) could

change considerably with sampling location and time, as well as differences in the

constituents of the POM, amongst other factors. For comparison, δ15N values for POM in the

Amundsen Gulf between September and October 2011 (˂ 100 m) ranged between 2.2 and

7.2 ‰ (J-É Tremblay, unpublished results).

4.6.4.5 C/N ratios

Interestingly, the presumably lower capacity of Parasagitta elegans to maintain a wax ester

supply (Connelly et al. 2012), did not cause its C/N ratios to differ from those of Eukrohnia

hamata (Table 4.4). Mean C/N ratios for hyperbenthic chaetognaths in the Beaufort Sea

during summer and autumn were 6.6 ± 0.8 for E. hamata and 5.2 ± 0.7 for P. elegans

(Connelly et al. 2012), whereas those of the pelagic chaetognaths in this study were 3.1-3.2.

This is similar to those of pelagic P. elegans in the Barents Sea in early summer (3.2; Ikeda

& Skjoldal 1989). A C/N ratio decrease in the pelagic populations could reflect enhanced

availability of nitrogen above the nepheloid layer (Connelly et al. 2012).

Page 124: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

102

4.6.5 Morphological explanations

Consistent with the above evidence that Eukrohnia hamata feeds at a lower trophic level than

Parasagitta elegans, this species has fewer posterior teeth and fewer hooks than P. elegans,

and lacks anterior teeth which exist in their counterpart (Furnestin 1965, Terazaki 1993).

Whilst P. elegans hooks are unserrated, the hooks in young E. hamata are serrated (serrations

disappear with age). We suspect that these serrated hooks may be particularly useful for

grabbing falling particles, early in the life cycle.

4.7 Concluding remarks and future studies

Owing to the large size and presumably high abundance of chaetognath fecal pellets,

chaetognath pellets could be important in the flux of carbon to the benthos (Dilling &

Alldredge 1993). However, Giesecke et al. (2010) suggested that a zooplankton-based diet

in taxa such as Parasagitta spp., Solidosagitta spp. and Pseudosagitta spp. cause their pellets

to sink at a slower velocity (5-10 ×), compared to those of other zooplankters known to

mainly consume algae. This could certainly limit the export of fecal-pellet carbon to depth,

as the likelihood of full pellet breakdown in the water column would rise. It is possible that

the pellets of omnivorous Eukrohnia hamata would sink faster, and make more of a

contribution to carbon sequestration. As carbon sequestration by zooplankton may be

important in regulating global climate (Turner 2015), this is an interesting avenue for future

studies. Instead of using intrusive nets, underwater traps that rapidly kill invertebrates with

toxic chemicals may help to reduce prey loss, allowing us to obtain a better impression of

recent feeding. Food choice experiments would help to evaluate preferred food items for E.

hamata and other species.

Page 125: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

103

5. Chapter 5 – General Conclusions

5.1 Resource use by arctic chaetognaths

The main purpose of this Ph. D. thesis was to shed light on the ecology of two major

chaetognath species that reside in Arctic waters (Eukrohnia hamata and Parasagitta

elegans), with a particular focus on how the species use and share habitat and food resources,

allowing them to co-exist. The bulk of studies on zooplankton ecology in Arctic waters have

focused on grazing copepods such as Calanus spp., whereas many aspects of chaetognath

ecology have received relatively less attention. The data used in Chapters 2-4 were based on

chaetognath samples collected during five years (2007, 2008, 2012, 2013, 2014) in European,

Canadian and Alaskan portions of the Arctic, including a full annual cycle in the Canadian

Arctic (2007-2008). This enabled the acquisition of detailed information on feeding habits

(Chapters 2 and 4), and growth and reproductive dynamics (Chapter 3), over broad spatial

and temporal scales. Relative to many other mesozooplankton, chaetognaths were abundant

components at all locations and times sampled. The following sections re-cap upon how P.

elegans had the most specialized requirements (its distribution and life cycle appeared to be

closely associated with copepod prey), whereas E. hamata showed greater flexibility for food

options and spawning times. Co-existence of the similar chaetognaths E. hamata and P.

elegans in the Arctic is possibly due to reduced temporal and spatial overlaps in their habitats

and feeding strategies; differences that reduce competition between species (Ross 1986).

5.2 Winter ecology in Svalbard

Insights into the activities of chaetognaths during the polar night are presented in all three

manuscripts. Winter ecology is considered a “new field of science” (Lønne et al. 2014), and

the polar night sampling dimension of all three articles represents a particularly important

step forward in our understanding of zooplankton communities. The first of the three articles

(Chapter 2) added to our understanding of Parasagitta elegans ecology during the Svalbard

polar night. We showed reduced winter activity through a low predation rate and the absence

of reproduction (both contrasting spring-summer reports in the European Arctic). Reduced

predation in winter offers one explanation for the low winter growth rates reported in the

same area (Grigor et al. 2014). This could possibly be due to reduced access to prey (although

Page 126: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

104

P. elegans individuals >12.5 mm can consume many prey size classes > 250 μm, Saito &

Kiørboe 2001), or reduced energy requirements for winter activities, which may be more

likely.

5.3 Life histories, habitats and spawning times

Based on our annual time series in the Canadian Arctic, mostly the Amundsen Gulf, we

predicted similar maximum body sizes (40-45 mm), lifespans (~2 years), and ages at maturity

(1 year+) for Eukrohnia hamata and Parasagitta elegans in this region of the Arctic (Chapter

3). These results suggested that egg laying in these species did not kill chaetognaths, at least

not immediately (contradictory to Kuhl 1938).

Our observations in the Amundsen Gulf showed that the bulk of Eukrohnia hamata

individuals remained in the Atlantic layer (below ~200 m) in all seasons (Chapter 3). E.

hamata bred during all seasons including winter at these meso-pelagic depths, though most

of its spawning occurred in spring. Hatching earlier in the season would give E. hamata a

longer time to feed and develop energy reserves required for the next winter, and this would

lead to higher fitness and a competitive advantage over chaetognaths born later in the season

with little time to store lipids (Varpe et al. 2007). In stark contrast, the bulk of Parasagitta

elegans individuals was restricted to shallower, epi-pelagic depths. P. elegans bred in the

shallower water masses (mostly the Pacific Mixed Layer) later in the season, during summer

and autumn, showing an overlap in both distribution and breeding phenology with specific

copepods (e.g. Calanus glacialis, Darnis & Fortier 2014). Offspring gain because they hatch

near high numbers of suitable copepod prey, ensuring healthy early growth.

I linked the less seasonally restricted spawning cycle of Eukrohnia hamata (compared to

Parasagitta elegans), to it being a capital breeder that can draw upon its considerable supply

of pre-accumulated wax esters (e.g. Lee & Hirota 1973, Falk-Petersen et al. 1987, Donnelly

et al. 1994, Lee et al. 2006, Connelly et al. 2012, Connelly et al. 2016), as its main

reproductive fuel. Capital breeding in E. hamata could also be supported by apparent declines

in lipid amounts in the oil vacuoles when winter breeding occurred (Chapter 3). Based on our

Page 127: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

105

interpretations, concurrent food could be more useful for sustenance or to allow E. hamata

to grow a little in winter (Varpe et al. 2007, Varpe 2012, Sainmont et al. 2014a). Peaks in the

frequency of E. hamata presenting oil reserves during the months of negative growth in

length (from January to March), suggest that reserves were not used to fuel growth.

Parasagitta elegans appears to be an income breeder dependent on high prey availability to

reproduce, perhaps because it lacks the storage mechanisms seen in its counterpart (e.g. Lee

& Hirota 1973, Falk-Petersen et al. 1987, Donnelly et al. 1994, Lee et al. 2006, Connelly et

al. 2012, Connelly et al. 2016). Income breeding in P. elegans could also be supported by the

elevated rates of growth and apparent predation in summer-autumn compared to winter-

spring (Chapters 3 and 4) and by seasonal overlaps in the habitats of P. elegans and small

copepods such as Calanus glacialis and Pseudocalanus spp. (Darnis & Fortier 2014, Chapter

3). Although some seasonal vertical migration was apparent in populations of both

chaetognath species in the Amundsen Gulf, the clearest ascent and descent signals were seen

in P. elegans. Ascensions by many P. elegans in the spring-time may have covered distances

close to 200 m. The downward migration of P. elegans in the water column during autumn,

to depths not much deeper than 200 m, also appeared to follow Calanus copepods which

overwintered here (Darnis & Fortier 2014). These copepods are likely to have fueled P.

elegans maturation in winter-spring.

5.4 Food resources

The two species ate copepods and chaetognaths, however predation rate estimates in both

Chapters 2 (Parasagitta elegans) and 4 (Eukrohnia hamata and Parasagitta elegans)

provided little evidence that either species controlled populations of copepods. Although

proportions of analysed P. elegans individuals with visible prey in guts remained below 5 %

in both feeding chapters, the higher proportions in Svalbard (4 %) compared to the Amundsen

Gulf (˂ 1 %) could be due to differences in digestion times, prey abundances, or body

conditions of the analysed chaetognaths.

Page 128: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

106

In some ways, the feeding strategies of Eukrohnia hamata and Parasagitta elegans differed

markedly. Our observations also documented thick particulate organic matter (POM) as a

food source for E. hamata across much of the year in the Amundsen Gulf (and during our

limited samplings in the Chukchi Sea). The same gut contents were not apparent in P. elegans

over the entire annual cycle in the Amundsen Gulf (2007-2008). Our observations follow

previous assertions that fine POM and dissolved organic matter (DOM) are important food

sources for chaetognaths (Casanova et al. 2012), as well as stable nitrogen results suggesting

E. hamata to be more omnivorous than P. elegans (Søreide et al. 2006, Pomerleau et al. 2011,

Connelly et al. 2014). The most likely source of these macroaggregates (˃500 µm) is falling

marine snow (grabbed or gulped in seawater), or consumption of material on the seabed.

There were high abundances of E. hamata near the seabed in the Amundsen Gulf, possibly

in the nepheloid layer, where some chaetognath species are known to ingest organic matter

from sediments (Casanova 1986). Grazing of algal cells from ice is unlikely given the meso-

pelagic distribution of E. hamata. Copepods attached to particles may be consumed with

marine snow, and both components provide nutritional benefits for a chaetognath.

Morphological characters in E. hamata, such as the absence of anterior teeth (Furnestin

1965), provide additional and indirect evidence for such omnivory. The head and mouthpart

enhancements of P. elegans (more teeth and hooks) are evidence of a more predatory life

cycle. In-vitro observations also showed this feeding mode in Pseudosagitta maxima, and to

investigate this further, the gut contents of more individuals would need to be analysed.

In the Chukchi Sea (autumn sampling), our observations of macroaggregate consumption by

Eukrohnia hamata in-vitro are useful, but limited by sample size. However, these

observations are reflected by E. hamata fatty acid and stable isotope signatures. In particular,

the high amounts of diatom fatty acid markers in this species offers strong evidence that

diatoms were consumed directly. In Parasagitta elegans, flagellate / dinoflagellate markers

were mixed in with relatively lower amounts of diatom markers (also in Svalbard, see

Chapter 2), and could reflect a direct or an indirect source of algae. However, 18:1 (n-9)/(n-

7) ratios, δ15N values, and δ13C values were all higher in P. elegans than in E. hamata

(especially in the Chukchi Sea), again reinforcing our conclusions that P. elegans tends more

towards carnivory.

Page 129: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

107

Marine snow could be a convenient food source for chaetognaths located below the sub-

surface chlorophyll maximum, avoiding the need for considerable migrations in the water

column to obtain high-quality food. Such food could be critical for the early survival of the

≥90 % of Eukrohnia hamata newborns that hatched at meso-pelagic depths in the months of

December, April, June and July. The specialized serrated hooks in young E. hamata could

help them to grab particles (Furnestin 1965). Interestingly, this feature does not remain in the

adults.

5.5 Limitations of the study

5.5.1 Sampling limitations

Chapters 3 and 4 were based mainly on sampling in the Amundsen Gulf, a restricted region

of the Canadian Arctic. However, since the Amundsen Gulf was not sampled during the

autumn, in Chapter 3 we included length data for Eukrohnia hamata and Parasagitta elegans

from other parts of the Canadian Arctic (the Baffin Bay area and Canadian archipelago) to

infer their growth dynamics in autumn. We have concluded that life histories of chaetognaths

may have slightly different in these regions. Ideally, future seasonal studies should aim to

remain in single sampling regions with restricted water mass advection to ensure single

populations are persistently sampled (e.g. Grigor et al. 2014).

5.5.2 Stable isotopes and trophic levels

In Chapters 2 and 4, we extracted lipids to stabilize carbon isotope values. For a given

species, comparisons between the carbon isotope signatures values reported in different

studies can be difficult, as other methods do not always take this additional step (see El-

Sabaawi et al. 2009). Trophic levels were inferred by comparing δ15N values of our samples

with those previously estimated for a food web baseline (we chose POM). We discussed how

differences in the choice of the baseline, and its assigned δ15N value, could alter trophic level

estimates in this and other studies. As for the chaetognaths, δ15N values of any baseline vary

with e.g. location and season. POM is also a mixture of various materials, with a mixture of

δ15N values.

Page 130: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

108

5.5.3 Length cohorts

Seasonal growth dynamics and lifespans of Eukrohnia hamata and Parasagitta elegans were

predicted from body length cohorts, identified by visual means in monthly length frequency

distributions (Chapter 3). We noted that cohorts were visually clearer (and therefore growth

and lifespan interpretations were likely to be better), when species produced a single annual

brood or had restricted breeding seasons. Methods that depend less on visual identifications

of cohorts could help to reduce subjectivity, for verifications of growth rates and lifespan.

5.6 Future research avenues

This study identified important differences in how two different chaetognath species utilize

available resources for co-existence in Arctic waters. Another interesting avenue for further

study could be the investigation of the feeding ecology of the two Heterokrohnia species that

reside at bathy-pelagic depths in the Arctic (Dawson 1968, Kapp 1991). In the future,

analyses of the amounts and composition of lipids in chaetognaths throughout a full annual

cycle (as done here for gut contents) would help to clarify how reserves are utilized.

For gut content analyses of chaetognaths sampled in nature, the use of sampling techniques

that are less obtrusive than nets (sediment traps, light traps and others) can be expected to

reduce the sampling biases that corrupt predation rate estimates (Baier & Purcell 1997). Since

small, rapidly digested copepods (e.g. Oithona spp.) may be key chaetognath prey, guts must

be scanned for small body parts such as copepod mandibles (Falkenhaug 1991). This will

certainly require high power microscopy such as Scanning Electron Microscopy (SEM),

which showed promising results in Chapter 4. Information on gut passage time and real

predation rates may also be obtained by incubating chaetognaths in large containers alongside

different prey, providing they can be kept alive, which can be a challenge. Such experiments

could check if Eukrohnia hamata show selectivity for different sizes and types of marine

snow particles, or if this mode of feeding is mostly opportunistic. E. hamata remain part-time

carnivores, so what is the energetic benefit of a diet based mainly on marine snow versus one

based mainly on zooplankton prey? Is marine snow frequently consumed in other E. hamata

Page 131: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

109

populations? Analyses of fatty acids and stable isotopes throughout a full annual cycle could

also provide insights into diet switching.

The role of arctic chaetognaths in mediating biogeochemical fluxes (their role in the

biological pump) should be re-considered. The fecal pellets of chaetognaths are large,

carbon-rich and produced year-round due to continued feeding. Large fecal pellets tend to

sink faster when loaded with algal contents and marine snow, and may be more likely to

reach seafloor depths than those of strict carnivores. This is especially the case when excreted

at meso-pelagic depths (Dilling & Alldredge 1993, Giesecke et al. 2010), so the fecal pellets

of Eukrohnia hamata could contribute more to carbon sequestration than previously thought.

A similar conclusion was reached by Casanova et al. (2012), based on their suggestions that

chaetognaths make use of the DOM products of viruses and bacteria.

5.7 A note on climate change

Climate change can affect the populations of grazers by influencing the phenology of primary

production and taxonomic composition of primary producers (Søreide et al. 2010, Leu et al.

2011). Our observations suggest that any future changes in copepod abundances, in response

to climate change, would have more severe effects on Parasagitta elegans than Eukrohnia

hamata, this being due to its greater reliance on copepods. Increased primary production in

ice-free waters (Pabi et al. 2008) could benefit E. hamata due to increased availability of

copepods, protozoans or freshly produced marine snow. Specifically, as the ice-free season

lengthens, the autumn brood of E. hamata (which currently comprises smaller numbers than

its main spring brood) may benefit from the increased food availability provided by new

phytoplankton blooms in autumn. On the other hand, P. elegans may also benefit, now

encountering more herbivorous copepod prey at epi-pelagic depths. Changes in chaetognath

abundances or distributions are expected to affect ecosystems through ‘bottom-up’ and/or

‘top-down’ effects.

Page 132: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

110

Bibliography

Alldredge AL, Silver MW (1988) Characteristics, dynamics, and significance of marine

snow. Progress in Oceanography 20:41-82

Alvarez-Cadena JN (1993) Feeding of the chaetognath Sagitta elegans Verrill. Estuarine

Coastal and Shelf Science 36:195-206

Alvarino A (1964) Bathymetric distribution of chaetognaths. Pacific Science 18:64-82

Alvarino A (1965) Chaetognaths. Oceanography and marine biology. An annual review

3:115-194

Alvarino A (1968) Egg pouches and other reproductive structures in pelagic Chaetognatha.

Pacific Science 22:488-492

Alvarino A (1990) 12: Chaetognatha. In: Adiyodi KG, Adiyodi RG (eds) Reproductive

biology of invertebrates: IV: sexual differentiation and behaviour. John Wiley & Sons,

Chichester, pp 255-282

Alvarino A (1992) 22: Chaetognatha. In: Adiyodi KG, Adiyodi RG (eds) Reproductive

biology of invertebrates. V. Sexual differentiation and behaviour. John Wiley & Sons,

Chichester, pp 425-470

Ardyna M, Babin M, Gosselin M, Devred E, Rainville L, Tremblay J-É (2014) Recent Arctic

Ocean sea ice loss triggers novel autumn phytoplankton blooms. Geophysical Research

Letters 41:6207-6212

Arim M, Naya DE (2003) Pinniped diets inferred from scats: analyses of biases in prey

occurrences. Canadian Journal of Zoology 81:67-73

Arnkvaern G, Daase M, Eiane K (2005) Dynamics of coexisting Calanus finmarchicus,

Calanus glacialis and Calanus hyperboreus populations in a high-Arctic fjord. Polar

Biology 28:528-538

Auel H, Harjes M, da Rocha R, Stübing D, Hagen W (2002) Lipid biomarkers indicate

different ecological niches and trophic relationships of the Arctic hyperiid amphipods

Themisto abyssorum and T. libellula. Polar Biology 25:374-383

Baier CT, Purcell JE (1997) Effects of sampling and preservation on apparent feeding by

chaetognaths. Marine Ecology Progress Series 146:37-42

Båmstedt U (1978) Studies on the deep-water pelagic community of Korsfjorden, western

Norway: seasonal variation in weight and biochemical composition of Chiridius armatus

(Copepoda), Boreomysis arctica (Mysidacea), and Eukrohnia hamata (Chaetognatha) in

relation to their biology. Sarsia 63:145-154

Banse (2013) Reflections about chance in my career, and on the top-down regulated world.

Annual Review of Marine Science 5:1-19

Bajkov AD (1935) How to estimate the daily food consumption of fish under natural

conditions. Transactions of the American Fisheries Society 65:288-289

Barber DG, Hanesiak JM (2004) Meteorological forcing of sea ice concentrations in the

southern Beaufort Sea over the period 1979 to 2000. Journal of Geophysical Research.

doi:10.1029/2003JC002027

Barber DG, Asplin MG, Gratton Y, Lukovich J, Galley RJ, Raddatz RL, Leitch D (2010) The

International Polar Tear (IPY) Circumpolar Flaw Lead (CFL) system study: overview

and the physical system. Atmosphere-Ocean 48:225-243

Page 133: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

111

Bates NR, Cai WJ, Mathis JT (2011) The ocean carbon cycle in the western Arctic Ocean:

distributions and air-sea fluxes of carbon dioxide. Oceanography 24:186-201

Bell T, Josenhans H (1997) The seismic record of glaciation in Nachvak Fiord, Northern

Labrador. In: Davies TA, Bell T, Cooper AK, Josenhans H, Polyak L, Solheim A, Stoker

MS, Stravers JA (eds) Glaciated continental margins: an atlas of acoustic images.

Chapman & Hall, London, pp 190-193

Berge J, Varpe Ø, Moline MA, Renaud PE, Daase M, Falk-Petersen S (2012) Retention of

ice-associated amphipods: possible consequences for an ice-free Arctic Ocean. Biology

Letters 8:1012-1015

Bergey MA, Crowder RJ, Shinn GL (1994) Morphology of the male system and

spermatophores of the arrow worm Ferosagitta hispida (Chaetognatha). Journal of

Morphology 221:321-341

Bergmann M, Dannheim J, Bauerfeind E, Klages M (2009) Trophic relationships along a

bathymetric gradient at the deep-sea observatory HAUSGARTEN. Deep-Sea Research

I 56:408-424

Bieri R (1959) The distribution of the planktonic Chaetognatha in the Pacific and their

relationship to the water masses. Limnology and Oceanography 4:1-28

Bieri R, Thuesen EV (1990) The strange worm Bathybelos. American Scientist 78:542-549

Bigelow HB (1926) Plankton of the offshore waters of the Gulf of Maine. Bulletin of the

Bureau of Fisheries 40:1-509

Bird KJ, Charpentier RR, Gautier DL, Houseknecht DW, Klett TR, Pitman JK, Moore TE,

Schenk CJ, Tennyson ME, Wandrey CR (2008) Circum-Arctic resource appraisal –

estimates of undiscovered oil and gas north of the Arctic Circle: U.S. Geological Survey

Fact Sheet 2008-3049, pp 1-4. Available, http://pubs.usgs.gov/fs/2008/3049/, retrieved

1 March 2017

Bochdansky AB, Herndl GJ (1992) Ecology of amorphous aggregations (marine snow) in

the Northern Adriatic Sea. III: zooplankton interactions with marine snow. Marine

Ecology Progress Series 87:135-146

Bogorov BG (1940) On the biology of Euphausiidae and Chaetognatha in the Barents Sea.

Bulletin of Moscow Society of Naturalists 4:3-18

Bogorov V (1946) Zooplankton collected by the Sedov expedition 1937-1939. Trudy Sedov

Expedition 3:336-370

Bollen M (2011) Feeding ecology of the arctic chaetognath Parasagitta elegans. MSc thesis,

University of Bremen

Boltovskoy J (1981) Atlas del zooplancton del Atlantico Sudocddental y mitodos de trabajo

con el zooplancton marino. Publicación Especial del INIDEP, Mar de Plata (in Spanish)

Bone Q, Kapp H, Pierrot-Bults AC (1991) The biology of chaetognaths. Oxford University

Press, New York

Bouchard C, Fortier L (2011) Circum-arctic comparison of the hatching season of polar cod

Boreogadus saida: a test of the freshwater winter refuge hypothesis. Progress in

Oceanography 90:105-116

Bouchard C, Mollard S, Robert D, Suzuki K, Fortier L (2016) Contrasting the early life

histories of sympatric polar cod Boreogadus saida and ice cod Arctogadus glacialis in

southeastern Beaufort Sea. Polar Biology 39:1005-1022

Page 134: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

112

Brodeur RB, Terazaki M (1999) Springtime abundance of chaetognaths in the shelf region

of the northern Gulf of Alaska, with observations on the vertical distribution and feeding

of Sagitta elegans. Fisheries Oceanography 8:93-103

Brotz L, Cheung WL, Kleisner K, Pakhomov E, Pauly D (2012) Increasing jellyfish

populations: trends in large marine ecosystems. Hydrobiologia 690:3-20

Buchanan RA, Sekerak AD (1982) Vertical distribution of zooplankton in Eastern Lancaster

Sound and Western Baffin Bay, July-October 1978. Arctic 35:41-55

Carmack EC, Macdonald RW (2002) Oceanography of the Canadian Shelf of the Beaufort

Sea: a setting for marine life. Arctic 55:29-45

Casanova J-P (1986) Quatre nouveaux chaetognathes atlantiques abyssaux (genre

Heterokrohnia). Description, remarques éthologiques et biogéographiques.

Oceanologica Acta 9:469-477

Casanova J-P, Barthélémy R-M, Duvert M, Faure E (2012) Chaetognaths feed primarily on

dissolved and fine particulate organic matter, not on prey: implications for marine food

webs. Hypotheses in the Life Sciences 2:20-29

Chapman WL, Walsh JE (2003) Observed climate change in the Arctic, updated from

Chapman and Walsh, 1993: Recent variations of sea ice and air temperatures in high

latitudes. Bulletin of the American Meteorological Society 74:33-47

Cheney J (1976) The spatial and temporal abundance patterns of chaetognaths in the western

North Atlantic Ocean. Ph. D. thesis, Woods Hole Oceanographic Institution

Choe N, Deibel D (2000) Seasonal vertical distribution and population dynamics of the

chaetognath Parasagitta elegans in the water column in the hyperbenthic zone of

Conception Bay, Newfoundland. Marine Biology 137:847-856

Choe N, Deibel D, Thompson RJ, Lee SH, Bushell VK (2003) Seasonal variation in the

biochemical composition of the chaetognath Parasagitta elegans from the hyperbenthic

zone of Conception Bay, Newfoundland. Marine Ecology Progress Series 251:191-200

Comeau S, Gorsky G, Jeffree R, Teyssie JL, Gattuso JP (2009) Impact of ocean acidification

on a key Arctic pelagic mollusc (Limacina helicina). Biogeosciences 6 :1877-1882

Connelly TL, Deibel D, Parrish CC (2012) Elemental composition, total lipid content, and

lipid class proportions in zooplankton from the benthic boundary layer of the Beaufort

Sea shelf (Canadian Arctic). Polar Biology 35:941-957

Connelly TL, Deibel D, Parrish CC (2014) Trophic interactions in the benthic boundary layer

of the Beaufort Sea shelf, Arctic Ocean: combining bulk stable isotope and fatty acid

signatures. Progress in Oceanography 120:79-92

Connelly TC, Businski TN, Deibel D, Parrish CC, Trela P (2016) Annual cycle of lipid

content and lipid class composition in zooplankton from the Beaufort Sea shelf,

Canadian Arctic. Canadian Journal of Fisheries and Aquatic Sciences 73:1-12

Conover RJ (1967) Reproductive cycle, early development, and fecundity in laboratory

populations of the copepod Calanus hyperboreus. Crustaceana 13:61-72

Conover RJ (1988) Comparative life histories in the genera Calanus and Neocalanus in high-

latitudes of the northern hemisphere. Hydrobiologia 167:127-142

Conway DVP, Williams R (1986) Seasonal population structure, vertical distribution and

migration of the chaetognath Sagitta elegans in the Celtic Sea. Marine Biology 93:377-

387

Daase M, Eiane K (2007) Mesozooplankton distribution in northern Svalbard waters in

relation to hydrography. Polar Biology 30:969-981

Page 135: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

113

Daase M, Falk-Petersen S, Varpe Ø, Darnis G, Søreide JE, Wold A, Leu L, Berge J, Philippe

B, Fortier L (2013) Timing of reproductive events in the marine copepod Calanus

glacialis: a pan-Arctic perspective. Canadian Journal of Fisheries and Aquatic Sciences

70:871-884

Daase M, Varpe Ø, Falk-Petersen S (2014) Non-consumptive mortality in copepods:

occurrence of Calanus spp. carcasses in the Arctic Ocean during winter. Journal of

Plankton Research 36:129-144

Daase M, Hop H, Falk-Petersen S (2016) Small-scale diel vertical migration of zooplankton

in the High Arctic. Polar Biology 39:1213-1223

Dagg MJ (1993) Sinking particles as a possible source of nutrition for the calanoid copepod

Neocalanus cristatus in the subarctic Pacific Ocean. Deep-Sea Research I 40:1431-1445

Dalsgaard J, St. John M, Gerard K, Muller-Navarra D, Hagen W (2003) Fatty acid trophic

markers in the pelagic marine environment. Advances in Marine Biology 46:227-318

Darby DA, Polyak L, Bauch HA (2006) Past glacial and interglacial conditions in the Arctic

Ocean and marginal seas – a review. Progress in Oceanography 71:129-144

Darnis G, Barber DG, Fortier L (2008) Sea ice and the onshore-offshore gradient in

zooplankton assemblages in early autumn in south-eastern Beaufort Sea. Journal of

Marine Systems 74:994-1011

Darnis G, Robert D, Pomerleau C, Link H, Archambault P, Nelson RJ, Geoffroy M,

Tremblay J-É, Lovejoy C, Ferguson S, Hunt BV, Fortier L (2012) Current state and

trends in Canadian Arctic marine ecosystems: II. Heterotrophic food web, pelagic-

benthic coupling, and biodiversity. Climatic Change 115:179-205

Darnis G, Fortier L (2014) Temperature, food and the seasonal vertical migration of key

arctic copepods in the thermally stratified Amundsen Gulf (Beaufort Sea, Arctic Ocean).

Journal of Plankton Research 36:1092-1108

Dawson JK (1968) Chaetognaths from the Arctic Basin, including the description of a new

species of Heterokrohnia. Bulletin of the Southern California Academy of Science

67:112-124

De Souza CS, Luz JAG, Mafalda PO (2014) Relationship between spatial distribution of

chaetognaths and hydrographic conditions around seamounts and islands of the tropical

southwestern Atlantic. Anais da Academia Brasileira de Ciências 86:1151-1165

DFO (2015) Ecologically and biologically significant areas. Canada's Eastern Arctic

biogeographic region. DFO Canadian Science Advisory Secretariat Science Advisory

Report 2015/049, pp 1-18. Available, http://www.dfo-mpo.gc.ca/csas-

sccs/publications/sar-as/2015/2015_049-eng.pdf, retrieved 1 March 2017

Dilling L, Alldredge AL (1993) Can chaetognath fecal pellets contribute significantly to

carbon flux? Marine Ecology Progress Series 92:51-58

Dilling L, Wilson J, Steinberg DK, Alldredge AL (1998) Feeding by the euphausiid

Euphausia pacifica and the copepod Calanus pacificus on marine snow. Marine Ecology

Progress Series 170:189-201

Dilling L, Brzezinski MA (2004) Quantifying marine snow as a food choice for zooplankton

using stable silicon isotope tracers. Journal of Plankton Research 26:1105-1114

Donnelly J, Torres JJ, Hopkins TL, Lancraft TM (1994) Chemical composition of Antarctic

zooplankton during austral autumn and winter. Polar Biology 14:171-183

Du J (2002) Combined algorithms for fitting finite mixture distributions. M. Sc. thesis,

McMaster University

Page 136: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

114

Dunbar MJ (1940) On the size distribution and breeding cycles of four marine planktonic

animals from the Arctic. Journal of Animal Ecology 9:215-226

Dunbar MJ (1962) The life cycle of Sagitta elegans in Arctic and subarctic seas, and the

modifying effects of hydrographic differences in the environments. Journal of Marine

Research 20:76-91

Dunton KH, Saupe SM, Golikov AN, Schell DM, Schonberg SV (1989) Trophic

relationships and isotopic gradients among arctic and subarctic marine fauna. Marine

Ecology Progress Series 56:89-97

Dupont N, Aksnes DL (2012) Effects of bottom depth and water clarity on the vertical

distribution of Calanus spp. Journal of Plankton Research 34:263-266

Duro A, Saiz E (2000) Distribution and trophic ecology of chaetognaths in the western

Mediterranean in relation to an inshore-offshore gradient. Journal of Plankton Research

22:339-361

El-Sabaawi L, Dower JF, Kainz M, Mazumder A (2009) Characterizing dietary variability

and trophic positions of coastal calanoid copepods: insight from stable isotopes and fatty

acids. Marine Biology 156:225-237

Falardeau M, Robert D, Fortier L (2014) Could the planktonic stages of polar cod and Pacific

sand lance compete for food in the warming Beaufort Sea? ICES Journal of Marine

Science 71:1956-1965

Falkenhaug T (1991) Prey composition and feeding rate of Sagitta elegans var. arctica

(Chaetognatha) in the Barents Sea in early summer. Polar Research 10:487-506

Falkenhaug T (1993) Population structure and metabolism of Sagitta elegans var. arctica in

the Barents Sea. In: Moreno I (ed) Proceedings of the 2nd international workshop of

chaetognaths. Universitat de les Illes Balears, Mallorca, pp 107-119

Falk-Petersen S, Sargent JR, Tande K (1987) Food pathways and life strategy in relation to

the lipid composition of sub-Arctic zooplankton. Polar Biology 8:115-120

Falk-Petersen S, Hopkins CCE, Sargent JR (1990) Trophic relationships in the pelagic, Arctic

food web. In: Barnes M, Gibson RN (eds) Trophic relationships in the marine

environment. Proceedings of 24th European Marine Biology Symposium. Aberdeen

University Press, Aberdeen, pp 315-333

Falk-Petersen S, Timofeev S, Pavlov V, Sargent JR (2007) Climate variability and possible

effects on Arctic food chains. The role of Calanus. In: Ørbæk JB, Tombre T, Kallenborn

R, Hegseth EN, Falk-Petersen S, Hoel AH (eds) Arctic-alpine ecosystems and people in

a changing environment. Springer Verlag, Berlin, pp 147-166

Falk-Petersen S, Mayzaud P, Kattner G, Sargent JR (2009) Lipids and life strategy of Arctic

Calanus. Marine Biology Research 5:18-39

Feigenbaum DL, Reeve MR (1977) Prey detection in the Chaetognatha: response to a

vibrating probe and experimental determination of attack distance in large aquaria.

Limnology and Oceanography 22:1052-1058

Feigenbaum DL (1982) Feeding by the chaetognath, Sagitta elegans, at low temperatures in

Vineyard Sound, Massachusetts. Limnology and Oceanography 27:699-706

Ferguson SH, Higdon J, Chmelnitsky EG (2010) The rise of killer whales as a major Arctic

predator. In: Ferguson S, Loseto L, Mallory M (eds) A little less Arctic: top predators in

the world’s largest northern inland sea, Hudson Bay. Springer, New York, pp 117-136

Folch J, Lees M, Sloane-Stanley GH (1957) A simple method for isolation and purification

of total lipids from animal tissue. Journal of Biological Chemistry 226:497-509

Page 137: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

115

Forest A, Galindo V, Darnis G, Pineault S, Lalande C, Tremblay J-É, Fortier L (2010) Carbon

biomass, elemental ratios (C:N) and stable isotopic composition (δ13C, δ15N) of

dominant calanoid copepods during the winter-to-summer transition in the Amundsen

Gulf (Arctic Ocean) Journal of Plankton Research 33:161-178

Foster L (2011) Chaetognaths (arrow worms). In: Lewbart GA (ed) Invertebrate Medicine:

II. Wiley-Blackwell, Oxford, pp 355-364

Froneman PW, Pakhomov EA (1998) Trophic importance of the chaetognaths Eukrohnia

hamata and Sagitta gazellae in the pelagic system of the Prince Edward Islands

(Southern Ocean). Polar Biology 19:242-249

Froneman PW, Pakhomov EA, Perissinotto R, Meaton V (1998) Feeding and predation

impact of two chaetognath species, Eukrohnia hamata and Sagitta gazellae, in the

vicinity of Marion Island (Southern Ocean). Marine Biology 131:95-101

Furnestin ML (1965) Variations morphologiques des crochets au cours du développement

dans le genre Eukrohnia. Revue des Travaux de l'Institut des Pêches Maritimes 29:275-

284

Gannefors C, Boer M, Kattner G, Graeve M, Eiane K, Gulliksen B, Hop H, Falk-Petersen S

(2005) The Arctic sea butterfly Limacina helicina: lipids and life strategy. Marine

Biology 147:169-177

Gibbons MJ, Stuart V, Verheye HM (1992) Trophic ecology of carnivorous zooplankton in

the Benguela. South African Journal of Marine Science 12:421-437

Geoffroy M, Robert D, Darnis G, Fortier L (2011) The aggregation of polar cod (Boreogadus

saida) in the deep Atlantic layer of ice-covered Amundsen Gulf (Beaufort Sea) in winter.

Polar Biology 34:1959-1971

Giesecke R, Gonzalez HE (2004) Mandible characteristics and allometric relations in

copepods: a reliable method to estimate prey size and composition from mandible

occurrence in predator guts. Revista Chilena de Historia Natural 77:607-616

Giesecke R, Gonzalez HE, Bathmann UV (2010) The role of the chaetognath Sagitta gazellae

in the vertical carbon flux of the Southern Ocean. Polar Biology 33:293-304

Giesecke R, Gonzalez HE (2012) Distribution and feeding of chaetognaths in the epipelagic

zone of the Lazarev Sea (Antarctica) during austral summer. Polar Biology 35:689-703

Goto T, Yoshida M (1981) Oriented light reactions of the arrow worm Sagitta crassa

Tokioka. Biological Bulletin 160:419-431

Goto T, Yoshida M (1985) The mating sequence of the benthic arrow worm Spadella

schizoptera. Biological Bulletin 169:328-333

Gause GF (1934) The struggle for existence. Williams & Wilkins, Baltimore

Graeve M, Albers C, Kattner G (2005) Assimilation and biosynthesis of lipids in Arctic

Calanus species based on 13C feeding experiments with a diatom. Journal of

Experimental Marine Biology and Ecology 317:109-125

Grainger EH (1959) The annual oceanographic cycle at Igloolik in the Canadian Arctic: 1:

the zooplankton and physical and chemical observations. Journal of the Fisheries

Research Board of Canada 16:453-501

Grebmeier JM, Cooper LW, Feder HM, Sirenko BI (2006) Ecosystem dynamics of the

Pacific-influenced Northern Bering and Chukchi Seas in the Amerasian Arctic. Progress

in Oceanography 71:331-361

Page 138: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

116

Grigor JJ, Søreide JE, Varpe Ø (2014) Seasonal ecology and life history strategy of the high-

latitude predatory zooplankter Parasagitta elegans. Marine Ecology Progress Series

499:77-88

Grigor JJ, Marais A, Falk-Petersen S, Varpe Ø (2015) Polar night ecology of a pelagic

predator, the chaetognath Parasagitta elegans. Polar Biology 38:87-98

Hagen W (1985) On distribution and population structure of Antarctic chaetognaths.

Meeresforsch 30:280-291

Hagen W (1999) Reproductive strategies and energetic adaptations of polar zooplankton.

Invertebrate Reproduction & Development 36:25-34

Haney JF (1988) Diel patterns of zooplankton behaviour. Bulletin of Marine Science 49:583-

603

Hannah CG, Dupont F, Dunphy M (2009) Polynyas and tidal currents in the Canadian Arctic

archipelago. Arctic 62:83-95

Hegseth EN, Sundfjord A (2008) Intrusion and blooming of Atlantic phytoplankton species

in the high Arctic. Journal of Marine Systems 74:108-119

Higdon JW, Hauser DDW, Ferguson SH (2012) Killer whales (Orcinus orca) in the Canadian

Arctic: distribution, prey items, group sizes, and seasonality. Marine Mammal Science

28:93-109

Hirche H-J, Niehoff B (1996) Reproduction of the Arctic copepod Calanus hyperboreus in

the Greenland Sea-field and laboratory observations. Polar Biology 16:209-219

Hirche H-J, Fetzer I, Graeve M, Kattner G (2003) Limnocalanus macrurus in the Kara Sea

(Arctic Ocean): an opportunistic copepod as evident from distribution and lipid patterns.

Polar Biology 26:720-726

Hirche H-J, Kosobokova KN (2011) Winter studies on zooplankton in Arctic seas: the

Storfjord (Svalbard) and adjacent ice-covered Barents Sea. Marine Biology 158:2359-

2376

Hirche H-J, Laudien J, Buchholz F (2015) Near-bottom zooplankton aggregations in

Kongsfjorden: implications for pelago-benthic coupling. Polar Biology.

doi:10.1007/s00300-015-1799-4

Hoberg EP, Kutz SJ, Cook JA, Galaktionov K, Haukisalmi V, Henttonen H, Laaksonen S,

Makarikov A, Marcogliese DJ (2013) Arctic biodiversity assessment: status and trends

in Arctic biodiversity. CAFF, Akureyri

Hobson KA, Welch HE (1992) Determination of trophic relationships within a high arctic

marine food web using δ13C and δ15N analysis. Marine Ecology Progress Series 84:9-18

Hobson KA, Fisk A, Karnovsky N, Holst M, Gagnon J-M, Fortier M (2002) A stable isotope

(δ13C, δ15N) model for the North Water food web: implications for evaluating

trophodynamics and the flow of energy and contaminants. Deep-Sea Research II

49:5131-5150

Hop H, Falk-Petersen S, Svendsen H, Kwasniewski S, Pavlov V, Pavlova O, Søreide JE

(2006) Physical and biological characteristics of the pelagic system across Fram Strait

to Kongsfjorden. Progress in Oceanography 71:182-231

Hopcroft RR, Kosobokova KN, Pinchuk AI (2004) Zooplankton community patterns in the

Chukchi Sea during summer 2004. Deep-Sea Research II 57:27-39

Hopcroft RR, Clarke C, Nelson RJ, Raskoff KA (2005) Zooplankton communities of the

Arctic’s Canada Basin: the contribution by smaller taxa. Polar Biology 28:198-206

Page 139: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

117

Hopcroft RR, Bluhm BA, Gradinger RR (2008) Arctic Ocean synthesis: analysis of climate

change impacts in the Chukchi and Beaufort Seas with strategies for future research: II.

North Pacific Research Board, Anchorage, pp 1-184

Howell KL, Pond DW, Billett DSM, Tyler PA (2003) Feeding ecology of deep-sea seastars

(Echinodermata: Asteroidea): a fatty acid biomarker approach. Marine Ecology Progress

Series 255:193-206

Huntington H, Weller G (2005) Ch. 1: An Introduction to the Arctic Climate Impact

Assessment. Arctic Climate Impact Assessment, Cambridge University Press, pp 1-19

Ikeda T (1974) Nutritional ecology of marine zooplankton. Memoirs of the Faculty of

Fisheries Hokkaido University 22:1-97

Ikeda T, Skjoldal HR (1989) Metabolism and elemental composition of zooplankton from

the Barents Sea during early Arctic summer. Marine Biology 100:173-183

Iken K, Bluhm BA, Gradinger R (2005) Food web structure in the high Arctic Canada Basin:

evidence from δ13C and δ15N analysis. Polar Biology 28:238-249

Ingram R, Bacle J, Barber D, Gratton Y, Melling H (2002) An overview of physical processes

in the North Water. Deep-Sea Research Part II 49:4893-4906

IPCC (2014) Summary for Policymakers. In: Edenhofer O, Pichs-Madruga R, Sokona Y,

Farahani E, Kadner S, Seyboth K, Adler A, Baum I, Brunner S, Eickemeier P, Kriemann

B, Savolainen J, Schlömer S, von Stechow C, Zwickel T, Minx JC (eds) Climate change

2014: mitigation of climate change. Contribution of Working Group III to the Fifth

Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge

University Press, Cambridge, United Kingdom and New York, NY, USA

Ji R, Jin M, Varpe Ø (2013) Sea ice phenology and timing of primary production pulses in

the Arctic Ocean. Global Change Biology 19:734-741

Kapp H (1991) Redescription of Heterokrohnia mirabilis Ritter-Zfihony, 1911

(Chaetognatha). Helgolander Meeresunters 45:253-262

Kattner G, Hagen W, Lee RF, Campbell R, Deibel D, Falk-Petersen S, Graeve M, Hansen

BW, Hirche H-J, Jónasdóttir SH, Madsen ML, Mayzaud P, Muller-Navarra D, Nichols

PD, Paffenhöfer G-A, Pond DW, Saito H, Stübing D, Virtue P (2007) Perspectives on

marine zooplankton lipids. Canadian Journal of Fisheries and Aquatic Sciences 64:1628-

1639

King KR (1979) The life history and vertical distribution of the chaetognath, Sagitta elegans,

in Dabob Bay, Washington. Journal of Plankton Research 1:153-167

Kjesbu OS, Bogstad B, Devine JA, Gjøsæter H, Howell D, Ingvaldsen RB, Nash RDM,

Skjæraasen JE (2014) Synergies between climate and management for Atlantic cod

fisheries at high latitudes. Proceedings of the National Academy of Sciences of the

United States of America 111:3478-3483

Kosobokova KN, Hopcroft RR (2009) Diversity and vertical distribution of

mesozooplankton in the Arctic’s Canada Basin. Deep-Sea Research II 57:96-110

Kotori M (1975) Newly-hatched larvae of Sagitta elegans. Bulletin of the Plankton Society

of Japan 21:113-114

Kraft A, Berge J, Varpe Ø, Falk-Petersen S (2013) Feeding in Arctic darkness: mid-winter

diet of the pelagic amphipods Themisto abysorrum and T. libellula. Marine Biology

160:241-248

Kramp PL (1939) The Godthaab Expedition 1928: Chaetognatha. Meddelelser om Grønland

80:1-38

Page 140: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

118

Kruse S (2009) Biology of meso- and bathy-pelagic chaetognaths in the Southern Ocean. Ph.

D. thesis, University of Bremen

Kruse J, Hagen W, Bathmann U (2010) Feeding ecology and energetics of the Antarctic

chaetognaths Eukrohnia hamata, E. bathypelagica and E. bathyantarctica. Marine

Biology 157:2289-2302

Kuhl W (1938) Chaetognatha. In: Bronns HG (ed) Klassen und Ordnungen des Tierreichs:

IV. Akademische Verlagsgesellschaft, Leipzig

Kwok R, Cunningham GF, Wensnahan M, Rigor I, Zwally HJ, Yi D (2009) Thinning and

volume loss of the Arctic Ocean sea ice cover: 2003-2008. Journal of Geophysical

Research Oceans. doi:10.1029/2009JC005312

Laliberté F, Howell SEL, Kushner PJ (2016) Regional variability of a projected sea ice-free

Arctic during the summer months. Geophysical Research Letters 43:256-263

Larson RJ (1986) Water content, organic content, and carbon and nitrogen composition of

medusae from the Northeast Pacific. Journal of Experimental Marine Biology and

Ecology 99:107-120

Larson ET, Shanks AL (1996) Consumption of marine snow by two species of juvenile

mullet and its contribution to their growth. Marine Ecology Progress Series 130:19-28

Lee RF, Hirota J (1973) Wax esters in tropical zooplankton and nekton and the geographical

distribution of wax esters in marine copepods. Limnology and Oceanography 18:227-

239

Lee RF (1974) Lipid composition of the copepod Calanus hyperboreus from the Arctic

Ocean. Changes with depth and season. Marine Biology 26:313-318

Lee RF, Hagen W, Kattner G (2006) Lipid storage in marine zooplankton. Marine Ecology

Progress Series 307:273-306

Legendre L, Ackley SF, Dieckmann GS, Gullicksen B, Horner R, Hoshiai T, Melnikov IA,

Reeburgh WS, Spindler M, Sullivan CW (1992) Ecology of sea ice biota part 2: global

significance. Polar Biology 12:429-444

Legendre P, Legendre LF (2012) Numerical ecology. III. Elsevier

Leu E, Søreide JE, Hessen DO, Falk-Petersen S, Berge J (2011) Consequences of changing

sea-ice cover for primary and secondary producers in the European Arctic shelf seas:

timing, quantity, and quality. Progress in Oceanography 90:18-32

Li WKW, McLaughlin FA, Lovejoy C, Carmack EC (2009) Smallest algae thrive as the

Arctic Ocean freshens. Science 326:539

Linstad H, Bright RM, Strømman AH (2016) Economic savings linked to future Arctic

shipping trade are at odds with climate change mitigation. Transport Policy 45:24-30

Loeng H, Brander K, Carmack E, Denisenko S, Drinkwater K, Hansen B, Kovacs K,

Livingston P, McLaughlin F, Sakshaug E (2005) Ch. 9: Marine Systems. Arctic Climate

Impact Assessment, Cambridge University Press, pp 453-538

Longhurst A, Sameoto D, Herman A (1984) Vertical distribution of Arctic zooplankton in

summer: eastern Canadian archipelago. Journal of Plankton Research 6:137-168

Lønne OJ, Falk-Petersen S, Berge J (2014) Introduction to the special issue on polar night

studies conducted onboard RV Helmer Hanssen in the Svalbard area. Polar Biology

38:1-3

Lovvorn JR, Cooper LW, Brooks ML, De Ruyck CC, Bump JK, Grebmeier JM (2005)

Organic matter pathways to zoo-plankton and benthos under pack ice in late winter and

Page 141: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

119

open water in late summer in the north-central Bering Sea. Marine Ecology Progress

Series 291:135-150

Makabe R, Hattori H, Sampei M, Darnis G, Fortier L, Sasaki H (2016) Can sediment trap-

collected zooplankton be used for ecological studies? Polar Biology.

doi:10.1007/s00300-016-1900-7

Margulis L, Chapman M (2010) Kingdoms and domains: an illustrated guide to the phyla of

life on Earth: IV. Academic Press, Philadelphia

Marazzo A, Machado CF, Nogueira CSR (1997) Notes on feeding of Chaetognatha in

Guanabara Bay, Brazil. Journal of Plankton Research 19:819-828

McConnaughey T, McRoy CP (1979) 13C label identifies eel-grass (Zostera marina) carbon

in an Alaskan estuarine food web. Marine Biology 53:263-269

McLaren IA (1961) The hydrography and zooplankton of Ogac Lake, a landlocked fjord on

Baffin Island. Ph. D. thesis, Yale University

McLaren IA (1969) Population and production ecology of zooplankton in Ogac Lake, a

landlocked fiord on Baffin Island. Journal of the Fisheries Research Board Canada

26:1485-1559

McLaughlin FA, Carmack EC (2010) Deepening of the nutricline and chlorophyll maximum

in the Canada Basin interior, 2003–2009. Geophysical Research Letters.

doi:10.1029/2010GL045459

Mehlum F, Gabrielsen GW (1993) The diet of high-arctic seabirds in coastal and ice-covered,

pelagic areas near the Svalbard archipelago. Polar Research 12:1-20

Michener RH, Schell DM (1994) Stable isotope ratios as tracers in marine aquatic food webs.

In: Lajtha K, Michener R (eds) Stable isotopes in ecology and environmental science.

Blackwell Scientific, Boston, pp 138-157

Minagawa M, Wada E (1984) Stepwise enrichment of 15N along food chains: Further

evidence and the relation between δ15N and animal age. Geochimica et Cosmochimica

Acta 48:1135-1140

Morata N, Seuthe L (2004) Importance of bacteria and protozooplankton for faecal pellet

degradation. Oceanologia 56:565-581

Nelson MM, Mooney BD, Nichols PD, Phleger CF (2001) Lipids of Antarctic Ocean

amphipods: food chain interactions and the occurrence of novel biomarkers. Marine

Chemistry 73:53-64

Newbury TK (1971) Adaptations of chaetognaths to subarctic conditions. Ph. D. thesis,

McGill University

Nilsen F, Cottier F, Skogseth R, Mattsson S (2008) Fjord-shelf exchanges controlled by ice

and brine production: the interannual variation of Atlantic Water in Isfjorden, Svalbard.

Continental Shelf Research 28:1838-1853

Nishino S, Kikuchi T, Fujiwara A, Hirawake T, Aoyama M (2016) Water mass

characteristics and their temporal changes in a biological hotspot in the southern Chukchi

Sea. Biogeosciences 13:2563-2578

Norkko A, Thrush SF, Cummings VJ, Gibbs MM, Andrew NL (2007) Trophic structure of

coastal Antarctic food webs associated with changes in sea ice and food supply. Ecology,

88:2810-2820

Øresland V (1987) Feeding of the chaetognaths Sagitta elegans and S. setosa at different

seasons in Gullmarsfjorden, Sweden. Marine Ecology Progress Series 39:67-79

Page 142: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

120

Øresland V (1990) Feeding and predation impact of the chaetognath Eukrohnia hamata in

Gerlache Strait, Antarctic Peninsula. Marine Ecology Progress Series 63:201-209

Øresland V (1995) Winter population structure and feeding of the chaetognath Eukrohnia

hamata and the copepod Euchaeta antarctica in Gerlache Strait, Antarctic Peninsula.

Marine Ecology Progress Series 119:77-86

Overland JE, Wang M, Walsh JE, Stroeve JC (2014) Future Arctic climate changes:

adaptation and mitigation timescales. Earth's Future 2 68-74

Pabi S, van Dijken GL, Arrigo KR (2008) Primary production in the Arctic Ocean, 1998–

2006. Journal of Geophysical Research Oceans. doi:10.1029/2007JC004578

Parsons TR, Maita Y, Lalli CM (1984) A manual of chemical and biological methods for

seawater analysis. Pergamon Press, Toronto

Pasternak A, Arashkevich E, Tande K, Falkenhaug T (2001) Seasonal changes in feeding,

gonad development and lipid stores in Calanus finmarchicus and C. hyperboreus from

Malangen, northern Norway. Marine Biology 138:1141-1152

Pearre S (1973) Vertical migration and feeding in Sagitta elegans Verrill. Ecology 54:300-

314

Pearre S (1982) Feeding by Chaetognatha: aspects of inter- and intra-specific predation.

Marine Ecology Progress Series 7:33-45

Pearre S (1991) Growth and reproduction. In: Bone Q, Kapp H, Pierrot-Bults AC (eds) The

biology of chaetognaths. Oxford University Press, Oxford, pp 61-75

Peterson CH (2001) The “Exxon Valdez” oil spill in Alaska: acute, indirect and chronic

effects on the ecosystem. Advances in Marine Biology 39:1-103

Peterson BJ, Holmes RM, McClelland JW, Vörösmarty CJ, Lammers RB, Shiklomanov AI,

Shiklomanov IA, Rahmstorf S (2002) Increasing river discharge to the Arctic Ocean.

Science 298:2171-2173

Philp K (2007) Ningaloo Reef as a plankton filter: changes in the size spectrum and

community structure of zooplankton across a fringing reef. Ph. D. Thesis, University of

Western Australia

Pierce EL (1941) The occurrence and breeding of Sagitta elegans Verrill and Sagitta setosa

J. Miller in parts of the Irish Sea. Journal of the Marine Biological Association of the

United Kingdom 25:113-124

Pierrot-Bults AC, Nair VR (1991) Distribution patterns in Chaetognatha. In: Bone Q, Kapp

H, Pierrot-Bults AC (eds) The biology of chaetognaths. Oxford University Press,

Oxford, pp 86-116

Pomerleau C, Winkler G, Sastri AR, Nelson RJ, Vagle S, Lesage V, Ferguson SH (2011)

Spatial patterns in zooplankton communities across the eastern Canadian sub-Arctic and

Arctic waters: insights from stable carbon (δ13C) and nitrogen (δ15N) isotope ratios.

Journal of Plankton Research 33:1779-1792

Pomerleau C, Lesage V, Ferguson SH, Winkler G, Petersen SD, Higdon JW (2012) Prey

assemblage isotopic variability as a tool for assessing diet and the spatial distribution of

bowhead whale Balaena mysticetus foraging in the Canadian eastern Arctic. Marine

Ecology Progress Series 469:161-174

Pond DW (2012) The physical properties of lipids and their role in controlling the distribution

of zooplankton in the oceans. Journal of Plankton Research 34:443-453

Purcell JE (2012) Jellyfish and ctenophore blooms coincide with human proliferations and

environmental perturbations. Annual Review of Marine Science 4:209-235

Page 143: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

121

Qi H, Coplen TB, Geilmann H, Brand WA, Bohlike JK (2003) Two new organic reference

materials for δ13C and δ15N measurements and a new value for the δ13C of NBS 22 oil.

Rapid Communications in Mass Spectrometry 17:2483-2487

R Development Core Team (2008) R: a language and environment for statistical computing.

R Foundation for Statistical Computing, Vienna

Rand KM, Whitehouse A, Logerwell EA, Ahgeak E, Hibpshman R, Parker-Stetter S (2013)

The diets of polar cod (Boreogadus saida) from August 2008 in the US Beaufort Sea.

Polar Biology 36:907-912

Reeve MR (1970) The biology of chaetognaths. I. Quantitative aspects of growth and egg

production in Sagitta hispida. In: Steele JH (ed) Marine foods chains. Oliver and Boyd,

Edinburgh, pp 168-189

Richardson AJ (2008) In hot water: zooplankton and climate change. ICES Journal of Marine

Science 65:279-295

Ross ST (1986) Resource partitioning in fish assemblages: a review of field studies. Copeia

1986:352-388

Roy V, Iken K, Gosselin M, Tremblay J-É, Belanger S, Archambault P (2015) Benthic faunal

assimilation pathways and depth-related changes in food-web structure across the

Canadian Arctic. Deep-Sea Research I 102:55-71

Russell FS (1932) On the biology of Sagitta. The breeding and growth of Sagitta elegans

Verrill in the Plymouth area, 1930–1931. Journal of the Marine Biological Association

of the United Kingdom 18:131-146

Sainmont J, Anderson KH, Varpe Ø, Visser AW (2014a) Capital versus income breeding in

a seasonal environment. The American Naturalist 184:466-476

Sainmont J, Gislason A, Heuschele J, Webster C, Sylvander P, Wang M, Varpe Ø (2014b)

Inter- and intra-specific diurnal habitat selection of zooplankton during the spring bloom

observed by Video Plankton recorder. Marine Biology 161:1931-1941

Saito H, Kiørboe T (2001) Feeding rates in the chaetognath Sagitta elegans: effects of prey

size, prey swimming behaviour and small-scale turbulence. Journal of Plankton

Research 23:1385-1398

Sakshaug E (2003) Primary and secondary production in the Arctic Seas. In: Stein R,

Macdonald RW (eds) The organic carbon cycle in the Arctic Ocean. Springer, Berlin,

pp 57-81

Saloranta TM, Haugen PM (2001) Interannual variability in the hydrography of Atlantic

water northwest of Svalbard. Journal of Geophysical Research 106:13931-13943

Sameoto DD (1971) life history, ecological production, and an empirical mathematical model

of the population of Sagitta elegans in St. Margaret's Bay, Nova Scotia. Journal of the

Fisheries Research Board of Canada 28:971-98

Sameoto DD (1972) Yearly respiration rate and estimated energy budget for Sagitta elegans.

Journal of the Fisheries Research Board of Canada 29:987-996

Sameoto DD (1987) Vertical distribution and ecological significance of chaetognaths in the

Arctic environment of Baffin Bay. Polar Biology 7:3170-328

Sampey A, McKinnon AD, Meekan MG, McCormick MI (2007) Glimpse into guts:

overview of the feeding of larvae of tropical shorefishes. Marine Ecology Progress

Series 339:243-257

Sands NJ (1980) Ecological studies on the deep-water community of Korsfjorden, Western

Norway: population dynamics of the chaetognaths from 1971-1974. Sarsia 65:1-12

Page 144: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

122

Sano M, Maki K, Nishibe Y, Nagata T, Nishida S (2013) Feeding habits of mesopelagic

copepods in Sagami Bay: Insights from integrative analysis. Progress in Oceanography

110:11-26

Saunders MI, Thompson PA, Jeffs AG, Sӓwstrӧm C, Sachlikidis N, Beckley LE, Waite AM

(2012) Fussy feeders: phyllosoma larvae of the Western Rocklobster (Panulirus cygnus)

demonstrate prey preference. PLoS ONE. doi:10.1371/journal.pone.0036580

Saupe SM, Schell DM, Griffiths WB (1989) Carbon-isotope ratio gradients in western arctic

zooplankton. Marine Biology 103:427-432

Schell DM, Barnett BA, Vinette KA (1989) Carbon and nitrogen isotope ratios in

zooplankton of the Bering, Chukchi and Beaufort seas. Marine Ecology Progress Series

162:11-23

Schmid MS, Aubry C, Grigor J, Fortier L (2016) The LOKI underwater imaging system and

an automatic identification model for the detection of zooplankton taxa in the Arctic

Ocean. Methods in Oceanography. doi:10.1016/j.mio.2016.03.003

Schulz J, Barz K, Ayon P, Ludtke A, Zielinski O, Mengedoht D, Hirche H-J (2010) Imaging

of plankton specimens with the Lightframe On-sight Key Species Investigation (LOKI)

System. Journal of the European Optical Society - Rapid Publications 5:1-9

Simo-Matchim AG, Gosselin M, Blais M, Gratton Y, Tremblay J-É (2016) Seasonal

variations of phytoplankton dynamics in Nunatsiavut fjords (Labrador, Canada) and

their relationships with environmental conditions. Journal of Marine Systems 156:56-75

Smith DM (1998) Recent increase in the length of the melt season of perennial Arctic sea

ice. Geophysical Research Letters 25:655-658

Søreide JE, Hop H, Falk-Petersen S, Gulliksen B, Hansen E (2003) Macrozooplankton

communities and environmental variables in the Barents Sea marginal ice zone in late

winter and spring. Marine Ecology Progress Series 263:43-64

Søreide JE, Hop H, Carroll ML, Falk-Petersen S, Hegseth EN (2006) Seasonal food web

structures and sympagic-pelagic coupling in the European Arctic revealed by stable

isotopes and a two-source food web model. Progress in Oceanography 71:59-87

Søreide JE, Falk-Petersen S, Hegseth EN, Hop H, Carroll ML, Hobson KA, Blachowiak-

Samolyk K (2008) Seasonal feeding strategies of Calanus in the high-Arctic Svalbard

region. Deep-Sea Research II 55:2225-2244

Søreide JE, Leu E, Berge J, Graeve M, Falk-Petersen S (2010) Timing of blooms, algal food

quality and Calanus glacialis reproduction and growth in a changing Arctic. Global

Change Biology 16:3154-3163

Spielhagen RF, Werner K, Sørensen SA, Zamelczyk K, Kandiano E, Budeus G, Husum K,

Marchitto TM, Hald M (2011) Enhanced modern hat transfer to the Arctic by warm

Atlantic water. Science 331:450-453

Stein R, Macdonald RW (2004) The organic carbon cycle in the Arctic Ocean. Springer,

Heidelberg

Stroeve J, Serreze M, Drobot S, Gearheard S, Holland M, Maslanik J, Meier W, Scambos T

(2008) Arctic sea ice extent plummets in 2007. Eos, Transactions American Geophysical

Union 89:13-14

Strömberg J-O (1989) Northern Svalbard Waters. In: Rey L, Alexander V (eds) Proceedings

6th Conference of the Comité Arctique International 13-15 May 1985, p 402-426

Page 145: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

123

Stübner E, Søreide JE, Reigstad M, Marquardt M, Blachowiak-Samolyk K (2016) Year-

round meroplankton dynamics in high-Arctic Svalbard. Journal of Plankton Research.

doi:10.1093/plankt/fbv124

Sullivan BK (1980) In situ feeding behavior of Sagitta elegans and Eukrohnia hamata

(Chaetognatha) in relation to the vertical distribution and abundance of prey at Ocean

Station “P”. Limnology and Oceanography 25:317-326

Suthers I, Dawson M, Pitt K, Miskiewicz AG (2009) Coastal and marine zooplankton:

diversity and biology. In Suthers IM, Rissik D (eds) Plankton, a guide to their ecology

and monitoring for water quality. CSIRO Publications, Melbourne, pp 181-222

Svendsen H, Beszczynska-Møller A, Hagen JO, Lefauconnier B, Tverberg V, Gerland S,

Ørbæk JB, Bischof K, Papucci C, Zajaczkowski M, Azzolini R, Bruland O, Wiencke C,

Winther J-G, Dallmann W (2002) The physical environment of Kongsfjorden-

Krossfjorden, an Arctic fjord system in Svalbard. Polar Research 21:133-166

Syvitski JPM, LeBlanc WG, Cranston RE (1990) The flux and preservation or organic carbon

in Baffin Island fjords. Geological Society, London, Special Publications 53:177-199

Terazaki M, Miller CB (1982) Reproduction of meso- and bathypelagic chaetognaths in the

genus Eukrohnia. Marine Biology 71:193-196

Terazaki M (1993) Deep-sea adaptation of the epipelagic chaetognath Sagitta elegans in the

Japan Sea. Marine Ecology Progress Series 98:79-88

Terazaki M (2004) Life history strategy of the chaetognath Sagitta elegans in the world

oceans. Coastal Marine Science 29:1-12

Terazaki M, Miller CB (1986) Life history and vertical distribution of pelagic chaetognaths

at ocean station P in the subarctic Pacific. Deep-Sea Research I 33:323-337

Terazaki M, Takahashi KT, Odate T (2013) Zonal variations in abundance and body length

of chaetognaths in the 140ºE seasonal ice zone during the austral summer of 2001/02.

Polar Science 7:39-47

Thuesen EV, Kogure K (1989) Bacterial production of tetrodotoxin in four species of

Chaetognatha. Biological Bulletin 176:191-194

Thuesen EV, Numachi K, Nemoto T (1993) Genetic variation in the planktonic chaetognaths

Parasagitta elegans and Eukrohnia hamata. Marine Ecology Progress Series 101:243-

251

Timofeev SV (1995) Parasagitta elegans Verrill (chaetognaths) in waters of Spitsbergen

archipelago. Oceanology 34:788-791

Tonnesson K Tiselius P (2005) Diet of the chaetognaths Sagitta setosa and S. elegans in

relation to prey abundance and vertical distribution. Marine Ecology Progress Series

299:177-190

Tremblay J-É, Gagnon J (2009) The effects of irradiance and nutrient supply on the

productivity of Arctic waters: a perspective on climate change. In: Nihoul JCJ,

Kostianoy AG (eds) Influence of climate change on the changing Arctic and Sub-Arctic

conditions. Elsevier, USA, pp 73-93

Tremblay J-É, Robert D, Varela DE, Lovejoy C, Darnis G, Nelson RJ, Sastri AR (2012)

Current state and trends in Canadian Arctic marine ecosystems I: Primary production.

Climatic Change 115:161-178

Turner JT (2015) Zooplankton fecal pellets, marine snow, phytodetritus and the ocean’s

biological pump. Progress in Oceanography 130:205-248

Page 146: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

124

Ussing HM (1939) Biology of some important plankton animals in fjords of East Greenland.

Meddelelser om Grønland 100:1-108

van Dongen BE, Schouten S, Sinninghe Damsté JS (2002) Carbon isotope variability in

monosaccharides and lipids of aquatic algae and terrestrial plants. Marine Ecology

Progress Series 232:83-92

Varpe Ø (2012) Fitness and phenology: annual routines and zooplankton adaptations to

seasonal cycles. Journal of Plankton Research 34:267-276

Varpe Ø, Jorgensen C, Tarling GA, Fiksen Ø (2007) Early is better: seasonal egg fitness and

timing of reproduction in a zooplankton life-history model. Oikos 116:1331-1342

Varpe Ø, Jørgensen C, Tarling GA, Fiksen Ø (2009) The adaptive value of energy storage

and capital breeding in seasonal environments. Oikos 118:363-370

Walcott CD (1911) Cambrian geology and paleontology II: Middle Cambrian annelids.

Smithson M. Sc. College Bulletin 57:109-144

Wallace MI, Cottier FR, Berge J, Tarling GA, Griffiths C, Brierley AS (2010) Comparison

of zooplankton vertical migration in an ice-free and a seasonally ice-covered Arctic

fjord: an insight into the influence of sea ice cover on zooplankton behavior. Limnology

& Oceanography 55:831-845

Wang SW, Budge SM, Iken K, Gradinger RR, Springer AM, Wooller MJ (2015) Importance

of sympagic production to Bering Sea zooplankton as revealed from fatty acid carbon

stable isotope analyses 518:31-50

Wassmann P, Hansen L, Andreassen I, Wexels Riser C, Urban-Rich J (1999) Distribution

and sedimentation of faecal pellets on the Nordvestbanken shelf, northern Norway, in

1994. Sarsia 84:239-252

Wassmann P, Reigstad M, Haug T, Rudels B, Carroll ML, Hop H, Gabrielsen GW, Falk

Petersen S, Denisenko SG, Arashkevich E, Slagstad D, Pavlova O (2006) Food webs

and carbon flux in the Barents Sea. Progress in Oceanography 71:232-287

Webster C, Varpe Ø, Falk-Petersen S, Berge J, Stübner E, Brierley A (2013) Moonlit

swimming: vertical distributions of macrozooplankton and nekton during the polar night.

Polar Biology. doi:10.1007/s00300-013-1422-5

Weingartner TJ, Aagard, Woodgate R, Danielson S, Sasaki Y, Cavalieri D (2005) Circulation

on the north-central Chukchi Sea shelf. Deep-Sea Research II 52:315-317

Welch HE, Bergmann MA, Siferd TD, Martin KA, Curtis MF, Crawford RE, Conover RJ,

Hop H (1992) Energy flow through the marine ecosystem of the Lancaster Sound region,

Arctic Canada. Arctic 45:343-357

Welch HE, Siferd TD, Bruecker P (1996) Population densities, growth, and respiration of the

chaetognath Parasagitta elegans in the Canadian high Arctic. Canadian Journal of

Fisheries and Aquatic Science 53:520-527 Williams R, Collins NR (1985) Chaetognaths and ctenophores in the holoplankton of the

Bristol Channel. Marine Biology 85:97-107

Winsor P, Bjork G (2000) Polynya activity in the Arctic Ocean from 1958 to 1997. Journal

of Geophysical Research 105:8789-8803

Wold A, Darnis G, Søreide JE, Leu E, Phillipe B, Fortier L, Poulin M, Kattner G, Graeve M,

Falk-Petersen S (2011a) Life strategy and diet of Calanus glacialis during the winter–

spring transition in Amundsen Gulf, south-eastern Beaufort Sea. Polar Biology 34:1929-

1946

Page 147: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

125

Wold A, Jæger I, Hop H, Gabrielsen GW, Falk-Petersen S (2011b) Arctic seabird food chains

explored by fatty acid composition and stable isotopes in Kongsfjorden, Svalbard. Polar

Biology 34:1147-1155

Zavolokin AV, Glebov II, Kosenok NS (2008) Distribution, quantitative composition, and

feeding of jellyfish in the Western Bering Sea in summer and autumn. Russian Journal

of Marine Biology 34:461-467

Zenkevich L (1963) Biology of the seas of the USSR. Georg Allen and Unwin, London

Zo Z (1973) Breeding and growth of the chaetognath Sagitta elegans in Bedford Basin.

Limnology and Oceanography 18:750-756

Page 148: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

126

Appendix A.

Information on the zooplankton samples included in Chapter 2. Sampling date (UTC), station name, latitude, longitude and bottom depth,

sampling gear (MIK or MPS), sampling time (UTC), sampling depths (m). Sampling was horizontal (MIK) in cases where the sampling

depth is given by one number, and vertical (MPS) when given by a lower and an upper depth. The ‘Analyses performed’ column indicates if

haul individuals were analysed for gut contents (GC), data on abundance, maturity, as well as for stable isotope and fatty acid composition.

Each X indicates that 75 Parasagitta elegans from the haul were analysed for stable isotopes (3 x 25 individuals; see Table 2.2), while each

O indicates that 30 individuals were analysed for fatty acid trophic markers (3 x 10 individuals; see Table 2.3). The last column gives numbers

and body lengths (mm) of individuals dissected for gut content analyses, where relevant.

Date

(UTC) Station

Latitude

(°N)

Longitude

(°E)

Bottom depth

(m) Gear

Time

(UTC)

Sampling depth(s)

(m) Analyses performed

n P. elegans dissected :

length range (mm)

12/01/2012 Rijpfjorden (R3) 80.19 22.16 282 MIK 2353 20 GC, XO, abundance 10 : 18-38.5

-"- -"- 80.19 22.11 211 MIK 1146 75 GC, XO, abundance, maturity 40 : 17-55

-"- -"- 80.19 22.15 280 MIK 1220 225 GC, XXO, abundance, maturity 30 : 21-44.5

-"- -"- 80.19 22.16 284 MPS 1310 260-200 Abundance

-"- -"- 80.19 22.16 284 MPS 1310 200-100 Abundance

-"- -"- 80.19 22.16 284 MPS 1310 100-50 Abundance

-"- -"- 80.19 22.16 284 MPS 1310 50-20 Abundance

-"- -"- 80.19 22.16 284 MPS 1310 20-0 Abundance

-"- -"- 80.19 22.16 284 MPS 1350 256-224 GC 10 : 22.5-33

-"- -"- 80.19 22.16 284 MPS 1350 192-160 GC 6 : 24-31.5

-"- -"- 80.19 22.16 284 MPS 1350 160-128 GC 3 : 21.5-29.5

-"- -"- 80.19 22.16 284 MPS 1350 128-0 GC 10 : 15-41

13/01/2012 Rijpfjorden (R3) 80.19 22.16 278 MIK 1104 20 Abundance

-"- -"- 80.19 22.16 275 MIK 0029 75 GC, XO, abundance 11 : 22-44

-"- -"- 80.19 22.16 278 MIK 1136 75 Abundance only

-"- -"- 80.19 22.16 278 MIK 1218 225 O, abundance

-"- -"- 80.19 22.16 281 MPS 0105 260-200 Abundance

-"- -"- 80.19 22.16 281 MPS 0105 200-100 Abundance

-"- -"- 80.19 22.16 281 MPS 0105 100-50 Abundance

-"- -"- 80.19 22.16 281 MPS 0105 50-20 Abundance

-"- -"- 80.19 22.16 281 MPS 0105 20-0 Abundance

-"- -"- 80.19 22.16 281 MPS 0157 256-224 GC 8 : 18-32

-"- -"- 80.19 22.16 281 MPS 0157 224-192 GC 4 : 23-28

Page 149: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

127

-"- -"- 80.19 22.16 281 MPS 0157 192-128 GC 10 : 20-31.5

-"- -"- 80.19 22.16 281 MPS 0157 128-20 GC 10 : 17-34.5

-"- -"- 80.19 22.16 284 MPS 0157 20-0 GC 1 : 10

17/01/2012 Mouth of Isfjorden 78.20 15.50 268 MIK 1957 250 GC, XO, maturity 20 : 20.5-42

17/01/2012 Isfjorden ISA 78.16 15.32 98 MIK 2208 35 GC, XO, maturity 10 : 18-46

-"- -"- 78.16 15.33 98 MIK 2238 60 GC, XO, maturity 10 : 20-46.5

18/01/2012 Isfjorden ISA 78.16 15.33 80 MIK 1142 30 GC, XO, maturity

-"- -"- 78.16 15.33 80 MIK 1207 60 GC, maturity 10 : 25-47

27/01/2012 Isfjorden ISA 78.16 15.33 80 MPS 1717 25-0 Abundance only

-"- -"- 78.16 15.33 80 MPS 1727 65-25 Abundance only

12/01/2013 Rijpfjorden (R3) 80.18 22.17 267 MIK 1249 20 GC, maturity 100 : 14-40

16/01/2013 Kongsfjorden 78.58 11.53 327 MIK 0952 30 GC 100 : 8-30

-"- -"- 78.58 11.52 309 MIK 1145 20 GC 100 : 16-41

-"- -"- 78.58 11.53 303 MIK 1255 20 GC, maturity 100 : 17-36

-"- -"- 78.58 11.50 298 MIK 1322 100 GC, maturity 100 : 12-42

-"- -"- 78.58 11.54 308 MIK 1634 20 GC 100 : 15-40

17/01/2013 Kongsfjorden 79.03 11.27 280 MIK 0549 20 GC 100 : 16-41

Page 150: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

128

Appendix B.

Vertical profiles of temperature and salinity in Chapter 2. a) Rijpfjorden on 13th January 2012; b) the mouth of Isfjorden on 17th January

2012; and c) Isfjorden station ISA on 18th January 2012. Profiles were obtained by CTD downcast.

Page 151: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

129

Appendix C.

Stations sampled in Chapter 3 for information on chaetognath body size and life history characteristics; dates, regions (NF = Nachvak Fjord, PC

= Parry Channel, BB = Baffin Bay, AG = Amundsen Gulf), positions, bottom depths, sampler used (Mn = Multinet, SCn = square-conical net),

sampling strata and sample splits analysed.

Station

ID

Date

(dd.mm.yy) Region Position

Bottom

depth (m)

Sampler

used Sampling strata (m)

Sample

split

600/602 02.08.07 NF 59° 5.25' N, 63° 26.20' W 215 Mn 0-50, 50-100, 100-150, 150-195 1

302 07.10.07 PC 74° 9.30' N, 86° 17.64' W 529 Mn 0-20, 20-40, 40-60, 61-160, 160-300, 300-460, 460-480, 480-500 0.5

308 09.10.07 PC 74° 8.19' N, 103° 7.30' W 346 SCn 0-335 1

1600 02.11.07 AG 71° 40.77'N, 130° 44.21' W 452 SCn 0-358 1

437 23.11.07 AG 71° 43.81' N, 126° 41.37' W 440 Mn 3-21, 21-40, 41-60, 61-161, 161-261, 261-360, 361-380, 381-396, 397-412 1

D3 30.11.07 AG 71° 2.39' N, 123° 58.46' W 320 SCn 0-310 1

D7 14.12.07 AG 71° 26.15' N, 125° 53.12' W 455 Mn 0-21, 21-40, 41-60, 61-150, 151-250, 251-350, 351-370, 372-390, 391-411 0.5

D12-A 26.12.07 AG 71° 14.96' N, 124° 27.10' W 289 Mn 10-30, 30-50, 50-70, 70-120, 120-170, 170-220, 220-240, 240-260, 260-280 1

D17-9 18.01.08 AG 71° 32.50' N, 125° 0.20' W 216 SCn 0-206 1

D17-I 21.01.08 AG 71° 36.20' N, 125° 9.40' W 242 Mn 10-20, 21-40, 40-60, 60-97, 97-133, 134-171, 172-191, 191-210, 211-233 1

D19-B 04.02.08 AG 71° 4.61' N, 124° 49.02' W 354 Mn 2-21, 22-41, 41-60, 61-125, 126-195, 196-264, 265-286, 286-306, 307-325 0.5

D19-E 10.02.08 AG 71° 3.93' N, 124° 46.15' W 365 Mn 0-20, 20-40, 40-60, 60-130, 130-200, 200-280, 300-320, 320-340 1

D26-12 28.02.08 AG 70° 55.79' N, 123° 51.10' W 359 SCn 0-345 1

D26-E 01.03.08 AG 70° 50.40' N, 123° 36.00' W 443 Mn 60-160 1

D27-B 03.03.08 AG 70° 48.87' N, 123° 53.50' W 352 Mn 4-22, 22-41, 41-62, 62-131, 132-206, 207-280, 280-300, 301-340 0.5

D29-27 19.03.08 AG 70° 54.50' N, 123° 28.60' W 401 SCn 0-385 1

D36-8 08.04.08 AG 71° 17.60' N, 124° 30.60' W 260 SCn 0-250 1

D41-D 18.04.08 AG 70° 37.90' N, 123° 57.90' W 499 Mn 0-20, 20-40, 40-60, 60-185, 185-309, 309-340, 376-404, 404-493 1

D41-E 18.04.08 AG 70° 37.30' N, 121° 55.40' W 500 Mn 0-20, 20-40, 40-60, 60-181, 181-303, 303-425, 424-445, 445-465, 465-485 1

D43-F 01.05.08 AG 70° 48.93 'N, 124° 13.33' W 447 Mn 11-20, 21-40, 40-60, 61-161, 161-263, 263-365, 365-385, 385-405, 405-428 1

405b 20.05.08 AG 70° 39.54' N, 122° 52.74' W 521 Mn 11-19, 20-40, 40-60, 60-179, 179-306, 307-430, 431-449, 450-470, 470-497 0.5

6010 26.05.08 AG 71° 31.86' N, 129° 34.20' W 696 SCn 0-684 1

D45-A 31.05.08 AG 71° 13.07' N, 124° 40.94' W 275 Mn 10-20, 20-40, 40-60, 60-105, 105-150, 150-195, 195-215, 215-235, 235-255 1

405c 02.06.08 AG 70° 37.22 'N, 123° 11.01' W 549 Mn 10-20, 20-40, 40-60, 60-197, 197-333, 433-470, 472-490, 490-510, 510-530 0.5

F7-5 08.06.08 AG 69° 49.45' N, 123° 37.97' W 78 SCn 0-70 1

1208-A 28.06.08 AG 71° 3.84' N, 126° 2.66' W 407 Mn 1-24, 24-40, 40-60, 60-156, 157-245, 246-336, 337-356, 357-372, 372-395 0.5

405 21.07.08 AG 70° 41.70' N, 122° 55.47' W 599 SCn 0-580 1

Page 152: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

130

405 21.07.08 AG 70° 41.92' N, 122° 55.91' W 592 Mn 20-40, 40-60 0.5

CA16-07 23.07.08 AG 71° 47.62' N, 126° 29.13' W 439 Mn 1-21, 21-40, 41-60, 61-161, 162-261, 261-371, 371-391, 392-411, 411-433 1

CA05-07 25.07.08 AG 71° 18.19' N, 127° 34.80' W 231 Mn 1-20, 40-61, 61-82, 82-121, 122-160, 181-202, 202-220 1

CA18-07 03.08.08 AG 70° 41.74' N, 122° 54.94' W 591 SCn 0-580 1

108 14.09.08 BB 76° 15.98' N, 74° 34.85' W 436 SCn 0-436 1

118 17.09.08 BB 77° 18.62' N, 76° 38.66' W 480 Mn 0-20, 40-60, 60-177, 177-294, 294-410, 410-430, 430-450, 450-470 1

Page 153: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

131

Appendix D.

Vertical profiles of temperature, salinity and chlorophyll a (chl a) in Chapter 3: Nachvak Fjord on 2 August 2007; Parry Channel on 7 October

2007; and Baffin Bay on 17 September 2008. Chl a data were provided by Michel Gosselin (Université du Québec à Rimouski).

Page 154: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

132

Appendix E-1. Amundsen Gulf stations sampled in Chapter 4: abbreviated station IDs, full station IDs, dates, positions, bottom depths, sampler used and sample

depths. Properties measured are abbreviated: chl a = chlorophyll a biomass; zoop = zooplankton abundance; gc = chaetognath gut contents; vd =

chaetognath vertical distributions. SCn = square-conical net, Mn = Multinet, CTD = conductivity-temperate-depth device. See Appendix C for the

complete list of Multinet sampling intervals.

Station ID

(our study)

Full

station ID

Date

(dd.mm.yy) Position

Bottom

depth (m) Sampler Sample depths (m) Properties measured

1 1600 02.11.07 71° 40.77' N, 130° 44.21' W 452 SCn 0-358 gc

2 1606 03.11.07 71° 33.47' N, 125° 41.51' W 345 CTD 0, 3, 5, 7, 11, 18, 24 chl a

3 1908 05.11.07 71° 8.64' N, 124° 21.17' W 259 CTD 0, 10, 20, 26, 30, 50, 50, 60, 75, 100 chl a

4 405 19.11.07 70° 37.30' N, 123° 0.09' W 640 CTD 10, 40, 80, 80, 120, 240, 500 chl a

5 1100 19.11.07 71° 2.23' N, 123° 15.11' W 281 CTD 10, 30, 60, 80, 100, 267 chl a, zoop

6 437 22.11.07 71° 44.37' N, 126° 38.85' W 420 CTD 10, 30, 50, 100, 200, 414 chl a

6 437 23.11.07 71° 43.81' N, 126° 41.37' W 440 Mn 3-412 zoop, vd, gc

7 D1 28.11.07 70° 25.85' N, 126° 27.44' W 285 CTD 10, 20, 40, 100, 160, 275 chl a

8 D3 30.11.07 71° 2.39' N, 123° 58.46' W 320 SCn 0-310 zoop, gc

9 D4 02.12.07 71° 43.93' N, 125° 33.80' W 247 CTD 10, 40, 60, 90, 160, 232 chl a

10 D5 05.12.07 71° 24.50' N, 124° 50.09' W 217 Mn 2-205 zoop

11 D5 07.12.07 71° 18.30' N, 124° 47.26' W 239 CTD 10, 50, 70, 110, 170, 232 chl a

12 D7 10.12.07 71° 16.22' N, 125° 13.76' W 323 CTD 10, 40, 80, 120, 190, 315 chl a

13 D7 13.12.07 71° 16.22' N, 125° 13.76' W 323 CTD 10 chl a

14 D7 14.12.07 71° 26.15' N, 125° 53.12' W 455 Mn 0-411 zoop, vd, gc

15 D8 16.12.07 71° 24.84' N, 126° 4.69' W 440 CTD 10, 30, 100, 150, 200, 441 chl a

16 D9 18.12.07 71° 46.22' N, 125° 57.94' W 247 Mn 2-230 zoop

17 D12-A 26.12.07 71° 14.96' N, 124° 27.10' W 289 Mn 10-280 vd, gc

18 D12 27.12.07 71° 12.91' N, 124° 28.63' W 280 CTD 10, 40 chl a

19 D12-B 28.12.07 71° 19.68' N, 124° 49.80' W 222 Mn 40-211 zoop

20 D14 03.01.08 71° 14.93' N, 124° 32.20' W 290 CTD 10, 40 chl a, zoop

21 D14-D 06.01.08 71° 32.60' N, 125° 32.10' W 343 Mn 1-333 zoop

22 D14-F 09.01.08 71° 31.60' N, 125° 42.90' W 360 Mn 10-347 zoop

23 D17 17.01.08 71° 31.53' N, 124° 57.97' W 215 CTD 10, 40 chl a

24 D17-E 18.01.08 71° 32.20' N, 124° 59.70' W 217 Mn 10-208 zoop

25 D17-9 18.01.08 71° 32.50' N, 125° 0.20' W 216 SCn 0-206 gc

26 D17-H 21.01.08 71° 36.20' N, 125° 9.40' W 242 Mn 10-232 zoop

Page 155: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

133

27 D17-I 21.01.08 71° 36.20' N, 125° 9.40' W 242 Mn 10-233 zoop, vd, gc

28 D17-J 21.01.08 71° 36.20' N, 125° 9.40' W 241 Mn 9-231 zoop

29 D17-K 21.01.08 71° 36.20' N, 125° 9.40' W 241 Mn 9-232 zoop

30 D17-L 21.01.08 71° 36.20' N, 125° 9.40' W 241 Mn 10-232 zoop

31 D19 24.01.08 71° 11.12' N, 125° 8.87' W 319 CTD 10, 40 chl a

32 D19 26.01.08 71° 6.65' N, 124° 56.70' W 312 CTD 10, 40 chl a

33 D19-D 27.01.08 71° 5.50' N, 124° 54.00' W 317 Mn 8-309 zoop

34 D19 03.02.08 71° 4.63' N, 124° 49.00' W 354 CTD 12, 30, 40, 160, 250, 300 chl a

35 D19-B 04.02.08 71° 4.61' N, 124° 49.02' W 354 Mn 2-325 zoop, vd

36 D19-D 08.02.08 71° 4.62' N, 124° 49.01' W 357 Mn 2-331 zoop

37 D19-E 10.02.08 71° 3.93' N, 124° 46.15' W 365 Mn 0-340 vd, gc

38 D19 11.02.08 71° 4.24' N, 124° 47.14' W 371 CTD 5, 12, 50, 60, 160, 250, 350 chl a

39 D19-G 12.02.08 71° 8.91' N, 125° 7.53' W 347 Mn 1-316 zoop

40 D19 14.02.08 71° 27.13' N, 126° 26.94' W 456 CTD 12 chl a

41 D22 18.02.08 71° 18.65' N, 124° 29.80' W 261 CTD 12, 20, 50, 63, 121, 201, 246 chl a, zoop

42 D24-A 23.02.08 71° 26.47' N, 125° 39.80' W 418 Mn 0-400 zoop

43 D26 25.02.08 70° 56.04' N, 123° 55.36' W 357 CTD 12, 20, 60, 80, 150, 200, 342 chl a, zoop

44 D26-12 28.02.08 70° 55.79' N, 123° 51.10' W 359 SCn 0-345 gc

45 D26-D 29.02.08 70° 50.97' N, 123° 39.92' W 425 Mn 2-405 zoop

46 D27 01.03.08 70° 47.43' N, 123° 4.16' W 370 CTD 1, 12, 40, 70, 150, 200, 363 chl a

47 D33-4 03.03.08 70° 48.66' N, 123° 54.90' W 390 CTD 0, 2, 5, 10, 25 chl a

48 D27-B 03.03.08 70° 48.87' N, 123° 53.50' W 352 Mn 4-340 zoop, vd, gc

49 D29-A 05.03.08 71° 0.80' N, 123° 23.30' W 236 Mn 1-230 zoop

50 D29-B 05.03.08 71° 2.52' N, 123° 26.34' W 249 Mn 3-230 zoop

51 D29-C 05.03.08 71° 3.61' N, 123° 27.11' W 252 Mn 2-240 zoop

52 D29-D 05.03.08 71° 4.02' N, 123° 27.13' W 242 Mn 2-230 zoop

53 D29-E 06.03.08 71° 3.94' N, 123° 27.00' W 270 Mn 2-230 zoop

54 D29-F 06.03.08 71° 4.30' N, 123° 27.01' W 236 Mn 3-255 zoop

55 D36 06.03.08 71° 3.97' N, 123° 27.05' W 250 CTD 0, 2, 5, 10, 25 chl a

56 D36 09.03.08 71° 1.09' N, 123° 50.69' W 323 CTD 0, 2, 5, 10, 25 chl a

57 D29 10.03.08 71° 2.32' N, 123° 54.65' W 320 CTD 12, 20, 40, 65, 150, 200, 310 chl a, zoop

58 D29 17.03.08 70° 54.48' N, 123° 28.62' W 401 CTD 2, 5, 10, 50 chl a

59 D31 19.03.08 70° 54.73' N, 123° 1.19' W 444 CTD 0, 10, 25 chl a

60 D29-27 19.03.08 70° 54.50' N, 123° 28.60' W 401 SCn 0-385 gc

61 D29M 19.03.08 70° 54.50' N, 123° 28.60' W 401 Mn 1-386 zoop

62 D33C 28.03.08 71° 3.90' N, 121° 47.20' W 187 Mn 5-181 zoop

Page 156: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

134

63 D35A 02.04.08 71° 4.00' N, 121° 56.30' W 213 Mn 0-203 zoop

64 D36C 08.04.08 71° 17.60' N, 124° 30.70' W 259 Mn 0-251 zoop

65 D36-8 08.04.08 71° 17.60' N, 124° 30.60' W 260 SCn 0-250 gc

66 D38 11.04.08 71° 15.52' N, 124° 37.23' W 259 CTD 0, 2, 5, 10, 25 chl a

67 D40 15.04.08 70° 47.89' N, 122° 27.95' W 471 CTD 0, 9, 15, 25, 37, 60, 78, 100 chl a

68 D41 16.04.08 70° 46.65' N, 122° 10.66' W 528 CTD 0, 2, 5, 10, 25 chl a

69 D41C 17.04.08 70° 44.20' N, 122° 8.10' W 541 Mn 2-527 zoop

70 D41-D 18.04.08 70° 37.90' N, 123° 57.90' W 499 Mn 0-493 vd, gc

71 D41-E 18.04.08 70° 37.30' N, 121° 55.40' W 500 Mn 0-485 zoop, vd

72 D41-F 18.04.08 70° 36.50' N, 121° 53.30' W 500 Mn 0-465 zoop

73 D41-G 18.04.08 70° 37.30' N, 121° 51.60' W 514 Mn 0-500 zoop

74 D41 19.04.08 70° 36.52' N, 121° 52.32' W 508 CTD 0, 2, 5, 10, 25 chl a

75 D41-H 19.04.08 70° 35.90' N, 121° 51.00' W 515 Mn 0-500 zoop

76 D41-I 19.04.08 70° 35.90' N, 121° 51.00' W 519 Mn 0-504 zoop

77 D41-J 19.04.08 70° 35.90' N, 121° 51.00' W 519 Mn 0-505 zoop

78 D43-A 23.04.08 70° 36.20' N, 122° 13.00' W 490 Mn 0-475 zoop

79 D43 26.04.08 70° 35.25' N, 122° 26.13' W 550 CTD 0, 2, 5, 10, 25 chl a

80 D43-C 28.04.08 70° 41.33' N, 123° 1.15' W 570 Mn 11-552 zoop

81 D43-1 29.04.08 70° 43.06' N, 123° 17.10' W 459 CTD 0, 2, 5, 10, 25 chl a

82 D43-F 01.05.08 70° 48.93' N, 124° 13.33' W 447 Mn 11-428 zoop, vd, gc

83 D43-2 02.05.08 70° 50.49' N, 124° 26.53' W 475 CTD 0, 2, 5, 10, 25 chl a

84 D43-H 04.05.08 70° 59.99' N, 125° 52.06' W 407 Mn 10-380 zoop

85 D43-3 05.05.08 71° 10.56' N, 126° 18.12' W 420 CTD 0, 2, 5, 10, 25 chl a

86 1020a 06.05.08 71° 1.24' N, 127° 3.10' W 254 CTD 0, 5, 11, 17, 26, 40, 45, 54, 100 chl a

87 6010 06.05.08 72° 31.58' N, 129° 34.50' W 699 Mn 11-680 zoop

88 O2 12.05.08 69° 59.03' N, 126° 2.91' W 205 CTD 11, 18, 26, 40, 54, 75, 100 chl a, zoop

89 405b 19.05.08 70° 39.55' N, 122° 50.74' W 537 CTD 0, 5, 11, 16, 18, 27, 37, 75, 100 chl a

90 405b 20.05.08 70° 39.54' N, 122° 52.74' W 521 Mn 11-497 zoop, vd

91 1011 22.05.08 70° 42.35' N, 124° 0.02' W 455 CTD 0, 6, 13, 21, 30, 32, 51, 69, 75, 100 chl a, zoop

92 6010 26.05.08 71° 31.86' N, 129° 34.20' W 696 SCn 0-684 gc

93 9008 27.05.08 74° 19.91' N, 126° 59.48' W 347 Mn 10-331 zoop

94 D46 30.05.08 71° 13.17' N, 124° 41.11' W 273 CTD 0, 5, 10, 25 chl a

95 D45-A 31.05.08 71° 13.07' N, 124° 40.94' W 275 Mn 10-255 vd, gc

96 405 02.06.08 70° 37.37' N, 123° 11.21' W 546 CTD 0, 10, 14, 22, 35, 53, 73, 100 chl a

97 405c 02.06.08 70° 37.22' N, 123° 11.01' W 549 Mn 10-530 zoop, vd, gc

98 F7-A 07.06.08 69° 49.45' N, 123° 37.97' W 78 Mn 1-78 zoop

Page 157: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

135

99 F7-5 08.06.08 69° 49.45' N, 123° 37.97' W 78 SCn 0-70 gc

100 405B 10.06.08 70° 39.85' N, 123° 0.04' W 562 CTD 0, 7, 12, 19, 30, 36, 46, 60, 75, 100 chl a, zoop

101 FB6 15.06.08 69° 58.95' N, 126° 5.18' W 205 CTD 0, 10, 25 chl a

102 FB6-A 16.06.08 69° 58.84' N, 126° 5.30' W 204 Mn 3-190 zoop

103 1216-A 23.06.08 70° 36.69' N, 127° 25.13' W 236 Mn 1-226 zoop

104 1200-A 28.06.08 71° 32.72' N, 124° 20.30' W 207 Mn 1-197 zoop

105 1208 28.06.08 71° 3.87' N, 126° 4.34' W 400 CTD 0, 7, 12, 18, 22, 34, 45, 60, 75, 100 chl a

106 1208-A 28.06.08 71° 3.84' N, 126° 2.66' W 407 Mn 1-395 zoop, vd

107 421-A 01.07.08 71° 27.95' N, 133° 53.78' W 1104 Mn 1-1080 zoop

108 435-A 02.07.08 71° 4.48' N, 133° 47.54' W 300 Mn 2-290 zoop

109 6006-A 04.07.08 72° 39.47' N, 128° 21.56' W 224 Mn 1-211 zoop

110 2010-A 06.07.08 75° 7.28' N, 120° 22.89' W 405 Mn 1-405 zoop

111 410 08.07.08 71° 42.34' N, 126° 29.26' W 395 CTD 0, 12, 29, 33, 54, 75, 80, 100, 108, 200, 388 chl a

112 410-A 08.07.08 71° 41.96' N, 126° 29.18' W 404 Mn 1-392 zoop

113 416 10.07.08 71° 42.16' N, 126° 7.42' W 236 CTD 0, 12, 21, 33, 52, 75, 78, 80, 100, 108 chl a

114 1100 11.07.08 71° 2.66' N, 123° 16.36' W 265 CTD 0, 7, 12, 18, 29, 46, 60, 75, 100, 200 chl a

115 1100-A 11.07.08 71° 2.58' N, 123° 15.58' W 270 Mn 2-261 zoop

116 403-10A 20.07.08 70° 5.46' N, 120° 8.90' W 410 CTD 5, 8, 14, 28, 35, 54, 73, 100, 200, 400 chl a

117 405-10A 21.07.08 70° 41.15' N, 122° 54.08' W 584 CTD 4, 9, 16, 26, 37, 41, 62, 75, 84, 100, 200, 400 chl a

118 405 21.07.08 70° 41.70' N, 122° 55.47' W 599 SCn 0-580 zoop, gc

119 437 23.07.08 71° 41.64' N, 126° 36.01' W 441 CTD 3, 11, 18, 29, 46, 55, 70, 75, 95 chl a

120 CA16-07 23.07.08 71° 47.62' N, 126° 29.13' W 439 Mn 1-433 zoop, vd, gc

121 408 25.07.08 71° 18.83' N, 127° 36.15' W 200 CTD 0, 9, 15, 25, 39, 55, 59, 75, 80, 100, 192 chl a

122 CA05-07 25.07.08 71° 18.19' N, 127° 34.80' W 231 Mn 1-220 zoop, vd

123 CA08-07 27.07.08 71° 8.73' N, 126° 8.12' W 402 Mn 1-393 zoop

124 CA04-07 30.07.08 71° 5.57' N, 133° 45.24' W 335 Mn 1-327 zoop

125 2008-10A 02.08.08 71° 19.95' N, 126° 13.09' W 442 CTD 5, 8, 13, 21, 51, 70, 75, 100, 200 chl a

126 405-10A 03.08.08 70° 40.03' N, 122° 59.01' W 536 CTD 0, 5, 8, 13, 21, 32, 43, 75, 100, 200 chl a

127 CA18-07 03.08.08 70° 41.74' N, 122° 54.94' W 591 SCn 0-580 gc

Page 158: Ecology and physiology of chaetognaths (semi-gelatinous ... · Les taux de prédation quotidiens évalués à partir des analyses du contenu du tube digestif sont généralement restés

136

Appendix E-2.

Chukchi Sea and Baffin Bay stations sampled in Chapter 4: abbreviated station IDs (see

Figure 4.1b and Figure 4.1c), full station IDs, dates, positions, bottom depths, sampler

used and sampling depths. For square-conical net sampling, selected mesh size and tow

type are also shown (O = oblique tow, V = vertical tow). FA = fatty acids, SI = stable

isotopes. *Analysis performed on Eukrohnia hamata, +analysis performed on Parasagitta

elegans (see ‘Method’).

Station ID

(our study)

Full

station ID Date Position

Bottom

depth (m)

Sampler

(SCn mesh, tow type)

Sampling

depths (m)

Parameters

measured

Chukchi Sea

CS1 1034 13.09.14 71° 54' N, 154° 58' W 379 SCn (750 µm, O) 0-90 FA+, SI+

CS2 1030 14.09.14 72° 13' N, 153° 58' W 2081 SCn (750 µm, O) 0-90 SI+

CS3 1085 16.09.14 75° 3' N, 167° 8' W 254 SCn (750 µm, O) 0-90 FA+, SI+

CS4 1100 18.09.14 75° 4' N, 161° 16' W 1985 SCn (750 µm, O) 0-90 FA+, SI+

CS4 1100 18.09.14 75° 4' N, 161° 16' W 1985 SCn (200 µm, V) 0-1972 FA*, SI*

CS6 1115 20.09.14 73° 54' N, 147° 11' W 3773 SCn (200 µm, V) 0-999 SI*

CS7 1130 22.09.14 72° 36'’ N, 144° 44' W 3234 SCn (200 µm, V) 0-1000 SI*

Scott Inlet Fjord

SIF PCBC-2 01.10.14 71° 5' N, 71° 50' W 693 SCn (500 µm, O) 0-90 FA+, SI+

SIF PCBC-2 01.10.14 71° 5' N, 71° 50' W 693 SCn (500 µm, V) 0-685 FA*, SI*

Gibbs Fjord

GF Gibbs-B 01.10.14 70° 46' N, 72° 14' W 440 SCn (500 µm, V) 0-430 FA*, SI*

Southern Baffin Bay

BB 180 03.10.14 67° 28' N, 61° 42' W 214 SCn (500 µm, O) 0-90 FA*+, SI*+